412
Investigation into mechanisms of biofilm formation by Klebsiella pneumoniae Thesis submitted by Nur Asyura Binti Nor Amdan For the degree of DOCTOR OF PHILOSOPHY in the Faculty of Medical Sciences University College London Department of Microbial Diseases UCL Eastman Dental Institute 256 Gray’s Inn Road London WC1X 8LD UK 2018

Investigation into mechanisms of biofilm formation by

Embed Size (px)

Citation preview

i |

Investigation into mechanisms

of biofilm formation by

Klebsiella pneumoniae

Thesis submitted by

Nur Asyura Binti Nor Amdan

For the degree of

DOCTOR OF PHILOSOPHY

in the

Faculty of Medical Sciences

University College London

Department of Microbial Diseases UCL Eastman Dental Institute

256 Gray’s Inn Road London WC1X 8LD

UK

2018

ii |

For mama and abah..

Verily, with every difficulty, there is relief (94:5)

iii |

Declaration

I hereby certify that the work embodied in this thesis is the result of my own

investigation except where otherwise stated.

iv |

Abstract

Klebsiella pneumoniae is a Gram-negative opportunistic pathogen that

normally causes nosocomial infections such as urinary tract infections,

septicaemia, and respiratory tract infections. K. pneumoniae has become

progressively resistant to the vast majority of antibiotics, and treatment of

these particular infections is very challenging. At present, the emergence of

antibiotic resistance is a widespread phenomenon that has been established

as a major global healthcare threat. The ability of K. pneumoniae to form

biofilms is known as one of the most important factors that contributes to the

spread of this antibiotic-resistant bacterium. Due to that fact, understanding

the mechanisms behind biofilm formation by K. pneumoniae is of the utmost

importance.

Numerous studies have shown that sub-inhibitory concentrations of antibiotics

could induce biofilm formation by clinically important pathogens, for example,

Escherichia coli, Pseudomonas aeruginosa, and Staphylococcus aureus. In

this study, we report for the first time on the effect of sub-minimum inhibitory

concentrations (sub-MICs) of gentamicin and ciprofloxacin on biofilm formation

by K. pneumoniae.

In addition, by using biofilm inhibition and disruption assays, we have also

assessed the role of proteins in formation and composition of K. pneumoniae

biofilms upon cultivation in nutritious and nutrient-poor media. As well as

proteins, the role of polysaccharides and extracellular DNA was also

determined.

v |

Furthermore, the role of type 1 fimbriae (T1F) in biofilm formation by

K. pneumoniae was also determined. Transcription of fimA (the major subunit

of T1F) is phase variable, as its promoter is on an invertible DNA element, fim-

switch (fimS). We investigated the orientation of fimS in all K. pneumoniae

strains used in this study, in planktonic and as biofilm cells, upon cultivation in

tryptic soy broth (TSB) and artificial urine medium (AUM).

Finally, we have also determined the role of outer membrane porins, OmpK35

and OmpK36 in biofilm formation by K. pneumoniae. We developed isogenic

mutants of ∆ompK35 and ∆ompK36 strains. The capacity to form biofilms by

the mutant strains compared to parental strains was examined in two different

models, a microtitre plate model and a model using catheter pieces.

Currently, we still lack a full understanding on the mechanisms of biofilm

formation by K. pneumoniae which is important in order to develop new

treatments to reduce the threat of biofilm-mediated infections by

K. pneumoniae.

vi |

Impact statement

An understanding of the mechanisms of biofilm formation by Klebsiella

pneumoniae is very important since this pathogen is classified as one of the

most prevalent bacterial species that cause nosocomial infections. Having the

capacity to form biofilms is among the contributing factors that result in

tolerance to antibiotics by K. pneumoniae.

The findings in this study have shown that exposure of K. pneumoniae to sub-

minimum inhibitory concentrations (sub-MICs) of commonly used antibiotics:

gentamicin and ciprofloxacin in three different conditions, did not induce biofilm

formation by K. pneumoniae. This contrasts with several studies which have

shown that antibiotics at sub-inhibitory concentrations could induce biofilm

formation by other clinically important pathogens such as Escherichia coli and

Pseudomonas aeruginosa.

Our data also suggest that proteins could play an important role in the

composition of extracellular polymeric substances (EPS) and biofilm formation

by K. pneumoniae. This data provides important insights on biofilm formation

by K. pneumoniae since very little is known with regard to the composition of

biofilms of K. pneumoniae.

We also investigated the orientation of the fim-switch (fimS) which controls

expression of type 1 fimbriae (T1F) in biofilm formation by K. pneumoniae.

The orientation of fimS was determined in planktonic and biofilm cells upon

cultivation of seven examined K. pneumoniae strains in two different media.

vii |

This data provide further knowledge on the orientation of fimS which controls

the expression of T1F in nutritious and nutrient-poor media.

Due to the fact that K. pneumoniae contributes to a high number of cases of

biofilm-associated infections such as catheter-associated urinary tract

infections, we have provided further knowledge on the causing factors to this

problem. The data in this study has also shown that OmpK35 and OmpK36

do not contribute to biofilm formation by K. pneumoniae upon growth on

clinically relevant surfaces such as silicone catheter surfaces.

Currently, commonly prescribed antibiotics for antibiotic resistant

K. pneumoniae infections are not very effective at eradicating these kinds of

infections. Thus, new approaches such as targeting the contributing factors

leading to biofilm formation by K. pneumoniae are important. This study

provides insight on biofilm formation by K. pneumoniae that could be useful for

future investigation.

viii |

Acknowledgements

My utmost gratitude and appreciation goes to my supervisors, Dr Sean P. Nair

and Dr Adam P. Roberts who have guided and supported me throughout my

research and dissertation. Their constant encouragement and great

supervision were truly inspiring.

This dissertation would not have been possible without the steadfast

encouragement and moral support from my beloved family especially my dad,

my mum, Along, Kak Yati, Kakak Alia, and Adik Aida. Thank you for always

being there for me.

I want to especially express my appreciation to my close friends (you know

who you are), my travel buddies (Najiah, Ayna, Zawa, and Hanim) and these

four ‘culprits’ (Supathep, Shirene, Supanan, and Sophia) whose

companionship has helped me tremendously in this journey.

I am indebted to my gracious and helpful laboratory members from Department

of Microbial Diseases, especially Dr Haitham Hussain, Dr Anna Tymon, and

also to Vanessa, Franky, Liam, Ajijur, Enas, Marika, Zeina, Ladan, Mehmet,

Deena, Hadeel, Shatha, Ingrid, Tracey, Arely, Khadija, Erni, Catie, Sarah,

Andre, Dallas, Ahmed, Carolina, and all my colleagues whose support,

assistance and opinion have enabled me to develop and undertake my

research as planned.

I would also like to express my gratitude to all Malaysian friends in London and

in the UK (Asrah, Shikin, Mazlina, Ashraf, Irina, Nisaa, Juzaili, Awis, Muzamir,

Shahrul, Hazeeq, Shafiq, Azizi, Ehsan, Iqbal, Siti, Kak Kurshiah, Faten,

ix |

Fadilah, Sudin, Firdaus, Asrar, Amal, and Afzal) for their unwavering mental

support and assistance.

Last but not least, it is with much honour I would like to thank the Government

of Malaysia for providing me the financial support under the Majlis Amanah

Rakyat (MARA) Sponsorship Award.

x |

Table of contents

Declaration .................................................................................................. iii

Abstract ....................................................................................................... iv

Impact statement ........................................................................................ vi

Acknowledgements .................................................................................. viii

Table of contents ......................................................................................... x

List of figures .......................................................................................... xviii

List of tables ............................................................................................. xxii

List of abbreviations ............................................................................... xxiv

1.0 General introduction ....................................................................... 1

1.1 Klebsiella genus ................................................................................ 4

1.1.1 Klebsiella pneumoniae ...................................................................... 5

1.2. Antibiotherapy for K. pneumoniae infection....................................... 6

1.3 Clinical manifestation of infections caused by K. pneumoniae ......... 8

1.4 Mechanisms of antibiotic resistance ............................................... 10

1.4.1 Extended spectrum β-lactamase (ESBL) and carbapenemase-

producing K. pneumoniae ............................................................... 14

1.4.2 Evolution of β-lactamases in K. pneumoniae .................................. 17

1.4.3 Epidemiology of multidrug-resistant K. pneumoniae ....................... 19

1.5 Mechanisms of K. pneumoniae pathogenicity ................................. 25

1.5.1 The capsule .................................................................................... 27

1.5.2 Lipopolysaccharides (LPS) ............................................................. 28

1.5.3 Siderophores ................................................................................... 29

1.5.4 Fimbriae .......................................................................................... 30

1.5.5 Outer membrane porins (Omps) ..................................................... 31

xi |

1.6 Biofilm formation by K. pneumoniae ............................................... 34

1.7 Mechanisms of antibiotic resistance in bacterial biofilms ................ 41

1.8 Aims and hypothesis of this study ................................................... 43

2.0 Materials and Methods ................................................................. 46

2.1 Sources of media, enzymes and reagent ........................................ 46

2.2 Bacterial strains and plasmids ........................................................ 47

2.3 Growth condition and storage of bacteria........................................ 56

2.4 Molecular biology methods ............................................................. 56

2.4.1 Genomic DNA and plasmid extraction ............................................ 56

2.4.2 Oligo synthesis ................................................................................ 57

2.4.3 Polymerase chain reaction (PCR) amplification .............................. 61

2.4.4 Agarose gel electrophoresis ............................................................ 61

2.4.5 PCR purification and DNA purification from gels ............................. 62

2.4.6 DNA sequencing ............................................................................. 62

2.4.7 In silico sequence analysis and protein structural analysis ............. 63

2.4.8 Restriction endonuclease reactions ................................................ 65

2.4.9 DNA ligation reaction ...................................................................... 65

2.4.10 Dephosphorylation reactions ........................................................... 65

2.4.11 Transformation of plasmids into E. coli -select silver efficiency

competent cells ............................................................................... 66

2.4.12 Blue/white screening ....................................................................... 66

2.5 Statistical analysis ........................................................................... 67

3.0 Investigation into the effect of antibiotic exposure on

Klebsiella pneumoniae biofilm formation ................................................ 68

3.1 Introduction ..................................................................................... 69

3.1.1 In vitro biofilm models ..................................................................... 69

xii |

3.1.2 The effects of antibiotic exposure on biofilm formation ................... 70

3.1.3 Justification of the experimental design .......................................... 73

3.1.4 Modes of action and mechanisms of resistance of antibiotics ......... 74

3.1.4.1 Gentamicin ............................................................................. 74

3.1.4.2 Ciprofloxacin ........................................................................... 77

3.1.5 Aims of this chapter ........................................................................ 79

3.2 Materials and methods .................................................................... 80

3.2.1 Biofilm formation under different growth conditions and in different

media .............................................................................................. 80

3.2.2 Pre-exposure of cultures to antibiotics prior to biofilm formation ..... 80

3.2.3 Exposure of mid-log phase cultures to antibiotics during biofilm

formation ......................................................................................... 81

3.2.4 Exposure of mature biofilms to antibiotics ....................................... 82

3.2.5 Crystal violet quantification of biofilm formation .............................. 82

3.3 Results ............................................................................................ 84

3.3.1 Optimisation of growth conditions and growth media for biofilm

formation assays ............................................................................. 84

3.3.2 Pre-exposure of cultures to gentamicin or ciprofloxacin prior to biofilm

formation ......................................................................................... 89

3.3.3 Exposure of mid-log phase cultures to gentamicin or ciprofloxacin

during 24 h biofilm formation ........................................................... 91

3.3.4 Exposure of mature biofilms to antibiotics ....................................... 93

3.4 Discussion ....................................................................................... 97

3.5 Conclusion .................................................................................... 102

4.0 Investigation into extracellular polymeric substances (EPS) of

Klebsiella pneumoniae biofilms ............................................................. 104

xiii |

4.1 Introduction ................................................................................... 104

4.1.1 The biofilm matrix .......................................................................... 104

4.1.1.1 Proteins ................................................................................ 105

4.1.1.2 Extracellular DNA (eDNA) .................................................... 105

4.1.1.3 Polysaccharides ................................................................... 106

4.1.1.3.1 Cellulose ............................................................................... 106

4.1.1.3.2 Poly-N-acetylglucosamine (PNAG) ....................................... 107

4.1.2 Extracellular polymeric substances (EPS) produced by

K. pneumoniae during biofilm formation ........................................ 108

4.1.2.1 Biofilm EPS of K. pneumoniae upon cultivation in nutrient-rich

media .................................................................................... 108

4.1.3 Artificial urine medium (AUM) ....................................................... 109

4.1.4. Aims of this chapter ...................................................................... 110

4.2 Materials and methods .................................................................. 111

4.2.1 Bioinformatic analysis ................................................................... 111

4.2.2 Media ............................................................................................ 111

4.2.3 Growth kinetics of examined K. pneumoniae strains in

TSB and AUM ............................................................................... 114

4.2.3.1 Determination of relationship between absorbance (OD600) and

colony forming unit per ml (CFU/ml) ..................................... 114

4.2.3.2 Determination of bacterial growth rate .................................. 115

4.2.4 Biofilm assays in TSB and AUM ................................................... 117

4.2.5 Biofilm inhibition and disruption assays ........................................ 117

4.2.5.1 Biofilm inhibition assays ....................................................... 117

4.2.5.2 Biofilm disruption assays ...................................................... 118

4.3 Results .......................................................................................... 119

4.3.1 Bioinformatic analysis ................................................................... 119

xiv |

4.3.1.1 Comparative analysis of bacterial cellulose synthase (bcs)

operon .................................................................................. 119

4.3.1.2 Comparative analysis of pga operon .................................... 128

4.3.2 Determination of the relationship between absorbance and colony

forming unit per ml (CFU/ml) ......................................................... 134

4.3.3 Growth kinetics of K. pneumoniae strains in TSB and AUM ......... 138

4.3.4 Quantification of K. pneumoniae biofilm upon cultivation in

TSB and AUM ............................................................................... 143

4.3.5 Biofilm inhibition and disruption assays ........................................ 145

4.3.5.1 Effect of 100 µg/ml proteinase K on biofilm formation and on

24 h pre-formed biofilms ....................................................... 146

4.3.5.2 Effect of 100 µg/ml of DNase on biofilm formation and on

24 h pre-formed biofilms ....................................................... 149

4.3.5.3 Effect of NaIO4 on 24 h pre-formed biofilms ......................... 152

4.3.5.4 Effect of cellulase on biofilm formation and on

24 h pre-formed biofilms ....................................................... 154

4.4 Discussion ..................................................................................... 157

4.5 Conclusion .................................................................................... 168

5.0 Investigation into the orientation of the Klebsiella pneumoniae

fim-switch which controls expression of type 1 fimbriae..................... 170

5.1 Introduction ................................................................................... 170

5.1.1 Major fimbriae of K. pneumoniae ...................................................... 170

5.1.1.1 Type 1 fimbriae (T1F) ........................................................... 173

5.1.1.1.1 Fim-switch (fimS) orientation ................................................ 177

5.1.1.2 Type 3 fimbriae (T3F) ........................................................... 179

5.1.2 Role of type 1 fimbriae (T1F) in biofilm formation by

K. pneumoniae .............................................................................. 181

xv |

5.2 Materials and Methods .................................................................. 187

5.2.1 DNA isolation from planktonic and biofilm cells ............................. 187

5.2.2 Determination of the invertible element (fimS) orientation............. 188

5.2.3 Bioinformatic analysis of the T1F operon ...................................... 188

5.3 Results .......................................................................................... 189

5.3.1 Determination of fimS orientation .................................................. 189

5.3.2 Analysis of the T1F operon sequences ......................................... 192

5.4 Discussion ..................................................................................... 197

5.5 Conclusion .................................................................................... 203

6.0 Role of OmpK35 and OmpK36 porins in biofilm formation by

Klebsiella pneumoniae ............................................................................ 205

6.1 Introduction ................................................................................... 205

6.1.1 K. pneumoniae outer membrane porins (Omps) ........................... 206

6.1.2 Major non-specific porins of K. pneumoniae ................................. 207

6.1.2.1 OmpK35 ............................................................................... 207

6.1.2.2 OmpK36 ............................................................................... 208

6.1.2.3 Structure of non-specific porins ............................................ 210

6.1.3 Role of outer membrane porins (Omps) in biofilm formation ......... 212

6.1.4 Regulation of OmpK35 and OmpK36 expression ........................ 217

6.1.4.1 EnvZ-OmpR two-component system .................................... 217

6.1.5 K. pneumoniae and urinary tract infections ................................... 219

6.1.5.1 Catheter-associated urinary tract infections (CAUTIs) .......... 220

6.1.5.2 Biofilms on catheters ............................................................ 220

6.1.6 Aims of this chapter ...................................................................... 221

6.2 Materials and methods .................................................................. 222

6.2.1 Bacterial strains and plasmid ........................................................ 222

xvi |

6.2.2 Determination of genetic context of ompK35 and ompK36 of

K. pneumoniae .............................................................................. 222

6.2.3 Analysis of protein structure .......................................................... 223

6.2.4 Development of ompK35 and ompK36 mutant strains .................. 225

6.2.4.1 Screening of ompK35 and ompK36 ...................................... 225

6.2.4.2 Construction of mutant constructs using SOE PCR .............. 225

6.2.4.3 Construction of new mutant cassettes using aad9 as a selective

marker .................................................................................. 230

6.2.4.4 Modification of the pKD46 helper plasmid by replacing bla with

aac(3)-IV ............................................................................... 233

6.2.4.5 Preparation of electrocompetent K. pneumoniae cells ....... 236

6.2.4.6 Electroporation of pKD46-apra into electrocompetent

K. pneumoniae cells ............................................................. 237

6.2.4.7 Preparation of electrocompetent K. pneumoniae harbouring

pKD46-apra cells .................................................................. 238

6.2.4.8 Electroporation of ompK35 and ompK36 mutant cassettes into

electrocompetent K. pneumoniae harbouring

pKD46-apra cells .................................................................. 239

6.2.5 Complementation of ompK35 and ompK36 strains .................. 241

6.2.5.1 Preparation of ompK35 and ompK36 complementation

constructs ............................................................................. 241

6.2.6 Epsilometer test (E-test) method ................................................... 243

6.2.7 Determination of MIC using agar and broth dilution methods ....... 243

6.2.8 Growth kinetics of wild type and mutant strains in TSB and AUM . 244

6.2.9 Biofilm formation assays in microtitre plate ................................... 244

6.2.10 Biofilm formation assays on catheter pieces ................................. 245

6.2.10.1 Staining and enumeration of biofilm cells on catheter pieces 247

6.3 Results .......................................................................................... 249

xvii |

6.3.1 Determination of genetic context of ompK35 and ompK36 ........... 249

6.3.2 Analysis of OmpC, OmpK35 and OmpK36 protein structure ........ 258

6.3.3 Amplification of ompK35 and ompK36 .......................................... 261

6.3.4 Development of mutant constructs ................................................ 263

6.3.5 Development of ΔompK35 or ΔompK36 strains ............................ 265

6.3.6 Development of complemented strains ........................................ 271

6.3.6.1 Construction of ompK35 complementation construct harbouring

tet(M) .................................................................................... 271

6.3.7 Antimicrobial susceptibility testing ................................................. 274

6.3.8 Growth kinetics of wild type and mutant strains upon cultivation in

TSB and AUM ............................................................................... 277

6.3.9 Role of OmpK35 and OmpK36 in biofilm formation by K. pneumoniae

in 96-well polystyrene tissue culture treated plates ....................... 284

6.3.10 Role of OmpK35 and OmpK36 in biofilm formation by K. pneumoniae

on catheter pieces ......................................................................... 286

6.4 Discussion ..................................................................................... 289

6.5 Conclusion .................................................................................... 300

7.0 General discussion and future directions ................................ 303

7.1 General discussion ....................................................................... 303

7.2 Future directions ........................................................................... 309

8.0 References ................................................................................... 313

Appendices............................................................................................... 369

xviii |

List of figures

Figure 1-1 Mechanisms of antibiotic resistance in Gram-negative bacteria. .. 2

Figure 1-2 Hypermucoviscous morphology of K. pneumoniae strain

Kpneu_110. ................................................................................... 5

Figure 1-3 Chemical structure of ampicillin. β- lactam ring is circled

in black. ....................................................................................... 14

Figure 1-4 The resistance movement of carbapenem-resistant

Enterobacteriaceae. .................................................................... 20

Figure 1-5 Surveillance atlas of the increasing number of K. pneumoniae

antibiotic resistance cases starting from A) 2005 to B) 2016

in Europe. .................................................................................... 22

Figure 1-6 Virulence factors of K. pneumoniae. ........................................... 26

Figure 1-7 Gram-positive and Gram-negative cell walls. ............................. 32

Figure 3-1 The modes of action of gentamicin. ............................................ 74

Figure 3-2 Chemical structure of gentamicin. .............................................. 76

Figure 3-3 Chemical structure of ciprofloxacin. ............................................ 77

Figure 3-4 Simplified diagram of methodology used in this chapter. ............ 83

Figure 3-5 Biofilm formation by K. pneumoniae under different growth

conditions. ................................................................................... 86

Figure 3-6 Biofilm formation by K. pneumoniae in two different media. ....... 88

Figure 3-7 Effect of pre-exposure of cultures to gentamicin or ciprofloxacin

prior to biofilm formation by K. pneumoniae. ............................... 90

Figure 3-8 Effect of antibiotic exposure during 24 h incubation on biofilm

formation by K. pneumoniae. ...................................................... 92

Figure 3-9 Effect of antibiotic exposure on growth of K. pneumoniae. ......... 94

xix |

Figure 3-10 Effect of antibiotic exposure on mature biofilms of

K. pneumoniae. ........................................................................... 96

Figure 4-1 Schematic presentation of bcs operons of K. pneumoniae. ...... 122

Figure 4-2 Partial multiple sequence alignment of bcsO and BcsO. .......... 123

Figure 4-3 Partial multiple sequence alignment of bcsQ and BcsQ. .......... 124

Figure 4-4 Partial multiple sequence alignment of bcsB_I and BcsB_I. ..... 125

Figure 4-5 Partial multiple sequence alignment of bcsC_I and BcsC_I. ..... 126

Figure 4-6 Schematic presentation of pgaABCD operon. .......................... 129

Figure 4-7 Pairwise sequence alignment of Pga amino acid. .................... 133

Figure 4-8 The relationship between absorbance and CFU/ml of

K.pneumoniae upon cultivation in TSB. .................................... 136

Figure 4-9 Growth of K. pneumoniae in two different growth media. ......... 141

Figure 4-10 Biofilm formation by K. pneumoniae in TSB and AUM............ 144

Figure 4-11 Biofilm inhibition and disruption assays in the presence of

100 µg/ml proteinase K. ............................................................ 148

Figure 4-12 Biofilm inhibition and disruption assays in the presence of

100 µg/ml DNase. ..................................................................... 151

Figure 4-13 Biofilm disruption assays in the presence of 10 mM NaIO4. ... 153

Figure 4-14 Biofilm inhibition and disruption assays in the presence of

0.1 % cellulase. ......................................................................... 156

Figure 4-15 Periodate oxidation of cellulose monomer. ............................. 165

Figure 5-1 The chaperon-usher (CU) pathway of T1F biogenesis. ............ 172

Figure 5-2 The structure of T1F mannose-sensitive fimbriae. .................... 174

Figure 5-3 Schematic diagram of T1F operon............................................ 176

Figure 5-4 Schematic diagram of fimE, fimS and fimA. .............................. 178

xx |

Figure 5-5 The structure of T3F mannose-resistant fimbriae. .................... 180

Figure 5-6 Determination of fimS orientation in biofilm cells. ..................... 191

Figure 5-7 Multiple sequence alignment of FimK. ...................................... 196

Figure 6-1 The structure of nonspecific porin. ............................................ 211

Figure 6-2 Regulation of the porin genes by EnvZ/OmpR system. ............ 218

Figure 6-3 Schematic diagram of development of mutant constructs using

ompK35 as an example. ........................................................... 228

Figure 6-4 Schematic diagram of erm1(B) replacement with aad9 in ompK35

cassette as an example. ........................................................... 232

Figure 6-5 Schematic diagram of the replacement of bla with aac(3)-IV

in pKD46. .................................................................................. 235

Figure 6-6 Flow diagram of the gene replacement by double cross over

homologous recombination in K. pneumoniae. ......................... 240

Figure 6-7 Construction of pUC19-tet(M)_ompK35 complementation

construct. .................................................................................. 242

Figure 6-8 Biofilm formation assay on catheter pieces .............................. 246

Figure 6-9 Genetic context of ompK35 and ompK36. ................................ 249

Figure 6-10 Multiple sequence alignment of upstream and downstream

regions of ompK35. ................................................................... 253

Figure 6-11 The upstream region of ompK36 with a putative promoter site

of rcsD. ...................................................................................... 254

Figure 6-12 Multiple sequence alignment of upstream and downstream

regions of ompK36. ................................................................... 257

Figure 6-13 Amino acid multiple sequence alignment of OmpC, OmpK36 and

OmpK35. ................................................................................... 260

xxi |

Figure 6-14 Screening for ompK35 and ompK36. ...................................... 262

Figure 6-15 Amplification of ompK35 and ompK36 mutant cassettes. ....... 264

Figure 6-16 Location of the external primers for confirmation of mutant

strains. ...................................................................................... 268

Figure 6-17 Confirmation of mutant strains by PCR. .................................. 270

Figure 6-18 Construction of ompK35 complementation constructs. ........... 272

Figure 6-19 Confirmation of the complementation constructs by digestion. 273

Figure 6-20 Growth curve of Kpneu_1, C3091, and isogenic mutants

in TSB. ...................................................................................... 279

Figure 6-21 Growth curve of Kpneu_1, C3091, and isogenic mutants

in AUM. ..................................................................................... 281

Figure 6-22 Biofilm formation of Kpneu_1, C3091, and their isogenic mutants

in TSB and AUM. ...................................................................... 285

Figure 6-23 Biofilm formation and cell viability of Kpneu_1 and its isogenic

mutants on silicone catheter surfaces. ...................................... 287

Figure 6-24 Biofilm formation and cell viability of C3091 and its isogenic

mutants on silicone catheter surfaces. ...................................... 288

xxii |

List of tables

Table 1-1 Classification schemes for bacterial β-lactamases. ..................... 16

Table 1-2 Involvement of capsule in biofilm formation by several

K. pneumoniae strains. ............................................................... 37

Table 2-1 Bacterial strains used in this study. .............................................. 47

Table 2-2 Antibiotic susceptibility profile of clinical isolates of K. pneumoniae

used in this study. ....................................................................... 50

Table 2-3 Determination of MIC of antibiotics for K. pneumoniae strains used

in this study using E-test. ............................................................ 51

Table 2-4 Distribution of antibiotic resistance genes in isolates used in this

study. .......................................................................................... 52

Table 2-5 Distribution of virulence clusters in isolates used in this study. .... 53

Table 2-6 Plasmids used in this study .......................................................... 54

Table 2-7 Primers used throughout this study. ............................................ 57

Table 2-8 Bioinformatic programmes used in this study. ............................. 63

Table 4-1 Media used in this study ............................................................ 112

Table 4-2 Amino acid variation in proteins encoded by the bcs operon. .... 127

Table 4-3 Amino acid variation in proteins encoded by pga operon........... 129

Table 4-4 Growth rate constant (µ) and doubling time (td) of K. pneumoniae

strains upon cultivation in TSB using OD600 and CFU/ml as

parameters ................................................................................ 137

Table 4-5 Growth rate constant (µ) and doubling time (td) of K. pneumoniae

strains upon cultivation in TSB and AUM. ................................. 142

Table 5-1 Involvement of fimbriae in biofilm formation by

Enterobacteriaceae. .................................................................. 183

xxiii |

Table 5-2 Orientation of fimS in planktonic and biofilm cells of

K. pneumoniae. ......................................................................... 190

Table 5-3 Amino acid substitutions identified in FimF, FimI, and FimK. ..... 194

Table 6-1 K. pneumoniae porins ................................................................ 207

Table 6-2 Involvement of Omps in initial adhesion or biofilm formation on

abiotic surfaces by Enterobacteriaceae. ................................... 215

Table 6-3 The MIC value of spectinomycin and apramycin against

K. pneumoniae. ......................................................................... 266

Table 6-4 Median MIC of antibiotics and antiseptics for Kpneu_1, C3091, and

their isogenic mutants. .............................................................. 276

Table 6-5 Growth rate constant (µ) and doubling time (td) of Kpneu_1 and

C3091, and their isogenic mutant strains upon cultivation in

TSB and AUM. .......................................................................... 283

xxiv |

List of abbreviations

α Alpha

β Beta

γ Gamma

λ Lambda

°C Degree Celsius

Å Ångström

aa Amino acid

AUM Artificial urine medium

BHI Brain heart infusion

BLAST Basic Local Alignment Search Tool

bp Base pair

CAUTI Catheter-associated urinary tract infection

c-di-GMP Cyclic dimeric guanosine monophosphate

CIAP Calf Intestinal Alkaline Phosphatase

cm Centimeter

CPC Cetylpyridium chloride

CTAB Cetyltrimethylammonium bromide

CFU Colony forming unit

CV Crystal violet

DNA Deoxyribonucleic acid

eDNA Extracellular DNA

EPS Extracellular polymeric substances

ESBL-E ESBL-producing Enterobacteriaceae

fimS Fim-switch

g-force (rcf) Gravitational force (relative centrifugal force)

h Hour

kb Kilobase

LB Luria-Bertani

LPS Lipopolysaccharide

µl Microlitre

µg Microgram

µM Micromolar

xxv |

µF Microfarad

ml Millimetre

mM Millimolar

min Minute

MQ H2O Milli-Q water

nM Nanometer

Ω Ohm

OD Optical density

PBS Phosphate buffered saline

PCR Polymerase chain reaction

PNAG Poly-N-acetylglucosamine

RBS Ribosome binding site

rpm Revolutions per minute

sec Second

SOC Super optimal broth with catabolite repression

SOE Splicing by overlap extension

T1F Type 1 fimbriae

T3F Type 3 fimbriae

TAE Tris acetate EDTA

TSB Tryptic soy broth

UTI Urinary tract infection

X-Gal 5-bromo-4-chloro-3-indolyl-β-D-galactopyranoside

V Volt

v/v volume/volume

v/w volume/weight

1 |

Chapter 1. 0

General introduction

2 |

1.0 General introduction

Antibiotic resistance is a widespread phenomenon that has become one of the

major healthcare problems faced by the world today. It is an increasingly

serious threat to global public health and requires attention from governments,

healthcare sectors and the public. Antibiotic resistance arises when bacteria

are able to grow in inhibitory concentrations of antibiotics. Resistance is due

to the modification of the antibiotic target sites by genetic mutations,

modification of the antibiotic molecules by antibiotic-modifying enzymes,

decreased membrane permeability to antibiotics, existence of efflux pumps,

overproduction of the antibiotic target, protection of the ribosome by ribosomal

protection protein, or acquisition of resistance genes from other resistant

bacteria that facilitate a population of bacteria to survive in the environment, in

the presence of antibiotics (Figure 1-1) (Allen, et al. , 2010; Connell, et al. ,

2003; Munita, et al. , 2016; Schmieder, et al. , 2012).

Figure 1-1 Mechanisms of antibiotic resistance in Gram-negative bacteria.

3 |

Several examples of mechanisms of antibiotic resistance in Gram-negative bacteria

include a) impermeability barrier (e.g. preventing erythromycin uptake), b) multidrug

resistance efflux pumps (e.g. for tetracyclines), c) resistance mutations (e.g.

fluoroquinolones), and d) drug inactivation (e.g. β-lactamases against β-lactams).

Figure is re-published with permission (Allen, et al., 2010).

Increasingly outbreaks caused by antibiotic resistant bacteria are attributed to

misuse of antibiotics (Sandora, et al. , 2012). Exposure to non-lethal

concentrations of antibiotics in environments (e.g. rivers, soils and lakes)

(Alonso, et al. , 2001; Baquero, et al. , 2008; Cabello, 2006) and during

antibiotherapy (Baquero, et al. , 1998) are also among the contributing factors

to the selection and evolution of antibiotic resistant microorganisms

(Andersson, et al. , 2014). In addition, the ability of bacteria to form a biofilm

is also established as one of the causative factors contributing to the

emergence of antibiotic-resistant bacteria (Vuotto, et al. , 2014). Klebsiella

pneumoniae is among the most prevalent bacterial species causing biofilm-

associated nosocomial infections, such as catheter-associated urinary tract

infections (CAUTIs) (Amalaradjou, et al. , 2013; Nicolle, 2014), hence,

understanding the underlying mechanisms of biofilm formation by

K. pneumoniae is of the utmost importance to prevent or control biofilms-

associated infections caused by this pathogen.

4 |

1.1 Klebsiella genus

Klebsiella is a genus in the Enterobacteriaceae family together with other

clinically relevant species such as Escherichia coli, Salmonella spp., Yersinia

pestis, Shigella spp., Proteus spp., Serratia spp., Citrobacter spp. and

Enterobacter spp.. Bacteria belonging to the Klebsiella genus are ubiquitous,

frequently colonise humans and animals as commensals, particularly on the

mucosal surfaces of the gastrointestinal tract (Hennequin, et al. , 2009). They

are also abundant in the environment (Struve, et al. , 2004), for instance, in

sewage, water reservoirs, cultivation and vegetation areas and industrial

effluent (Bagley, 1985; Brown, et al. , 1973). The Klebsiella species are

characterised according to the diseases they cause and their distinct habitats.

K. pneumoniae itself contains three subspecies, which have been subdivided

according to the disease they caused: K. pneumoniae subspecies

pneumoniae, K. pneumoniae subspecies K. ozaenae, and K. pneumoniae

subspecies K. rhinoscleromatis (Ørskov, 1984). K. pneumoniae infections

commonly occur in blood, the urinary tract and within lung tissues, which will

eventually turn to a thick, bloody-mucoid sputum known as currant jelly

sputum. While K. ozaenae causes chronic atrophic rhinitis (ozaenae) and

K. rhinoscleromatis, causes chronic upper airways infections or rhinoscleroma

(Botelho-Nevers, et al. , 2007). Other than those prevalent subspecies, there

are seven additional Klebsiella species that have been identified (Brisse, et al.

, 2006): K. granulomatis (Carter, et al. , 1999) K. oxytoca (Lautrop, 1956),

K. planticola (Bagley, et al. , 1981), K. ornithinolytica (Sakazaki, et al. , 1989),

K.terrigena (Izard, et al. , 1981), K. mobilis (Bascomb, et al. , 1971), and

K. variicola (Rosenblueth, et al. , 2004).

5 |

1.1.1 Klebsiella pneumoniae

The genus Klebsiella was named after the famous German physician and

bacteriologist, Edwin Klebs. Discovered by the German bacteriologist and

pathologist, Carl Friedlander in 1882 (Brisse, et al., 2006), K. pneumoniae is a

non-motile, encapsulated, Gram-negative, rod-shaped (Abbott, 2007), and

facultative anaerobic pathogen. K. pneumoniae is also lactose positive, and

its colonies are normally dome-shaped, about 3-4 mm in diameter with mucoid

and sometimes sticky phenotypes, depending on the medium composition and

strains (Brisse, et al., 2006). K. pneumoniae strains which have a distinct

mucoid appearance (Figure 1-2) also known as hypermucoviscous phenotype

are frequently associated with hypervirulence (Shon, et al. , 2013). This

pathogen also has been identified as a common cause of community-acquired

bacterial infections, albeit, it is more prominently associated with nosocomial

infections, and mainly infects those immunocompromised individuals who are

hospitalised and severely-ill patients (Podschun, et al. , 1998).

Figure 1-2 Hypermucoviscous morphology of K. pneumoniae strain

Kpneu_110.

This strain exhibited a viscous string > 5 mm in length when probed by an inoculation

loop post-cultivation at 37 °C for 16 h on a blood agar.

6 |

1.2. Antibiotherapy for K. pneumoniae infection

K. pneumoniae has become progressively resistant to the vast majority of

antibiotics. Treatment of these particular infections is very challenging not only

because this pathogen is intrinsically resistance to ampicillin, but also

resistance to the majority of the antibiotics (Stahlhut, et al. , 2012) used for

treating bacterial infections in human. By harbouring blaSHV-1 (sulphydryl

variable type 1) on the chromosome, K. pneumoniae is known be intrinsically

resistant to ampicillin, amoxicillin, piperacillin, and also to early classes of

cephalosporins, for example, cephalothin (French, et al. , 1996; Haeggman, et

al. , 1997; Simpson, et al. , 1980).

Severely ill-patients who are infected by extended spectrum β-lactamase

(ESBL)-producing isolates are treated with 4th generation cephalosporins (e.g.

cefepime) (LaBombardi, et al. , 2006), aminoglycosides (e.g. gentamicin),

fluoroquinolones (e.g. ciprofloxacin) and the antibiotics of last resort,

carbapenems (e.g. imipenem or meropenem) (Paterson, et al. , 2004). These

antimicrobial agents can be prescribed as a monotherapy or a combination

therapy. Clinical success rates, however, are much higher when combination

therapies are used for treatment.

A previous study by Benenson, et al. (2009) showed that the combination of

gentamicin and a colistin was effective for carbapenem resistant

K. pneumoniae endocarditis. The combination of polymyxin B and rifampicin

proved to have a synergistic effect against K. pneumoniae carbapenemase-

producing strains in vitro (Bratu, Tolaney, et al. , 2005). Treatment options for

Klebsiella pneumoniae carbapenemase (KPC)-producing K. pneumoniae are

7 |

limited to colistin or tigecycline (Roy, et al. , 2013), and combination of both of

these antibiotics (Hirsch, et al. , 2010; Humphries, et al. , 2010). Furthermore,

studies by Deris, et al. (2012) and Zhang, et al. (2017) showed that the

combination of colistin or tigecycline with doripenem has a synergistic effect in

vitro. Ceftazidime/avibactam has also been proven to have excellent in vitro

activity, and might be suitable as a treatment to combat KPC infections (van

Duin, et al. , 2016).

In addition, bacterial infections that are caused by the non-ESBL producing

K. pneumoniae isolates are usually treated using 3rd generation

cephalosporins (e.g. cefotaxime), fluoroquinolones or a combination of a

β-lactam with a β-lactamase inhibitor, for example, ampicillin/sulbactam,

piperacillin/tazobactam or ticarcillin/clavulanate (Bhavnani, et al. , 2006).

8 |

1.3 Clinical manifestation of infections caused by

K. pneumoniae

K. pneumoniae is a commensal microorganism ubiquitously found colonising

human mucosal surfaces, for example, the gastrointestinal tract and

nasopharynx. Infections of K. pneumoniae can be divided into two groups:

community-acquired infections and hospital-acquired infections. Most

community-acquired infections cause pneumoniae or urinary tract infections.

A distinct K. pneumoniae invasive syndrome that causes pyogenic liver

abscess is being increasingly reported in Asian countries such as Taiwan in

the past two decades and have since emerged as a global disease (Ko, et al.

, 2002). It is also spreading rapidly as one of the most important causes of

hospital-acquired infections (nosocomial infections) in hospitals worldwide

(Onori, et al. , 2015; Podschun, et al., 1998). Hospital-acquired infections can

be defined as infections which develop in a patient at least 48 h after admission

to a hospital, without any symptoms prior to hospitalisation (Ducel, et al. ,

2002).

As an opportunistic pathogen, K. pneumoniae not only causes urinary tract

infections (UTIs), respiratory tract infections (RTIs), and liver abscess (Siu, et

al. , 2012), but it is also able to cause osteomyelitis (Prokesch, et al. , 2016),

septicaemia (Fane, et al. , 2005), and wound or surgical site infections (Virgilio,

et al. , 2016). K. pneumoniae is the second most prominent pathogen that

cause bacteraemia after E. coli (Yinnon, et al. , 1996) and amongst the most

common uropathogens that cause UTIs (Ronald, 2002). While invasive liver

abscess is established to be prevalent in Asian countries, the cause of this

9 |

predominance in Asian populations is still unknown. The strains isolated from

infected Asian people have been found to be more virulent than the strains

isolated from the patients outside Asia. Most of the patients with liver abscess

were observed to be infected exclusively by K1 and K2 capsular type

K. pneumoniae isolates (Siu, et al., 2012). These particular hypervirulent

strains are also known to be carried by healthy adult carriers in their intestines

and are suggested to be the origin of the infecting isolates (Fung, et al. , 2012).

Other than this antibiotic resistance attribute, the invasive syndrome that

causes liver abscess is also alarming since it has been increasingly reported

in Taiwan during the past few decades and has the highest prevalence of

cases related to liver abscesses (Siu, et al., 2012). Due to this fact, one of our

strains used in this study was isolated from a pyogenic liver abscess patient

and also was charaterised to have a hypermucoviscous phenotype and K2

capsular type (see Table 2-1).

In addition, K. pneumoniae is also known to cause medical device-associated

infections, for example, infections on urinary catheters (Nicolle, 2014),

endotracheal tubing (Boddicker, et al. , 2006), prostheses (de Sanctis, et al. ,

2014), central venous catheters (Buckley, et al. , 2007), artificial heart valve

(Raymond, et al. , 2014), and ventilators (Sbrana, et al. , 2016).

10 |

1.4 Mechanisms of antibiotic resistance

Antibiotic resistant bacteria have emerged not only in hospital settings, but also

in community settings. This suggests that reservoirs of antibiotic resistance

are also present outside hospitals, for example, in wastewater systems,

pharmaceutical and manufacturing effluent, and aquaculture and animal farms

(Berendonk, et al. , 2015). The emergence of antibiotic resistant bacteria is a

growing problem, while development of novel and efficient antibiotics is

progressing very slowly.

There are three primary antibiotic resistance categories: intrinsic, acquired and

adaptive. Intrinsic also referred to as ‘natural’ resistance of a bacterial species,

is due to inherent properties and functional characteristic of the cells. These

include harbouring of β-lactamases genes, and the capacity to reduce the

permeability of the bacterial cell membrane by alteration, modification and

decreased expression of porins (Mallea, et al. , 1998; Nikaido, 2003; Pangon,

et al. , 1989).

Numerous bacterial species are known to have intrinsic resistance to

β-lactams because they harbour β-lactamase genes on their chromosome.

Among the genes are AmpC β-lactamase gene, blaCMY-2, which is likely to be

originated from the chromosome of Citrobacter freundii (Bauernfeind, et al. ,

1996) and β-lactamase resistance genes that encode for CTX-M enzymes

which are suggested to have originated from the chromosome of Kluyvera

ascorbata (Humeniuk, et al. , 2002). In addition, intrinsic β-lactam resistance

in Stenotrophomonas maltophilia is mediated by two chromosomal-mediated

inducible β-lactamases resistance genes, blaL1 and/or blaL2 (Lin, et al. , 2009).

11 |

Other than these examples, intrinsic resistance in K. pneumoniae to penicillin,

ampicillin, amoxicillin, carbenicillin, oxacillin, and ticarcallin is mediated by

blaSHV-1, a gene that encodes for SHV-1 enzyme which is prevalently found on

the chromosome of vast majority of K. pneumoniae isolates (Paterson, et al. ,

2014).

The existence of an outer membrane in Gram-negative bacteria acts as an

intrinsic barrier to large antibiotic molecules that are effective against Gram-

positive bacteria (Blair, et al. , 2015; Miller, 2016). For example, the large

molecular size of vancomycin precludes it from being efficient to be used

against E. coli (Zhou, et al. , 2015) and K. pneumoniae (Pultz, et al. , 2005).

Another example of intrinsic resistance is resistance against erythromycin by

several Gram-negative bacteria for example K. pneumoniae (Sugawara, et al.

, 2016). This is due to the hydrophilic part of LPS at the outer membrane of

several Gram-negative bacteria, which interrupts the diffusion of erythromycin

into the cells, and subsequently provides a membrane impermeability to this

lipophilic antibiotic (Nikaido, 2003).

The second type is acquired resistance, where mobile genetic elements

(MGEs) are incorporated into the genome of susceptible bacteria, leading to

antibiotic resistance (Tenover, 2006). MGEs can be classified into two

different groups: elements that can transfer from one bacterial cell to another

cell, for example, plasmids and transposons, and those elements that can

translocate from one particular location to another location in the same

bacterial cell such as integrons and insertion sequences (Bennett, 2008). The

incorporation of MGEs into bacterial genomes happens by transferring the

12 |

genetic elements e.g. conjugative plasmid from one bacterium to another by

conjugation, translocation of mobile genetic elements e.g. transposons from

one location to another in the genome of a bacterium, and transducing the

portion of DNA from one bacterium to another by bacteriophages (van Hoek,

et al. , 2011). Conjugation by plasmids and transposons is known as one of

the most important mechanisms of acquired resistance as it could promote to

the direct transfer of antibiotic resistance genes from chromosome to

chromosome (Munita, et al., 2016).

The capacity to inactivate antibiotics is one of the significant resistance

mechanisms (Ammor, et al. , 2007), which is caused by acquisition of genes

encoding for antibiotic-degradation enzymes such as NDM-1 and antibiotic-

modifying enzymes for example, aminoglycosides N-acetyltransferases

(AAC). These AAC enzymes can modify aminoglycosides by catalysing

acetylation on the substrates (e.g. gentamicin), which subsequently leads to

poor binding of modified aminoglycosides to the 30S ribosomal subunit

(Ramirez, et al. , 2010) (details in Chapter 3, section 3.1.4.1).

Acquisition of genes encoding for β-lactamases and/or efflux pumps which can

be located on plasmids can also contribute to antibiotic resistance (Weinstein,

et al. , 2005). For instance, the spread of ESBL and KPC-producing

K. pneumoniae strains is usually facilitated by the transfer of plasmids

harbouring β-lactam and carbapenem resistance genes, for example,

blaCTX-M-15, blaKPC-1 and blaNDM-1. The blaCTX-M-15 gene was first discovered by

Karim, et al. (2001) on large plasmids with an insertion sequence, IS Ecp1

located upstream of the 5’ end of blaCTX-M-15 while blaKPC-1 was found to be

13 |

encoded on an approximately 50-kb non-conjugative plasmid (Yigit, et al. ,

2001). In addition, blaNDM-1 was discovered on a 180 kb plasmid of

K. pneumoniae 05-506 in 2008 by Yong, et al. (2009).

The AcrAB efflux pump, as an example, constitutes a major drug efflux system

towards large and lipophilic antimicrobial agent e.g. erythromycin and fusidic

acid (Nikaido, 1996a). Additionally, the capability to change the specific target

structure by genetic mutation or post translational modifications are also

proven to ultimately hinder the binding of the antibiotics to their target site

(Blair, et al., 2015). Addition of a chemical group to the target site of antibiotics,

for example by erythromycin ribosome methyltranferases (erm), which alter

the nascent 23S rRNA ribosomal binding site can also prevent the binding of

erythromycin and other antibiotics such as lincosamides and streptogramins B

(Leclercq, 2002). Furthermore, protection of the ribosome by ribosomal

protection proteins (RPPs), for example, by Tet(M) can prevent the binding of

tetracycline to the ribosome and allow the ribosome to continue the protein

synthesis (Connell, et al., 2003).

The third and final resistance category is adaptive resistance, which occurs

when bacteria obtain a transient resistant phenotype by altering their genes

expression and/or protein expression (Fernández, et al. , 2012). This can

happen when there is exposure to unfavourable environmental factors, such

as, nutrient starvation, low or high temperature and pH levels, and exposure

to sub-inhibitory concentrations of particular antibiotics (Poole, 2012). In

contrast to intrinsic and acquired resistance which are more stable, adaptive

14 |

resistance is transitory, and the bacteria normally revert back to their initial

phenotype after the inducing factor is removed (Fernández, et al., 2012).

1.4.1 Extended spectrum β-lactamase (ESBL) and

carbapenemase-producing K. pneumoniae

Beta-lactam family are a large group of antibiotics that includes penicillins,

cephalosporins, carbapenems, and monobactams (Danziger, et al. , 2011).

They are the most widely used antibiotics for bacterial infections particularly

for K. pneumoniae infections. There are two major antibiotic resistance

mechanisms that are common in K. pneumoniae: production of β-lactamases,

specifically expression of extended spectrum β-lactamases (ESBLs) which

render resistance to a broad range of β-lactam antibiotics including 3rd

generation cephalosporins (Paterson, et al., 2004) and production of

carbapenemases which render resistance to most β-lactam antibiotics

including carbapenems (Nordmann, et al. , 2009). These enzymes are

universally carried on the plasmids and also on chromosomes (Rawat, et al. ,

2010) They hydrolyse β-lactam by cleaving the β-lactam ring which is located

on the molecular structure of β-lactam antibiotics (Figure 1-3).

Figure 1-3 Chemical structure of ampicillin. β- lactam ring is circled in black.

15 |

Beta-lactamases can be classified into four major classes following the Ambler

or Jacoby and Bush classifications. Ambler classification is based on

conserved and distinguishing amino acid homology and divides β-lactamases

in four different classes. Class A, C and D are classified as serine beta-

lactamases, and hydrolyse β-lactams through a serine active site, whilst, class

B (metallo-β-lactamases) hydrolyse their substrates by utilising a zinc ion at

their active site (Bush, et al. , 2010; Paterson, 2006). While Jacoby and Bush

classification is based on the functional properties of enzymes, for example,

the substrate and the inhibitor profiles. The classification are divided into

group 1, 2, 3 (Bush, et al., 2010) and 4 (only in Bush-Jacoby Medeiros group

(Bush, et al. , 1995)) (refer Table 1-1).

16 |

Table 1-1 Classification schemes for bacterial β-lactamases.

The table shows classification schemes for bacterial β-lactamases and it is based on

the publication by Bush, et al. (2010).

17 |

1.4.2 Evolution of β-lactamases in K. pneumoniae

To date, antibiotic resistant bacterial infections have spread widely and have

become one of the most problematic nosocomial infections (Nordmann, et al.

, 2011). Rapid emergence of ESBL-producing K. pneumoniae and

carbapenemase-producing Enterobacteriaceae are occurring worldwide and

their threat is classified as serious and concerning health issue. Most ESBLs

are normally encoded by plasmids or chromosomes and generally are

derivatives of non-ESBL enzymes, such as, blaTEM-1, blaTEM-2 or blaSHV-1 (Bush,

et al., 1995). They underwent simple point mutations which resulted to

alteration of the amino acid sequence around the enzyme active sites and

become ESBLs, such as, blaTEM-3 or blaSHV-2. TEM and SHV-type ESBLs are

found in E. coli and K. pneumoniae and other genera of Enterobacteriaceae.

The first plasmid-borne β-lactamase (blaTEM-1) in the Enterobacteriaceae was

discovered early 1965 in Athens, Greece. It was originally found in

E. coli which was isolated from the blood culture of a patient named

Temoneira, hence, it got its designation, TEM (Datta, et al. , 1965; Medeiros,

1984). The first TEM-type enzymes, blaTEM-3, that conferred ESBL phenotype

was reported in 1988. This variant was found to have two amino acid

substitutions in comparison to blaTEM-2 at position E102K and G236S

(Sougakoff, et al. , 1988). In addition, the first case of ESBL-producing

K. pneumoniae was reported at University Hospital in Frankfurt, Germany in

1983 (Knothe, et al. , 1983). This enzyme could hydrolyse oxyimino-

cephalosporins and was designated as blaSHV-2, which was derived from

blaSHV-1. It differs from blaSHV-1 of E. coli p452 by one amino acid substitution

18 |

at position G213S (Barthelemy, et al. , 1988) and all blaSHV variants reported

nowadays are derivation of blaSHV-1.

In the last few years, a newly discovered family, the CTX-M-family enzymes

have replaced TEM and SHV-derived enzymes as the most prevalent in

clinical Enterobacteriaceae isolates. The first CTX-M-1 was discovered in

1989 in Munich from E. coli strain GRI, which was isolated from an exudate of

the ear of a newborn suffering from otitis media (Bauernfeind, et al. , 1990). In

addition, CTX-M-15 and CTX-M-14 are by far is the most important enzymes

and the most common CTX-M variants identified worldwide (Zhao, et al. ,

2013). These enzymes were designated as CTX-M (cefotaximase, Munich)

because of their capacity to hydrolyze cefotaxime more efficiently than

ceftazidime (Tzouvelekis, et al. , 2000).

Apart from these enzymes, carbapenemase such as KPC-1 is also identified

as one of the most important β-lactamases and this resistance mechanism is

established as a causative factor contributing to the global spread of antibiotic

resistant bacterial infections. The first K. pneumoniae carbapenemase (KPC)-

producing K. pneumoniae isolate was discovered in a North Carolina hospital

in 1996 (Yigit, et al., 2001) and emerged rapidly in the New York City hospitals

outbreak (Bratu, Landman, et al. , 2005). A recent case of pan-resistant

K. pneumoniae was reported in January 2017. The report on the 70-years-old

woman in Nevada, United States was deeply worrying since she was infected

with K. pneumoniae which harboured the New Delhi metallo-β-lactamase

(blaNDM-1) (Yong, et al., 2009) resistance gene and could not be treated with

any relevant and available antibiotics in the United States (Chen, et al. , 2017).

19 |

The ESBL-producing strains of K. pneumoniae are able to hydrolyse

β-lactams, for instance 3rd generation cephalosporins (e.g. cefotaxime,

ceftazidime, and ceftriaxone), and monobactams (e.g aztreonam).

Additionally, they can also possess resistance against different groups of

antibiotics, as an example, aminoglycosides and fluoroquinolones (Turner,

2005). In addition, worse than ESBL-producing strains, bacterial isolates that

harbouring carbapenemases are resistant to nearly all antibiotics including to

β-lactams, such as penicillins, 3rd and 4th generation cephalosporins,

monobactams, fluoroquinolones, aminoglycosides, and carbapenems (Arnold,

et al. , 2011).

1.4.3 Epidemiology of multidrug-resistant K. pneumoniae

The expansion of antibiotic resistance cases involving multidrug-resistant

K. pneumoniae have been increasingly reported over the past few decades.

The rapid spread of particular strains harbouring antibiotic resistance genes,

such as carbapenemase resistance genes across a region and a country is

also worrying.

The first isolate of K. pneumoniae harbouring blaKPC on a plasmid was isolated

in 1996 in North Carolina and it was then observed to spread rapidly across

New York City in 2003. By 2007, it was reported that 21 % of Klebsiella strains

in the city carried this blaKPC gene in comparison to the average of 5 % across

the rest of the United States (Hidron, et al. , 2008). This rapid dissemination

was thought to travel from person to person since Enterobacteriaceae such as

K. pneumoniae inhabit the human gut and it can be carried away by

asymptomatic patients. Bacteria harbouring blaKPC were then moved to Tel

20 |

Aviv’s Sourasky Medical Centre in 2005 and subsequently to Italy, Colombia,

the UK and Sweden (McKenna, 2013). The warning alarm was triggered in

2008 where doctors discovered a K. pneumoniae strain from a urine culture of

a 59-year old patient who came back from India and was hospitalised in

Sweden. The strain was highly resistant to carbapenems more than KPC-

producing K. pneumoniae (see Figure 1-4). This particular strain hydrolysed

most of the antibiotics with a unique enzyme which was encoded by blaNDM-1

and designated as NDM-1 (Yong, et al., 2009).

Figure 1-4 The resistance movement of carbapenem-resistant

Enterobacteriaceae.

Figure is re-published with permission (McKenna, 2013).

21 |

In addition to the movement of carbapenem-resistant Enterobacteriaceae

worldwide, the movement of other resistance mechanisms of K. pneumoniae

has also been reported in Europe for over the last decade. The antibiotic

resistance rates in Europe have been observed to increase particularly from

year 2005 until year 2016 for most of the countries except the UK (Figure 1-

5).

22 |

Figure 1-5 Surveillance atlas of the increasing number of K. pneumoniae antibiotic resistance cases starting from A) 2005 to B) 2016

in Europe.

The percentage (%) of invasive isolates with combined resistance to fluoroquinolones, aminoglycosides, and carbapenems by country, EE/EEA

countries. The figures were downloaded from https://ecdc.europa.eu/en/surveillance-atlas-infectious-diseases.

23 |

The largest increase was observed for Italy which was reported to have less

than 1 % of resistance cases in 2005, which increased to somewhere between

25 % to < 50 % of cases in 2016. There was no increase seen for Greece

which had high number of cases in 2005 until 2016 (25 % to < 50 %). Italy and

Greece were also the only countries in the Europe to have highest incidence

of carbapenemase-producing Enterobacteriaceae infections. It had been

reported that both of these countries were having an ‘endemic situation’ for

these particular bacterial infections in 2013-2014 which hospitalisation cases

in most of the hospital in both of the countries were from local population and

sources (Albiger, et al. , 2015).

There was a slight reduction in the number of cases that were reported in the

UK in 2016 (1 % to < 5 %) in comparison to 5 % to < 10 % in 2005. Based on

the ECDC Surveillance Report of Antimicrobial Resistance in Europe 2016, at

the European Union/ European Economic Area (EU/EEA) level, 34.5 % of

K. pneumoniae isolates were reported to The European Antimicrobial

Resistance Surveillance Network (EARS-Net) in 2016 were resistant to at least

one of the regular surveillance antibiotics groups, fluoroquinolones, third-

generation cephalosporins, aminoglycosides, and carbapenems. The highest

percentage of antibiotic resistance was reported for 3rd generation

cephalosporins (25.7 %). This is followed by fluoroquinolones with 24.6 %

cases, aminoglycosides with 19.0 % and carbapenems with 6.1 % cases

(ECDC, 2017).

24 |

Beside Europe, the spread of multidrug-resistant K. pneumoniae is increasing

globally including in the Middle East and also in Asian region countries. Egypt

was found to be one of the countries with the highest rates of ESBL among

Enterobacteriaceae (38.5 %) (Abdallah, et al. , 2015) according to the report

of study in 2001 – 2002 by Pan European Antimicrobial Resistance Local

Surveillance (PEARLS) in 13 European countries, three Middle East countries,

and also South Africa.

Owing to high rates of antibiotic resistance cases in Egypt, it is important to

investigate the prevalence of resistance mechanisms from year to year in this

country. Since information about ESBL-producing Enterobacteriaceae (ESBL-

E) in Egypt is still limited (Abdallah, et al., 2015) and to obtain more information

on this, all of the clinical K. pneumoniae isolates used in this study are the

strains that were isolated from community or hospital-acquired infections in

2000, 2003, and 2009 from three different university hospitals in Egypt (see

Table 2-1). All of the strains were also confirmed as multidrug resistant strains

(see Table 2-2).

Apart from Europe and Middle East countries, the emergence of ESBL-E has

also been reported in Asian countries. The Study for Monitoring Antimicrobial

Resistance Trends (SMART) in 2008 which examined the in vitro antimicrobial

susceptibility to 2370 unique facultative-aerobic Gram-negative bacilli isolates,

which associated with intra-abdominal infections to 12 antimicrobial agents,

high rates of ESBL-producing E. coli and K. pneumoniae isolates were

observed in China with 59.1 % and 34.4 % respectively, while in India, the

25 |

rates were 61.2 % and 48.8 % respectively, and in Thailand, they were

53.0 % and 23.1 % respectively (Hsueh, et al. ).

In addition, there was also an evidence of increasing number of ESBL-

producing K. pneumoniae isolates from 15 % – 23.4 % in a 7-year study

starting from 2001 to 2008 at the largest tertiary hospital in Taiwan (Shu, et al.

, 2009). The emerging threat of antibiotic resistant bacterial infections in

Taiwan is shown to be related with nosocomial infections that are caused by

ESBL-producing K. pneumoniae, however, the data regarding community-

acquired infections are still limited (Lin, et al. , 2016).

1.5 Mechanisms of K. pneumoniae pathogenicity

Understanding the mechanisms of virulence are pivotal to determine their

complex interaction with host cells and could explain the underlying process

that leads to bacterial pathogenesis. The most common site for

K. pneumoniae infection is the urinary tract and K. pneumoniae possesses a

number of virulence determinants to colonise that particular tract by adhering

and invading the host cells (Struve, et al. , 2008), and to escape the natural

host defense mechanisms, such as protection from phagocytosis (Clegg, et al.

, 2016; Schembri, et al. , 2005). Those factors are capsules,

lipopolysaccharides, siderophores and fimbriae (Figure 1-6). Even though the

outer membrane porins (Omps) have not yet been thoroughly characterised

and have received less attention, some have previously been found to play a

role in virulence mechanisms of K. pneumoniae (Tsai, et al. , 2011).

26 |

Figure 1-6 Virulence factors of K. pneumoniae.

There are four well-studied virulence factors of K. pneumoniae: capsule, LPS,

siderophores and fimbriae (type 1 and type 3). Omps (e.g. OmpK35 and OmpK36)

have received less attention, even though, their role in virulence in murine model has

been demonstrated (Tsai, et al., 2011). Figure is re-published with permission

(Fernández, et al., 2012; Paczosa, et al. , 2016).

OMP

27 |

1.5.1 The capsule

The capsule, also known as the K antigen, is a dense and thick

polysaccharides matrix (~160 nm of capsule thickness) (Amako, et al. , 1988).

It is known as a major surface-located virulence factor, and encases the whole

bacterial cell (Schembri, et al., 2005). There are around 78 different capsular

types which have been reported up until 2013 (Pan, et al. , 2013).

Overproduction of the capsule is known as a hypercapsule, or a

hypermucoviscous phenotype and is associated with hypervirulent strains of

K. pneumoniae. Such strains have a mucoviscous polysaccharide which is

more robust than the typical capsule (refer to Figure 1-2). The capsule

protects K. pneumoniae against desiccation and phage attack (Schembri, et

al. , 2004), and the role of capsule in pathogenicity of K. pneumoniae infections

is also well understood. The capsule is well-known to protect

K. pneumoniae against the host immune response during infection through

several important mechanisms: evading phagocytosis by immune cells such

as polymorphonuclear leukocytes (i.e. neutrophils), dendritic cells and

macrophages (Collins, et al. , 1996), preventing activation of early immune

responses which leads to a more severe infections, and also escaping cell

lysis mediated by complement and antimicrobial peptides (Doorduijn, et al. ,

2016).

Hypervirulent K. pneumoniae strains are also considered to be more resistant

to phagocytosis because of modified polysaccharides in the capsule structure,

and indirectly circumvent the binding to phagocytic cells. As an example,

Klebsiella strains which produce capsule with certain repeating mannose and

28 |

rhamnose sequences, for instance, mannose-α2/3-mannose or L-rhamnose-

α2/3- L-rhamnose disaccharide can bind efficiently to a mannose specific lectin

which is normally found on tissue macrophages. Whilst, capsule types K1, K2,

K7, K12, K21a K24, K32, and K55 which do not contain these repeating units

are not recognised easily by macrophages, and this possibly causes

resistance to host immune responses (Athamna, et al. , 1991; Collins, et al.,

1996). As reported previously, K2 serotypes have a modified glycan

composition to thwart recognition by the lectin pathways (Sahly, et al. , 2009).

1.5.2 Lipopolysaccharides (LPS)

K. pneumoniae endotoxin also called lipopolysaccharides (LPS) is a major and

necessary component of the outer leaflet of the Gram-negative bacteria cell

membranes, which is important in maintaining the structure and outer

membrane stability, and also for environmental adaptation. It is comprised of

three distinct structural domains: i) the hydrophobic lipid A, ii) the core

oligosaccharide antigen, which is connected to lipid A and provides the

attaching side for iii) O- antigen which is the long chain of polysaccharides

(Clements, et al. , 2008).

Lipid A, a hydrophobic membrane anchor is a pathogen associated molecular

pattern (PAMP) which is recognised by a host pattern recognition receptor

(PRR), for example, toll-like receptor 4 (TLR4) which is present on many cell

types including macrophages and dendritic cells, and initiates the clearance of

bacterial infection (Steimle, et al. , 2016). The capsule is proposed to play a

synergistic role with lipid A by partially shielding the PAMP from detection by

PRR. The core antigen, which connects O antigen to lipid A molecules

29 |

consists of negatively charged mono, di or oligosaccharides, for instance, two

to three 3-deoxy-D-manno-oct-ulosonic acid (KDO) residues (Clements, et al.,

2008). The outmost component of LPS, glycan polymer or O antigen consists

of a polymer of oligosaccharide repeating units. There are 11 different

O-antigen types that have been identified in K. pneumoniae isolates in contrast

to the large group of capsular serotypes (Vinogradov, et al. , 2002).

1.5.3 Siderophores

Siderophores are low molecular weight iron-chelating molecules that maintain

a higher affinity for iron compared to host iron transport system. Siderophores

scavenge iron in the environment and sequester iron from host-iron chelating

proteins to grow efficiently in that particular host tissue. Iron is critical for many

cellular mechanisms of K. pneumoniae, including oxygen metabolism, DNA

replication and as a cofactor for the electron transport chain (Messenger, et al.

, 1983). Because of this, K. pneumoniae needs to acquire iron from the host

in order to propagate or to survive in low-iron environments.

There are several siderophores that are produced by K. pneumoniae:

enterobactin (Ent), yersiniabactin (Ybt), salmochelin (Sal) and aerobactin.

Enterobactin, a prototypic catecholate siderophore is the primary iron uptake

system in K. pneumoniae, and is prevalent in respiratory tract isolates of

K. pneumoniae. Yersiniabactin is a phenolate siderophore, while salmochelin

is a c-glucosylated derivative form of enterobactin (Holden, et al. ; Russo, et

al. , 2015). Aerobactin, an uncommon citrate hydroxamate iron-chelating

compound is the dominant siderophore produced by hypermucoviscous

K. pneumoniae strains (Russo, et al., 2015).

30 |

1.5.4 Fimbriae

Bacteria adhere to host cells in order to aid initial colonisation. Fimbriae

mediate adhesion of bacterial cells to host cells through interaction of the

adhesins with receptors on host surfaces. There are many types of bacterial

appendages which are important for adhering to host cell surfaces, for

example, fimbriae, curli and also flagella. Those surface appendages must be

sufficiently sturdy and firm, to anchor on the host cell surface to resist forces,

such as the impact of aqueous flushes or gaseous flushes in the urethra or the

respiratory tract, respectively (Chen, et al. , 2011). Fimbriae are among the

most important surface appendages in K. pneumoniae. The most common

K. pneumoniae fimbriae are the type 1 fimbriae (T1F) and type 3 fimbriae (T3F)

which were initially identified in Klebsiella strains as ‘MS adhesin’ (mannose-

sensitive) and ‘MR adhesin’ (mannose-resistant), respectively (Duguid, 1959).

Additionally, K. pneumoniae has other types of fimbriae, for example, Kpc

fimbriae (Wu, et al. , 2010), and E. coli common pilus (EPC) (Alcántar-Curiel,

et al. , 2013). The details of T1F and T3F are discussed in Chapter 5.

31 |

1.5.5 Outer membrane porins (Omps)

The outer membrane of Gram-negative bacteria consists of a lipid bilayer,

which is essential as a first line shield against any toxic substances, and it is

also impermeable to large and charged molecules, such as some antibiotic

molecules (Pagès, et al. , 2008).

The influx transport system embedded in bacterial outer membrane is partly

facilitated by porins (Delcour, 2003; Nikaido, 2003; Pagès, et al., 2008),

proteins found in the outer membrane layer of Gram-negative bacteria

(Sutcliffe, et al. , 1983). Porins are abundant water-filled channels, which

permit the passive penetration of hydrophilic molecules through the outer

membrane layer (Figure 1-7) (Galdiero, et al. , 2008; Nikaido, 2003). Outer

membrane porins (Omps) are crucial to allow different substances to penetrate

(e.g. ions, amino acid, nucleosides, iron, and nutrients) the cells and to extrude

molecules (e.g. toxic and waste products) from the cells across the outer

membrane barrier, specifically in Gram-negative pathogens (Martínez‐

Martínez, 2008).

32 |

Figure 1-7 Gram-positive and Gram-negative cell walls.

A) The cell wall of Gram-positive bacteria consists of a thick peptidoglycan layer

located outside of the cytoplasmic membrane. Teichoic acids are anchored in the

peptidoglycan layer while the lipoteichoic acids are protruded and extended into the

cytoplasmic membrane. B) The cell wall of Gram-negative bacteria is comprised of

an outer membrane, cytoplasmic or inner membrane, and peptidoglycan layer which

is located in the periplasmic space between the outer and inner membrane. The outer

membrane structure includes porins and LPS. Porins act as channels for small

molecules to pass through the membrane and LPS or endotoxin which is extended

into extracellular space act as a structural support and physical barrier for Gram-

negative bacteria. Figure is re-published with permission (Cabeen, et al. , 2005).

33 |

Generally, porins are divided into two distinct types according to their activity:

non-specific porins and specific porins. Nonspecific porins (e.g. OmpC, OmpF

and PhoE) for example, in E. coli and K. pneumoniae, are responsible for a

general diffusion of small and polar molecules, by having eyelet (also known

as L3 structure) within the porins that permit permeating solutes of

approximately 600 Da (Nikaido, 1994). While specific porins (e.g. sucrose-

specific porin (ScrY) and maltose-induced maltoporin, (LamB)) aid in diffusion

of specific substances (Doménech-Sánchez, et al. , 1999; Hardesty, et al. ,

1991). Most hydrophilic and small antibiotics, for instance, β-lactams (Nikaido,

1988), tetracycline, fluoroquinolones, and chloramphenicol (Nikaido, 1996a)

penetrate the outer membrane layer through porin channels. β-barrel porins

not only provide exclusion limits of typically 600 Da for many antibiotic

molecules but they also limit the passive diffusion of large-scaffold antibiotics

molecules, for example, erythromycin (734 Da), rifampicin (823 Da) and

vancomycin (1450 Da) into Gram-negative bacterial cells. These large

antibiotics are thought to diffuse through LPS instead of porins, but the process

is inefficient and considered to be ineffective to be used against Gram-negative

bacteria (Muheim, et al. , 2017).

Depending on the species and environment, bacteria can have up until 105

porins per cell (Nikaido, 1988). Delcour (2003) suggested that OmpF and

OmpC families allow diffusion of cationic molecules, whereas PhoE favours

inorganic phosphate and anions (Nikaido, 1996b). The details of OMP of

K. pneumoniae are discussed in Chapter 6.

34 |

1.6 Biofilm formation by K. pneumoniae

Biofilms are microbial aggregates, which adhere to inert, living or non-living

surfaces (i.e. metals, plastics, soil particles, indwelling medical devices, and

host tissues). They are enclosed in a self-produced matrix of hydrated

extracellular polymeric substances (EPS). The biofilms can be made up of

single and multiple species of microorganisms that adhere to substratum in

aqueous environment and produce slimy EPS matrix (Flemming, et al. , 2010;

Mah, et al. , 2001) for protection. EPS can consist of polysaccharides,

proteins, extracellular DNA (eDNA) and glycolipids. These biopolymers are

responsible for interactions including for cohesion between cells in the biofilms,

adhesion of cells to surfaces, providing mechanical stability to the biofilms, and

protection from dehydration and the host immune defense cells (Flemming,

2011). EPS is also important in bacterial pathogenesis since it mediates

resistance of bacteria to antibiotics, protects the biofilms from host immune

responses, and promotes the development of persister cells (Bales, et al. ,

2013). The structure of biofilms depends on several factors including the

environment, nutrients, shear forces, hydrodynamic conditions, intercellular

communications and metabolic pathways (Flemming, et al., 2010).

The identification of the main components of K. pneumoniae biofilm EPS is still

yet to be established. However, Flemming, et al. (2010) suggested that

cellulose might play an important part in the biofilm EPS of K. pneumoniae.

This is supported by the study by Huertas, et al. (2014), showing that deletion

of yfiR, increased the expression of yfiN and resulted in increased cellulose

production. Increased cellulose production led to robust biofilm formation by

35 |

the mutant strain, LM21ΔyfiR in comparison to the wild type LM21 strain. yfiR

encodes for YfiR, a negative regulator which controls the YfiN diguanylate

cyclase (DGC) protein activity, while YfiB is a sensor protein. This yfiRNB

operon encodes for a set of proteins which are known to be involved in the

regulation of the second messenger, cyclic di-guanosine monophosphate (c-

di-GMP) synthesis (Wiebe, et al. , 2015).

Lipopolysaccharides, capsular polysaccharides, type 1 fimbriae (T1F) and

type 3 fimbriae (T3F) are the important virulence factors that have been shown

to be associated with biofilm formation by K. pneumoniae. A surface

exopolysaccharide, LPS was shown to be involved in biofilm formation by

K. pneumoniae. A study by Balestrino, et al. (2008) showed that several

mutants obtained from the signature-tagged mutagenesis mutant library

lacked the genes which were involved in LPS and capsule production. The

biofilm assay comparing between mutant and parental strains and microscopic

analyses showed that LPS was involved in the initial adhesion of

K. pneumoniae to both glass and polyvinyl-chloride, while, capsule was shown

to play an important role in initial coverage of the substratum and development

of mature biofilm structure.

In addition, the involvement of another surface polysaccharide, capsular

polysaccharides in biofilm formation by K. pneumoniae was also supported by

a previous study by Boddicker, et al. (2006). This study showed that mutations

in several loci including the K12 capsular gene cluster by transposon Tn5

caused a deficiency in biofilm formation by K. pneumoniae strain 43816.

36 |

Even though capsular polysaccharide has been shown to play a significant role

during formation of mature biofilm, Schembri and colleagues (2005) have

shown that capsulation can also interfere with, but did not abolish, the function

of T1F, which is known to be predominantly involved in the initial adhesion

stage of biofilm formation. Expression of capsule was demonstrated to notably

affect the FimH adhesin, thereby interfering with adhesion of K. pneumoniae

to surfaces. The interference might be due to several reasons: i) the FimH

adhesins might be coated and masked by the polysaccharide capsular

material, ii) the molecules of FimH adhesins that penetrate the capsule might

be coated by polysaccharide capsular material thus impeding the function for

adhesion and attachment of K. pneumoniae onto the surface, iii) fimbrial

organelles are proned to breakage due to the presence of capsule that could

result in structural weak points, and iv) the flexibility of fimbriae that is important

for optimal adhesion with other receptors on the cell surface could be reduced

by the expression of capsule (Schembri, et al., 2005). However, there is still

conflicting report regarding the role of capsule in biofilm formation by

K. pneumoniae. For instance, encapsulated bacteria was also previously been

reported to adhere weakly to epithelial cells and had reduced capacity for

invasion compared to non-encapsulated K. pneumoniae strains (Sahly, et al. ,

2000).

37 |

Table 1-2 Involvement of capsule in biofilm formation by several K. pneumoniae strains.

Bacterial strains Details of involvement of capsule in biofilm formation References

MGH 78578 Capsule deletion was shown to perturb biofilm formation since two mutant strains,

Δwzabc and Δcps that had less capsule, formed higher amounts of biofilm in

comparison to the wild type strains. In the Δwzabc mutant strain, overexpression

of pga alone was suggested to be sufficient for increased biofilm formation by

K. pneumoniae, while in the Δcps mutant strain, overexpression of pga and T1F

were thought to play a synergistic role in biofilm formation by

K. pneumoniae.

(Huang, et al. , 2014)

LM21 Capsule was shown to play a role in biofilm formation by K. pneumoniae. Mutant

strains which had disruption in genes involved in capsule biosynthesis formed

less biofilms relative to wild type strain.

(Balestrino, et al., 2008)

43816 Capsule was shown to likely to be involved in biofilm formation by

K. pneumoniae. A strain with a mutation in ORF12 of the capsule gene cluster

had decreased capsule production and subsequently formed less biofilm in

comparison to wild type strain.

(Boddicker, et al., 2006)

38 |

Bacterial strains Details of involvement of capsule in biofilm formation References

C105 Capsule could interfere with biofilm formation by K. pneumoniae. A non-

encapsulated strain formed more biofilm than the parental strain and this

observation was also correlated with expression of T1F.

(Schembri, et al., 2005)

Kp25, Kp57 and Kp65 wcaG which encodes for WcaG protein that involved in the biosynthesis of fucose

was found to be associated with K. pneumoniae bacteremia biofilm formation.

WcaG was speculated to promote biofilm formation of K. pneumoniae by altering

the composition of capsular polysaccharide

(Zheng, et al. , 2018)

39 |

Beside capsule, previous studies have shown that T1F and T3F are

coordinately controlled in biofilm formation by K. pneumoniae. A study by

Murphy, et al. (2013) found both of these fimbriae contributed to bacterial

adherence in a murine catheter associated urinary tract infection model, and

Stahlhut, et al. (2012) also showed that T1F and T3F were coordinately

controlled under a strict co-regulation system and the capacity to form biofilm

was severely diminished when both of them were absent.

However, the role of T1F in biofilm formation is shown to be dependent on the

experimental conditions such as the experimental model, characteristics of the

substratum, and growth medium used for cultivation of the biofilm. For

example, by using well defined isogenic mutants, Schroll, et al. (2010) reported

that T3F but not T1F promoted biofilm formation by K. pneumoniae C3091

upon cultivation in a fastidious anaerobic medium in a flow system. T1F was

found to be downregulated in the biofilm community upon cultivation under this

conditions. While, Stahlhut, et al. (2012) reported that T1F are as important

as T3F in promoting biofilm formation on catheter upon cultivation in pooled

human urine in a catheterised bladder model. This study also clearly showed

that both fimbrial types are coordinately controlled which means that T1F and

T3F do not simultaneously express in one individual cells; if T1F is expressed,

T3F is not expressed and vice versa. For example, during biofilm formation

on urinary catheters, each individual cell was observed to express either T1F

or T3F, but not both. However, neither of the fimbriae were found to be

expressed in planktonic cells.

40 |

To date, the relationship between Omps and biofilm formation by

K. pneumoniae is still unclear. However, there is evidence from several

studies that demonstrated that OmpC of E. coli and OmpC of Salmonella

enterica serovar Typhi, homologs of OmpK36 of K. pneumoniae could play a

role in initial adhesion or biofilm formation by E. coli, S. enterica ser. Typhi or

S. enterica ser. Typhimurium on biotic and abiotic surfaces (Gonzalez-

Escobedo, et al. , 2011; Rolhion, et al. , 2007). In addition, the ompC of E. coli

K-12 was observed to be among the upregulated genes in sessile bacteria

compared to planktonic cells in a high osmolarity environment (Prigent-

Combaret, et al. , 1999). The OmpC of S. enterica ser. Typhimurium, which is

identical to the OmpC S. enterica ser. Typhi was also shown to be essential

for binding and biofilm formation by S. enterica ser. Typhimurium on

cholesterol surfaces in the presence of bile (Crawford, et al. , 2010). OmpC is

thought to be essential for biofilm formation in the presence of bile possibly

because it mediates the electrostatic interaction between bacterial cells and

the cholesterol surface, acts as a nutrient channel which is involved in biofilm

health and also controls cell-cell interactions in the biofilm as it does in biofilms

of E. coli (Pruss, et al. , 2006). This finding is also supported by Gonzalez-

Escobedo, et al. (2011), who demonstrated that loss of OmpC negatively

affected the initial binding of S. enterica ser. Typhi on cholesterol gallstones.

In the light of these previous findings, the role of OmpK35 and OmpK36 in

pathogenic K. pneumoniae biofilm formation was investigated in this study.

41 |

1.7 Mechanisms of antibiotic resistance in bacterial biofilms

The formation of biofilms is an important resistance mechanism towards

antibiotics and confers protection against host defense mechanisms. Bacteria

have an increased capacity to withstand antibiotic-mediated killing once a

biofilm is formed. The ability of biofilms to endure in the environment in the

presence of high concentrations of antibiotics is defined as recalcitrance

(Lebeaux, et al. , 2014). The recalcitrance phenomenon toward antibiotics in

biofilms is still clinically difficult to manage. It can lead to failure of treatment,

recurrence of infection, and mortality. Tissue-related and device-related

infections by K. pneumoniae are very problematic and have been associated

with recurring infections. There is a lot of evidence showing that biofilm

communities of K. pneumoniae are more resistant to antimicrobial agents than

their planktonic counterparts (Davies, 2003; Lewis, 2001). The resistance

mechanisms which enable biofilms to be less susceptible to antibiotics than

planktonic cells have been attributed to several main hypotheses (Mah, et al.,

2001; Stewart, 2002).

The first hypothesis is the delay or failure of the antibiotic to penetrate through

biofilms. The biofilm matrix layer which is complex and dense matrix or EPS

can consist of polysaccharides, proteins, eDNA, and lipids which interrupts the

penetration of antibiotics through the thick layer of matrix (Stewart, et al. ,

2001).

The second hypothesis posits that bacteria grow slower in biofilms. Bacteria

which are deep within biofilms are dividing at relatively slower rates than those

which are closer to the surface. This is probably due to gradient of nutrients

42 |

and oxygen, signalling factors, waste products, and other substrates within

biofilms that leads to metabolically inactive and slow growth of the cells in the

biofilm community (Wimpenny, et al. , 2000). Slow growing or non-growing

bacterial cells are resistant to antimicrobial challenge as most of the antibiotics

act upon rapidly dividing cells. Penicillin and ampicillin, for example, only kill

growing cells instead of non-dividing cells (Lewis, 2001).

43 |

1.8 Aims and hypothesis of this study

K. pneumoniae is an opportunistic pathogen that primarily causes urinary tract

infections, septicaemia, and respiratory tract infections. The ability to form

biofilms on a medical device, for example, urinary catheters, is known as one

of the important causative factors contributing to the nosocomial

K. pneumoniae infections, and subsequently, the treatment of K. pneumoniae

infections becomes unmanageable.

We hypothesised that exposure of K. pneumoniae cultures and mature biofilms

to sub-minimum inhibitory concentrations (sub-MICs) of antibiotics would

enhance biofilm formation by K. pneumoniae. The aims of the first chapter of

this PhD project were to identify the effect of antibiotic (with different modes of

actions) exposure on biofilm formation by K. pneumoniae since there are

studies showing that antibiotic exposure could enhance and stimulate biofilm

formation in other pathogenic bacteria.

We also hypothesised that the capacity to form biofilm in different

environments and the main components of the biofilm matrix of K. pneumoniae

biofilms are strain dependent. To test this hypothesis, in the second chapter,

we determined the capacity of K. pneumoniae to form biofilm in two different

media, nutrient-rich and nutrient-poor media and identify the EPS produced by

tested K. pneumoniae strains during biofilm formation in both of the media.

We conducted enzymatic biofilm inhibition assays and biofilm disruption

assays upon cultivation of the examined K. pneumoniae strains in nutrient-rich

medium, tryptic soy broth (TSB), and nutrient-poor medium, artificial urine

medium (AUM).

44 |

In the third chapter, we hypothesised that T1F possibly plays a role in biofilm

formation by K. pneumoniae upon cultivation in TSB and AUM. The role of

T1F in biofilm formation was investigated by determining the orientation of fimS

which controls the expression of fimA, the major subunit of T1F, in planktonic

and biofilm cells cultivated in TSB and AUM. To examine this, biofilm mass

for all of the examined strains was compared upon cultivation in TSB and AUM,

and the orientation of fimS was then determined. This knowledge could

improve the understanding of T1F regulation and their involvement in biofilm

formation by K. pneumoniae in different environments.

In addition, based on the findings in Chapter 4, we also hypothesised that outer

membrane porins (Omps), could possibly be involved in biofilm formation by

K. pneumoniae since other studies have shown a role for Omps. For example,

OmpC was shown to be involved in biofilm formation by S. enterica ser. Typhi

and was also found to be upregulated in biofilm cells of E. coli. The role of

Omps in biofilm formation was investigated by developing ∆ompK35 and

∆ompK36 mutant strains of K. pneumoniae, and comparing their capacity to

form biofilm on abiotic surfaces in comparison to the parental wild type strains.

Kpneu_1 and C3091, and its mutant strains were subsequently tested for their

capacity to form biofilms on silicone catheter cultivated in TSB and AUM. To

examine this, biofilm assays using catheter pieces inoculated with mutant or

wild type strains cultivated in TSB and AUM were conducted.

45 |

Chapter 2.0

Materials and Methods

46 |

2.0 Materials and Methods

This chapter describes the materials and methods that were used throughout

the study. The specific materials and methods for each experiment within a

chapter are outlined at the beginning of that particular chapter.

2.1 Sources of media, enzymes and reagent

Tryptic soy broth (TSB), Luria-Bertani (LB), agar bacteriological no 1, Mueller-

Hinton (MH) broth, and brain heart infusion (BHI) were obtained from Sigma

(Sigma, UK). All antibiotics were obtained from Sigma-Aldrich (Poole, UK) and

were used at a concentration of 100 µg/ml ampicillin (amp), 100 µg/ml

spectinomycin (spc), 50 µg/ml apramycin (apra) and 12.5 µg/ml tetracycline

(tet). Erythromycin (erm), gentamicin (gen), and ciprofloxacin (cipro) were

used at different concentrations depending on the experiments. All of the

restriction enzymes were obtained from New England Biolabs (NEB),

(Hertfordshire, UK).

47 |

2.2 Bacterial strains and plasmids

Table 2-1 Bacterial strains used in this study.

All of the bacterial strains and plasmids used in this study are listed in Table 2-1 and Table 2-6.

Strain Relevant characteristics Year and place of isolation References/sources

K. pneumoniae

Kpneu_1

O2a:NT

Community-acquired infection

ST441, NT, O2a

Clinical blood isolate

Collected in 2000 from Cairo University Hospital, Egypt

Obtained from Dr Enas Newire (UCL Eastman Dental Institute, and on behalf of US NAMRU-3 and CDC partnership)

Kpneu_63

O3:K14

ST11, K14, O3

Clinical blood isolate

Collected in 2003 from Kafr El Shiek General Hospital, Egypt

Kpneu_64

O3:K14

ST11, K14, O3

Clinical blood isolate

Kpneu_65

O1:K47

ST15, K47, O1

Clinical blood isolate

Kpneu_110

O3:K38

Hospital-acquired infection

ST37, K38, O3

Clinical urinary tract infection (UTI) isolate

Hypermucoviscous phenotype

(Figure 1-2)

Collected in 2009 from Alexandria University Teaching Hospital, Egypt

48 |

Strain Relevant characteristics Year and place of isolation References/sources

C3091

O1:K16

ST14, K16, O1

Laboratory strain, originally isolated from UTI patient

Collected in 1997 from Walter Reed Army Medical Center, Washington

Obtained from Dr Carsten Struve, Statens Serum Institute, Denmark (Oelschlaeger, et al. , 1997; Struve, et al., 2008)

CG43S3

O1:K2

ST86, K2, O1

Laboratory strain, originally isolated from pyogenic liver abscess (PLA)

Collected in 1996 from Chang Gung Hospital, Taoyuan, Taiwan

Obtained from Dr Wang Zhe Chong, National Chiao Tung University, Taiwan (Wang, et al. , 2013)

Kpneu_1∆ompK35 Isogenic ompK35 mutant derived from Kpneu_1

This study

Kpneu_1∆ompK36 Isogenic ompK36 mutant derived from Kpneu_1

This study

C3091∆ompK35 Isogenic ompK35 mutant derived from C3091

This study

C3091∆ompK36 Isogenic ompK36 mutant derived from C3091

This study

Kpneu_1/pKD46-apra Kpneu_1 harbouring pKD46-apra

This study

C3091/pKD46-apra C3091 harbouring pKD46-apra This study

49 |

Strain Relevant characteristics Year and place of isolation References/sources

E. coli

α-select silver efficiency Competent cells Bioline (London, UK)

C. difficile

CD 630 Laboratory strain

ErmR containing two erm(B) genes: erm1(B)

& erm2(B)

Obtained from Dr Haitham Hussain, UCL Eastman Dental Institute (Mullany, et al. , 1990)

* ST – sequence typing based on putative MLST – Multi Locus Sequence Typing as predicted by MLST 1.8

(https://cge.cbs.dtu.dk/services/MLST/)

** K – capsular typing is based on wzi gene

NT- non-typeable (either it is a novel capsule or noncapsulated) (Brisse, et al. , 2013)

*** O – antigen typing is based on wzm and wzt gene

Differentiation between O1 and O2a is based on wbbY gene (Fang, et al. , 2016)

50 |

Table 2-2 Antibiotic susceptibility profile of clinical isolates of K. pneumoniae used in this study.

The determination of antibiotic susceptibility of these isolates was conducted using Kirby-Bauer agar disc diffusion assays. The interpretation

was done according to Clinical & Laboratory Standards Institute (CLSI) Guidelines (CLSI, 2013) . The assay was based on using 6 mm ampicillin

(AM), ceftazidime (CAZ), cefotaxime (CTX), imipenem (IPM), aztreonam (ATM), ampicillin with sulbactam (SAM), ticarcillin with clavulanic acid

(TIM), tetracycline (TE), sulfamethoxazole (SXT), chloramphenicol (C), amikacin (AN), gentamicin (GM), ciprofloxacin (CIP) and nalidixic acid

(NA) discs (Becton Dickinson BD, Sparks, MD, USA). All disc strengths (disc content of the antibiotic) were either (30 μg), (10μg) or (5 μg), based

on the CLSI instruction (CLSI, 2013).

51 |

Table 2-3 Determination of MIC of antibiotics for K. pneumoniae strains used in this study using E-test.

E-test MIC

(µg/ml)

Strain

Gentamicin

Tetracycline

Ciprofloxacin

Cefotaxime

Kpneu_1

6

0.5

0.016

>32

Kpneu_63

>256

32

12

>32

Kpneu_64

>256

32

8

>32

Kpneu_65

>256

>256

>32

2

Kpneu_110

>256

>256

>32

>32

C3091

0.38

0.75

0.012

0.047

CG43S3

0.5

0.5

0.012

0.016

52 |

Table 2-4 Distribution of antibiotic resistance genes in isolates used in this study.

53 |

Table 2-5 Distribution of virulence clusters in isolates used in this study.

54 |

Table 2-6 Plasmids used in this study

Plasmids

Characteristics

Resistance marker

References/sources

pGEM®-T Easy Cloning vector AmpR Promega (UK)

pUC19 Cloning vector AmpR (Yanisch-Perron, et al. , 1985)

pUC19_tet(M) Cloning vector AmpR, TetR Obtained from Dr Supathep Tansirichaiya,

UCL Eastman Dental Institute

pKD46 λ-Red-recombinase plasmid AmpR Obtained from Dr Sophie Helaine,

Imperial College (Datsenko, et al. , 2000)

pKD46-apra Modified pKD46 by replacing amp with

aac(3)-IV

ApraR This study

pSET152 Shuttle vector ApraR Obtained from Dr Haitham Hussain,

UCL Eastman Dental Institute

pLR16T Shuttle vector SpcR Obtained from Dr Haitham Hussain,

UCL Eastman Dental Institute

pGEMT::∆ompK35-1_ermB ΔompK35 mutant cassette derived from

Kpneu_1 cloned into pGEM® T-Easy

AmpR, ErmR This study

55 |

Plasmids

Characteristics

Resistance marker

References/sources

pGEMT::ΔompK36-63_ermB ΔompK36 mutant cassette derived from

Kpneu_63 cloned into pGEM® T-Easy

AmpR, ErmR This study

pGEMT::ΔompK36-65_ermB ΔompK36 mutant cassette derived from

Kpneu_65 cloned into pGEM® T-Easy

AmpR, ErmR This study

pGEMT::ΔompK36-

C3091_ermB

ΔompK36 mutant cassette derived from

C3091 cloned into pGEM® T-Easy

AmpR, ErmR This study

pGEMT::ΔompK35-1_spc Modified pGEMT::∆ompK35-1_ermB

by replacing ermB with aad9

AmpR, SpcR This study

pGEMT::ΔompK36-

C3091_spc

Modified pGEMT::ΔompK36-C3091_ermB

by replacing ermB with aad9

AmpR, SpcR This study

56 |

2.3 Growth condition and storage of bacteria

The bacterial strains used in this study are listed in Table 2-1. All the

K. pneumoniae strains were cultivated on LB agar unless stated otherwise

(with appropriate concentrations of antibiotics when needed) and incubated

aerobically at 37 °C for 16 h statically or on a rotary shaker at 200 rpm when

grown in broth unless stated otherwise. The C. difficile strain was grown in

BHI supplemented with C. difficile selective supplement and 5 % horse blood.

The culture was incubated in a MACS-MG-1000 anaerobic workstation (Don

Whitley Scientific) containing 10 % H2, 10 % CO2 and 80 % N2 at 37 °C for

48 h. All of the bacterial strains were stored at -80 °C in growth medium

containing 20 % (v/v) sterile glycerol.

2.4 Molecular biology methods

2.4.1 Genomic DNA and plasmid extraction

Genomic DNA of K. pneumoniae strains was extracted using Puregene

Yeast/Bact Kit for Gram-negative bacteria (Qiagen, UK), and genomic DNA for

C. difficile 630 was isolated using Puregene Yeast/Bact Kit for Gram-positive

bacteria (Qiagen, UK). All the plasmids used in this study were extracted using

QIAprep Spin Miniprep Kit (Qiagen, UK). All the methods were conducted

according to manufacturer’s protocol with some modifications. The DNA was

eluted using molecular grade water (Sigma-Aldrich) instead of the elution

buffer supplied by the manufacturer.

57 |

2.4.2 Oligo synthesis

The oligonucleotide/primers used in this study are listed in Table 2-7. The primers were synthesised by Sigma-Aldrich.

Table 2-7 Primers used throughout this study.

Primers Sequences References

OmpK35_FL_F

OmpK35_FL_R

5’ CACCAAACTCTCATCAATGG 3’

5’ CTCACGAATACAGAAAAGC 3’

This study

OmpK36_FL_F

OmpK36_FL_R

5’ AGCCGACTGATTAGAAGG 3’

5’ GACAAGAGTATACCAGCGAG 3’

This study

OmpK35_scr_F

OmpK35_scr_R

5’ TATTCTGGCAGTGGTGATCCC 3’

5’ TGGTTATCTTCTTCCGGAGTCAT 3’

This study

OmpK36_scr_F

OmpK36_scr_R

5’ ATCAGGCATGGACTCGTC 3’

5’ ACGTCGTCGGTAGAGATAC 3’

This study

ermBF

ermBR

5’ GAAAAGGTACTCAACCAAATA 3’

5’ AGTAACGGTACTTAAATTGTTTAC 3’

(Ojo, et al. , 2003)

58 |

Primers Sequences References

35_KO1 5’ TTATGTTACGCACTGTTTCGGTGCAT 3’ This study

35_KO1_2 5’ TCACAATTATGTTACGC 3’ This study

35_KO1_3 5’ GATCGCACTTTATGTTCATAACTG 3’ This study

35_KO2 5’ GAAGTGATTACATCTATAAATATTTATTACCCTCATTAAT 3’ This study

35_KO3 5’ TTCTATGAGTCGCTTTTTTATCTGCAGTACACCCCTTCGT 3’ This study

35_KO4 5’ CGGGCGCGTTTATCAGCGATAATC 3’ This study

35_KO4_2 5’ GCAGCGACCAAATTCAGG 3’ This study

35_KO4_3 5’ AGCTCGGATTACTTAC 3’ This study

35_KO5 5’ ATTAATGAGGGTAATAAATATTTATAGATGTAATCACTTC 3’ This study

35_KO6 5’ ACGAAGGGGTGTACTGCAGATAAAAAAGCGACTCATAGAA 3’ This study

36_KO1 5’ CAAAAGGGAAGAATCGCACG 3’ This study

36_KO1_2 5’ TAATTGATTGATTAATAGTCG 3’ This study

36_KO2 5’ GAAGTGATTACATCTATAAAGTTATTAACCCTCTGTTTGT 3’ This study

59 |

Primers Sequences References

36_KO3 5’ TTCTATGAGTCGCTTTTTTAGTTGCAAGCTGCATAACAAA 3’ This study

36_KO4 5’ GCTGTTTATCCGGGCCAAAGCC 3’ This study

36_KO4_2 5’ CCACAGGTTGACCAGCGGC 3’ This study

36_KO4_3 5’ CACGGTAATATCCATCG 3’ This study

36_KO4_4 5’ GCTATCGCCTCGGCATC 3’ This study

36_KO5 5’ ACAAACAGAGGGTTAATAACTTTATAGATGTAATCACTTC 3’ This study

36_KO6 5’ TTTGTTATGCAGCTTGCAACTAAAAAAGCGACTCATAGAA 3’ This study

35D_BB_HindIII_GC_F 5’ GCGGAAGCTTTCTGCAGTACACCCC 3’ This study

35U_BB_HindIII_GC_R 5’ GCCGAAGCTTATTTATTACCCTCAT 3’ This study

36D_BB_HindIII_GC_F 5’ CCGGAAGCTTGTTGCAAGCTGCATA 3’ This study

36U_BB_HindIII_GC_R 5’ GCGCAAGCTTGTTATTAACCCTCTGTTTG 3’ This study

Spect_HindIII_GC_F 5’ CGGGAAGCTTGTGAGGAGGATATATTTGAATAC 3’ This study

Spect_HindIII_GC_R 5’ CGCGAAGCTTGTGTTTCCACCATTTTTTC 3’ This study

60 |

Primers Sequences References

Apra_HindIII_GC_F 5’ GCGGAAGCTTCCAACGTCATCTCGT 3’ This study

Apra_HindIII_GC_R 5’ CCCGAAGCTTGCTATGATCGACTGATG 3’ This study

pKD46_HindIII_GC_F 5’ GCCCAAGCTTACTCTTCCTTTTTCAATATT 3’ This study

pKD46_HindIII_GC_R 5’CCGGAAGCTTCTGTCAGACCAAGTTTACTC 3’ This study

OmpK35_SphI_GC_F_2 5’CGCCGCATGCACATCTTGCAAAGACAGA 3’ This study

OmpK35_SphI_GC_R_2 5’ GCCAGCATGCAAGTCCTGCTTTGAA 3’ This study

OmpK36_SphI_GC_F 5’ CGCCGCATGCGACGGTAATAAATAA 3’ This study

OmpK36_SphI_GC_R 5’ GCCAGCATGCGCCTTTTTGTTATGC 3’ This study

61 |

2.4.3 Polymerase chain reaction (PCR) amplification

Polymerase chain reactions were performed in 20 μl reaction volumes

containing 1 μl of 100 ng/μl of DNA template, 10 μl of 2X BioMix Red (contains

Taq DNA polymerase) (Bioline, UK) or Q5 High Fidelity 2X Master Mix (NEB,

UK) (contains Q5® High-Fidelity DNA polymerase for high fidelity

amplification), 1 μl of 10 μM of each primer and 7 μl of molecular grade water

(Sigma-Aldrich, UK) using a Biometra T3000 Thermocycler (Biometra,

Netherlands) with these following programmes:

A) Standard condition for 2X BioMix Red: initial denaturation at 94 °C for

2 min, and 30 cycles of denaturation step at 94 °C for 2 min (depending on the

template size), annealing step at 40 °C – 71.4 °C (depending on -5 °C of the

melting temperature) for 30 sec, extension step at 72 °C for 2 min to 5 min

(depending on the expected product sizes) and a final extension step at 72 °C

for 2 min to 5 min.

B) Standard condition for Q5 High Fidelity 2X Master Mix: initial

denaturation at 98 °C for 30 sec, and 35 cycles of denaturation step at 98 °C

for 10 sec, annealing step at 40 °C – 71.4 °C (depending on -5 °C of the melting

temperature) for 30 sec, extension step at 72 °C for 45 sec – 2 min (depending

on the expected product sizes) and a final extension step at 72 °C for 2 min.

2.4.4 Agarose gel electrophoresis

DNA was separated using a 1 % (unless stated otherwise) agarose gel by

electrophoresis. One percent (w/v) agarose gels were prepared in 1 X Tris

Acetate EDTA (TAE) buffer, and 0.3 µg/ml of ethidium bromide (EtBr) (Life

Technologies, UK) was added for visualization purposes. One microliter of

62 |

HyperLadderTM 1 kb (Bioline, UK) was loaded as a marker. The DNA sample

was mixed with 5 X loading dye (Qiagen, UK) and loaded onto the prepared

gel. The gel was electrophoresed in 1 X TAE at 90 V for appropriate times

(60 - 100 min). The gels were visualised under UV illumination using an Alpha

Imager (Alpha InnoTech, UK), and the image was captured by Alpha View

version 1.0.1.10 (Alpha InnoTech, UK). The size of the DNA fragments was

determined by comparing the obtained bands with DNA marker.

2.4.5 PCR purification and DNA purification from gels

The separated amplified regions were subjected to gel purification or PCR

purification using a QIAquick Gel Extraction kit (Qiagen, UK) or a QIAquick

PCR Purification kit (Qiagen, UK), respectively, according to the manufactures

recommendations. The elution step was done using molecular grade water

(Sigma-Aldrich) instead of the elution buffer supplied by the manufacturer.

2.4.6 DNA sequencing

All of the purified DNA samples were sequenced using Sanger sequencing

method (ABI 3730XL) performed by a DNA sequencing service, GeneWiz

(formerly known as Beckman Coulter Genomics), UK. Whole genome

sequencing by Illumina MiSeq platform was provided by MicrobesNG

(http://www.microbesng.uk), which is supported by the BBSRC (grant number

BB/L024209/1).

63 |

2.4.7 In silico sequence analysis and protein structural analysis

Table 2-8 Bioinformatic programmes used in this study.

Name Functions used for this study References/sources

Database searches

BLAST

BLASTn

BLASTp

Search for similar DNA/protein sequences on database

Compare two different nucleotide or protein sequences and

identify the homologous regions

(Altschul, et al. , 1990)

Sequence analysis

BioEdit 7.2.0

GeneDoc 2.7

Clustal Omega

SnapGene 3.2.1

(GSL Biotech LLC,

Chicago)

Sequence assembly and chromatogram analysis tool

Multiple DNA/protein sequences alignment tool

Multiple DNA/protein sequences alignment tool

Simulates DNA manipulations, visualises open reading

frames and illustrates DNA/plasmids post manipulations

(Hall, 1999)

(Nicholas, et al. , 1997)

(Sievers, et al. , 2011)

64 |

Name Functions used for this study References/sources

Protein structural analysis

EMBOSS-transeq

RCSB Protein Data Bank

SWISS-MODEL

UCSF Chimera 1.12rc

Predicts deduced amino acid sequences

Retrieves the information regarding the protein sequences

and crystal structures (https://www.rcsb.org/pdb)

Predicts protein structures by comparing the query structure

with the published structure on database

Structural-based alignment software

Illustrates and generates 3D protein structure

Aligns protein structures using the template-based homology

modelling

(Rice, et al. , 2000)

(Berman, et al. , 2000)

(Biasini, et al. , 2014)

(Pettersen, et al. , 2004)

Others

BPROM

ARNold Finding

Terminators

Reverse-complement

MLST 1.8

Predicts promoter sites (Softberry, Inc., NY)

Predicts terminator sites

(http://rna.igmors.u-psud.fr/toolbox/arnold/)

Converts a DNA sequence into its reverse form

(https://www.bioinformatics.org/sms/rev_comp.html)

Predicts K. pneumoniae multi-locus sequence typing (MLST)

(https://cge.cbs.dtu.dk/services/MLST/)

(Solovyev, et al. , 2011)

(Naville, et al. , 2011)

65 |

2.4.8 Restriction endonuclease reactions

Digestion of DNA was conducted using restriction endonucleases (NEB,

Hertfordshire, UK) according to the manufacturer’s protocols. The digested

product was purified using QIAquick PCR Purification Kit prior to ligation steps.

2.4.9 DNA ligation reaction

DNA was ligated using T4 DNA ligase (NEB, Hertfordshire, UK) according to

the manufacturer’s instructions with some modifications. The amount of vector

and insert DNA was added according to the molar ratio of 1:3 vector to insert.

The ligation mixture was incubated at 16 °C for 16 h and the enzyme was heat-

inactivated at 65 °C for 10 min in a Biometra T3000 Thermocycler (Biometra,

Netherlands).

2.4.10 Dephosphorylation reactions

Dephosphorylation of plasmids was carried out to minimise self-ligation of the

vector during ligation step. Seven microlitres of linearised vector (~200 ng/µl)

(to be used for ligation) was incubated with 1 μl (1 U/µl) of Calf Intestine

Alkaline Phosphatase (CIAP), 2 μl of 10X CIAP Reaction buffer (Promega, UK)

and 10 µl of molecular grade water and incubated at 37 °C for 30 min in a

water bath. After 30 min, 1 μl of CIAP was added to the mixture and re-

incubated for an additional 30 min. The dephosphorylated vector was purified

using QIAquick PCR Purification Kit prior to ligation.

66 |

2.4.11 Transformation of plasmids into E. coli -select silver

efficiency competent cells

E. coli transformations were conducted using E. coli α-select silver efficiency

competent cells (Bioline, London, UK). Five microlitres of ~100 ng/µl DNA

(ligation mix) was added to 50 µl of thawed competent cells, and the mixture

was mixed gently by tapping the bottom of the tube. The mixture was

incubated on ice for 30 min and heat shocked in a 42 °C water bath for

45 sec. The cells were transferred quickly to an ice bucket and incubated for

2 min. Nine hundred microliters of SOC broth (NEB, UK) was added to the

mixture and incubated at 37 °C for 1 h with shaking at 200 rpm. Two hundred

microlitres of cells were spread onto an LB agar supplemented with an

appropriate antibiotic for selection and LB agar without any supplement as a

control. All the plates were then incubated for at 37 °C for 16 h.

2.4.12 Blue/white screening

The pGEMT-easy plasmids containing a DNA fragment of interest were

transformed into E. coli α-select silver efficiency competent cells (Bioline, UK).

Putative transformants were selected on an LB agar containing 100 µg/ml

ampicillin, 40 µg/ml 5-bromo-4-chloro-indolyl-β-D-galactopyranoside (X-gal),

100 mM isopropyl-β-D-thiogalactopyranoside (IPTG), and additional

antibiotics (if any). Plates were incubated at 37 °C for 16 h. White colonies

indicated the presence of recombinant plasmids containing the insert of

interest, while blue colonies indicated self-ligated plasmids. White colonies

were then selected for further confirmation by colony PCR or plasmid

extraction using QIAprep Spin Miniprep Kit (Qiagen, UK).

67 |

2.5 Statistical analysis

All data were presented as mean ± standard deviations (SD). Nonparametric

Mann-Whitney test was used to analyse differences between two different

groups. Nonparametric Kruskal-Wallis with post-hoc Dunn’s test was used to

compare differences between more than two groups with one parameter. Both

of these analyses were used for non-normally distributed data. Analysis of

variance, one-way ANOVA was carried out to compare differences between

more than two groups with one parameter, and two-way ANOVA was carried

out to compare differences between more than two groups with more than one

parameters. Both of the ANOVAs were corrected using Dunnett’s or

Bonferroni’s post hoc tests depending on the experiments. Dunnett’s post hoc

test was conducted to compare one group (control group) with the others, and

Bonferroni’s post hoc test was chosen for multiple comparison between the

groups. A difference was considered significant if p<0.05 (*). All statistical

analyses were conducted using GraphPad Prism version 6.0 (GraphPad

Software, San Diego, CA, USA).

68 |

Chapter 3.0

Investigation into the effect of

antibiotic exposure on

Klebsiella pneumoniae

biofilm formation

69 |

3.0 Investigation into the effect of antibiotic exposure

on K. pneumoniae biofilm formation

3.1 Introduction

3.1.1 In vitro biofilm models

Surface-adhered bacterial communities or biofilms play an important role in

nosocomial infections. The ability of bacteria to form a biofilm on medical

devices leads to biofilm-associated infections, which are recalcitrant toward

antibiotics (Lebeaux, et al., 2014; Lynch, et al. , 2008). Hence, it is very

important to develop experimental models that mimic real environmental

conditions for a better understanding of the mechanisms of biofilm formation.

Although these models only approximate real conditions, findings obtained

from the experimental models, can still provide us with information on the

phenotypic and genotypic differences between motile and sessile bacterial

communities.

There are several models/apparatuses that researchers use to examine

biofilms. Among them are the microtitre plate (O'Toole, 2011), the Calgary

device (Ceri, et al. , 1999), the Biofilm Ring test (Olivares, et al. , 2016), the

modified Robbins device (McCoy, et al. , 1981), flow chambers (Wolfaardt, et

al. , 1994), the constant depth film fermenter (CDFF) (Peters, et al. , 1987), the

drip flow biofilm reactor, the rotary biofilm device (BioSurface Technologies

Corporation) (Schwartz, et al. , 2010) and microfluidic biochips (Richter, et al.

, 2007). Biofilm formation assays using microtitre plates are the most common

model because this method is cost-effective, easy to handle, reproducible,

suitable for high-throughput screening, and less complicated in comparison to

70 |

the other methods (Azeredo, et al. , 2017; Lebeaux, et al. , 2013). Despite

these advantages, biofilm formation assays in microtitre plates do have some

pitfalls. For example, the biofilm yield obtained from wells of a microtitre plate

is very limited and it is normally inadequate to be used for gene or protein

expression studies specifically in 96-well microtitre plates.

3.1.2 The effects of antibiotic exposure on biofilm formation

As a rule of thumb, matrix-encased biofilms are more resistant to antibiotics

than their planktonic counterpart (Davies, 2003). Exposure to low dose

antibiotics has been suggested to be a contributing factor to stimulation of

biofilm formation (Kaplan, 2011) and this phenomenon might have clinical

relevance. Several studies have proven that exposure to sub-inhibitory

concentrations of antibiotics in vitro can affect various bacterial activities

including: increase in virulence (Geisinger, et al. , 2015), increase in random

mutagenesis (Foster, 2007), and stimulation of biofilm formation (Kaplan,

2011). Formation of biofilm upon exposure to low-level antibiotics also is

proposed to be one of the defensive mechanisms and adaptation responses

to the presence of antibiotics in the environment (Hoffman, et al. , 2005).

Many studies have investigated the mechanisms of antibiotic-induced biofilm

formation upon exposure of bacteria in different growth states to sub-MICs of

antibiotics in several pathogens, for example in Staphylococcus aureus

(Bisognano, et al. , 1997; Blickwede, et al. , 2004), Staphylococcus epidermidis

(Rachid, et al. , 2000), Pseudomonas aeruginosa (Bagge, et al. , 2004;

Hoffman, et al., 2005; Linares, et al. , 2006), Escherichia coli (Hoffman, et al.,

2005), and Acinetobacter baumanni (Nucleo, et al. , 2009).

71 |

As an example, one-quarter of the MIC of ciprofloxacin was shown to induce

the expression of fibronectin adhesins such as fibronectin binding proteins

(FnBPs) in fluoroquinolone resistant S. aureus strains upon growth in Mueller-

Hinton (MH) broth. This subsequently resulted in increased adhesion of

fluoroquinolone resistant strains on fibronectin-coated coverslips compared to

the strains which were grown in antibiotic-free medium (Bisognano, et al.,

1997).

Rachid, et al. (2000) also showed that exposure of S. epidermidis cultures to

sub-inhibitory concentrations of tetracycline or streptogramin could also

promote the production of exopolysaccharides in S. epidermidis. The

transcription of the ica operon, which contains genes that encode for proteins

involved in the synthesis of polysaccharide intercellular adhesin (PIA) was

observed to be induced after treatment with a very low dosage of tetracycline

or combination of streptogramin compounds, (quinupristin-dalfopristin) ranging

from 1/70 to 1/2 X MIC. This subsequently resulted in an enhancement of

biofilm formation by S. epidermidis.

Exposing P. aeruginosa biofilms to no more than 1/10 X MIC of imipenem

(0.2 µg/ml) was also shown to upregulate the genes that encode for alginate

biosynthesis. Enhancing production of alginate stimulated biofilm formation by

this pathogen (Bagge, et al., 2004). Besides imipenem, biofilms of

P. aeruginosa were also found to be induced upon exposure to 1 µg/ml of

tobramycin which is equal to less than 0.3 X MIC of tobramycin for this strain

(Hoffman, et al., 2005). The aminoglycoside response regulator (arr) that was

predicted to encode an inner-membrane phosphodiesterase that catalysed

72 |

hydrolysis of c-di-GMP was believed to be involved in the biofilm induction

upon exposure to tobramycin. This study proposed that tobramycin either

directly or indirectly enhanced the enzymatic activity of Arr cytoplasmic EAL

domain, which resulted in increased biofilm formation by decreasing the c-di-

GMP level in the cytoplasm.

This finding, however, is in contradiction to other studies which have shown

that higher levels of c-di-GMP are important in bacterial biofilm formation. As

an example, a previous study showed that intracellular levels of c-di-GMP are

involved in regulation of EPS production by P. aeruginosa. Increased levels

of c-di-GMP were shown to correlate with increased biofilm formation and

decreased levels of c-di-GMP resulted in reduction of biofilm production by

P. aeruginosa. Low levels of c-di-GMP were also suggested to prevent the

initiation step of biofilm formation (Hickman, et al. , 2005).

Exposure of E. coli to sub-inhibitory concentrations of translation inhibitor

antibiotics (erythromycin, chloramphenicol, tetracycline, and streptomycin)

were shown to induce biofilm formation by upregulating poly-β-1, 6-N-acetyl-

glucosamine (PNAG) through two biosynthetic proteins, PgaA and PgaD.

Enhanced production of PgaA and PgaD leads to increase PNAG and

enhances the formation of biofilm. PNAG production is controlled by bacterial

intracellular signaling molecules guanosine-bis 3’ 5’ diphosphate (ppGpp)

and c-di-GMP (Boehm, et al. , 2009). Owing to the fact that K. pneumoniae

also has high degree of amino acid sequence identity (see Chapter 4, section

4.3.1.2) and machinery for the synthesis and secretion of PNAG with E. coli,

(Chen, et al. , 2014), exposure of K. pneumoniae to sub-inhibitory

73 |

concentrations of protein translation inhibitor, gentamicin was also

hypothesised to induce biofilm formation by K. pneumoniae.

3.1.3 Justification of the experimental design

In this study, two different antibiotics with two different modes of action: a

protein translation inhibitor and a DNA replication inhibitor, were chosen to be

tested for possible effects of sub-MICs exposure of these antibiotics on biofilm

formation by K. pneumoniae. The antibiotics were gentamicin, and

ciprofloxacin. The concentrations of antibiotics used in this study were sub-

MICs and these included the breakpoint and sub-breakpoint concentrations of

the particular antibiotics, for K. pneumoniae. These were based on the

European Committee on Antimicrobial Susceptibility Testing (EUCAST) (2014)

Guidelines when this study was conducted.

Gentamicin and ciprofloxacin were selected since they are normally used to

treat K. pneumoniae infections and are also commonly used as antimicrobial

prophylaxis. Gentamicin is commonly used for perioperative antimicrobial

prophylaxis for urological surgeries such as ureterostomy and it has also been

used prior to catheter change (Parkinson, et al. , 2016). Whilst, ciprofloxacin

has been one of the most widely used antibiotics as prophylaxis for UTIs

(Enzler, et al. , 2011; Hickerson, et al. , 2006).

74 |

3.1.4 Modes of action and mechanisms of resistance of

antibiotics

3.1.4.1 Gentamicin

Gentamicin is an aminoglycoside. The aminoglycosides family includes

amikacin, tobramycin, streptomycin, kanamycin, and neomycin. Gentamicin

acts by binding irreversibly to the 16S rRNA subunit of the bacterial 30S

ribosomal unit which results in misreading of the amino acid codons that have

already been initiated (Shoji, et al. , 2009; Tai, et al. , 1979) and interrupting

the translocation of peptidyl-tRNA from A-site on the 16S rRNA to the P-site

(Figure 3-1) (Feldman, et al. , 2010; Kotra, et al. , 2000). This subsequently

inhibits protein synthesis (Wilson, 2014).

Figure 3-1 The modes of action of gentamicin.

Gentamicin binds to the 30S ribosomal subunit and leads to misreading of codon and

also interrupts the translation process.

75 |

There are several aminoglycoside resistance mechanisms including pumping

out aminoglycosides by active efflux system, alteration of the ribosome target

sites, (Magnet, et al. , 2005) and inactivation of aminoglycosides by

aminoglycosides-modifying enzymes (AMEs). AcrD in E. coli (Rosenberg, et

al. , 2000) and MexXY-OprM in P. aeruginosa (Hocquet, et al. , 2003) are

amongst the most effective efflux pumps which are necessary for resistance

against aminoglycosides. Alteration of the ribosome target sites can prevent

the binding of the aminoglycosides to the ribosome. For example,

methyltransferases which are known as 16S rRNA methylases can modify 16S

ribosomal RNA which acts as the aminoglycosides binding site. These

methyltransferases include aminoglycoside resistance methyltransferase A

which encodes by armA (Galimand, et al. , 2003). AMEs are the most

important resistance mechanisms found in K. pneumoniae (Ramirez, et al.,

2010) and are typically found on MGEs such as plasmids or transposons. The

modifying enzymes that are involved in gentamicin inactivation include

aminoglycosides acetyltransferases (AAC), aminoglycosides acetylnucleotidyl

transferases (ANT), and aminoglycosides acetylphosphotransferases (APH)

(Figure 3-2) (Mingeot-Leclercq, et al. , 1999; Tangy, et al. , 1985).

76 |

Modifying enzymes and the targets

are as labelled.

Figure 3-2 Chemical structure of gentamicin.

The gentamicin resistance genes, aac(3)-II and aac(6’)-Ib, aac(6’)-IV,

ant(2’’)-Ia, and aph(3’)-Ia, are known to be among the most prevalent

determinants of gentamicin resistance found in K. pneumoniae (Almaghrabi,

et al. , 2014; Naparstek, et al. , 2014; Ruiz, et al. , 2012).

77 |

3.1.4.2 Ciprofloxacin

Figure 3-3 Chemical structure of ciprofloxacin.

Ciprofloxacin is a bactericidal fluoroquinolone (Figure 3-3). Quinolones disturb

DNA replication by inhibiting two essential type II topoisomerases: DNA gyrase

and topoisomerase IV activities which are essential for bacterial DNA

replication. DNA gyrase is a tetrameric enzyme consisting of two A (GyrA) and

two B (GyrB) subunits (Redgrave, et al. ). The GyrA subunit plays a role in

DNA breakage and religation, and GyrB is involved in ATPase binding and

hydrolysis (Qi, et al. , 2002). Topoisomerase IV has a similar structure as DNA

gyrase. It is made of two A subunits and two B subunits which are encoded

by parC and parE, respectively (Kato, et al. , 1990). This enzyme has two

specific functions: to relax the positively supercoiled strands together with DNA

gyrase and to unlink newly replicated daughter chromosomes (Fabrega, et al.

, 2009).

78 |

Mutations or alterations in specific domains for both gyrA and gyrB, and

secondary drug targets ParC and ParE, respectively (M. E. Jones, et al. ,

2000), and also, protection of DNA gyrase and topoisomerase IV by Qnr

protein are among the most common resistance mechanisms to ciprofloxacin

(Tran, et al. , 2002).

The first plasmid (pMG252) mediated harbouring quinolone resistance gene,

qnrA was found in a clinical K. pneumoniae strain isolated in 1994 at the

University of Alabama at Birmingham Medical Centre, USA (Jacoby, et al. ,

2003; Martinez-Martinez, et al. , 1998). qnrA encodes for the 218-amino-acid

protein, QnrA, which belongs to pentapeptide repeat family. This protein

protects the quinolone targets, DNA gyrase and topoisomerase IV from

inhibitory activity of the antibiotics. (Tran, et al. , 2005).

Another additional quinolone resistance mechanism which is common in

Enterobacteriaceae and also K. pneumoniae (Ruiz, et al., 2012) is mediated

by an aminoglycosides-modifying enzyme, AAC(6)-Ib-cr, which can cause

resistance to aminoglycosides and fluoroquinolones simultaneously. This

enzyme is a variant of aminoglycoside acetyltransferase with two amino acid

differences, Trp102Arg and Asp179Tyr and is thought to be more prevalent

than the Qnr protein. The alterations allow it to inactivate ciprofloxacin through

the acetylation of its piperazinyl substituent (Robicsek, et al. , 2005;

Strahilevitz, et al. , 2009).

79 |

3.1.5 Aims of this chapter

To determine the effect of exposure of: i) cultures during bacterial growth until

mid-log phase, ii) cultures during biofilm formation, and iii) mature biofilms of

K. pneumoniae to sub-MICs of antibiotics under shaking condition. The

antibiotics, gentamicin and ciprofloxacin were tested at sub-MICs. The usage

of these concentrations can be considered clinically relevant since the sub-

MICs specifically breakpoint concentration is correlated to achievable serum

drug levels using standard dosing for treatment (Levison, et al. , 2009).

Bacteria are also possibly exposed to low antibiotic concentrations several

hours at the beginning, between doses and end of antibiotherapy (Craig,

1998), or after continuous low-dose antibiotics prophylaxis which is normally

used to treat UTI and surgical wounds (Parkinson, et al., 2016). In this study,

we wanted to determine whether exposure of these different condition of

culture and mature biofilm to these antibiotics would increase biofilm

formation by K. pneumoniae.

80 |

3.2 Materials and methods

3.2.1 Biofilm formation under different growth conditions and

in different media

A 16 h old culture of each isolate was adjusted to an optical density at

600 nm (OD600) of 0.05 by adding 100 µl of the culture into 10 ml of TSB.

The cultures were incubated at 37 °C with agitation at 200 rpm (SANYO MIR-

22RU, Japan) until they reached mid-log phase (approximately after 2 h).

Bacteria at mid-log phase (see Chapter 4, section 4.3.2) were diluted to a

OD600 of ~1.0 (~109 cfu/ml) and 20 μl of the suspension was added into

wells of a flat-bottom 96-well polystyrene tissue culture treated plate

(Sarstedt, Germany) containing 180 μl of fresh TSB or TSB + 2 % glucose.

TSB only (without bacterial culture) was included in each test as a blank

control. The cultures were then incubated at 37 °C for 24 h under static

conditions or with shaking at 200 rpm on an orbital shaker (Luckham, UK).

The biofilms were then quantified using a crystal violet assays.

3.2.2 Pre-exposure of cultures to antibiotics prior to biofilm

formation

To examine the effect of pre-exposure of cultures to sub-MICs of antibiotics

on biofilm formation, a 16 h old culture of each isolate was adjusted to OD600

of 0.05 by adding 100 µl of the culture into 10 ml of TSB. The antibiotics,

gentamicin (4 μg/ml, 2 μg/ml or 1 μg/ml) or ciprofloxacin (1 μg/ml, 0.5 μg/ml

or 0.25 μg/ml) were added into the suspension. The tube with TSB inoculated

with culture only (without any antibiotic) acted as a control. The cultures were

incubated until the bacteria reached mid-log phase and they were diluted to

OD600 of ~1.0 (~109 cfu/ml). Twenty microlitres of the suspension were

81 |

inoculated into the wells of a flat-bottom 96-well polystyrene tissue culture

treated plate (Sarstedt, Germany) containing 180 μl of fresh TSB. The cover

was replaced with a microtitre plate breathable film cover (Anachem, UK) to

minimise evaporation and ‘edge effects’. The breathable film covers were

used for all subsequent biofilm formation assays in this study. The plate was

incubated at 37 °C for 24 h under shaking conditions at 200 rpm on an orbital

shaker (Luckham, UK) and the biofilms were then quantified using a crystal

violet assays.

3.2.3 Exposure of mid-log phase cultures to antibiotics during

biofilm formation

To investigate the effect of exposure of mid-log phase cultures to sub-MICs

of antibiotics during biofilm formation, 180 μl of TSB only or TSB with

respective concentrations of antibiotics [(4 μg/ml, 2 μg/ml or 1 μg/ml) for

gentamicin, or (1 μg/ml, 0.5 μg/ml or 0.25 μg/ml) for ciprofloxacin)] were

added to the wells of microtiter plates. Twenty microlitres of diluted mid-log

phase cultures (OD600 of ~1.0 (~109 cfu/ml) were used to inoculate wells

containing TSB only or TSB with antibiotics. The TSB only wells were

labelled as control groups. The plate was incubated as mentioned before

and biofilms were then quantified using crystal violet.

82 |

3.2.4 Exposure of mature biofilms to antibiotics

Biofilms were grown at 37 °C for 24 h as previously mentioned, and after

24 h incubation, planktonic cells were removed from the wells containing 24 h

old biofilms. The wells were washed three times carefully (to avoid disturbance

to the biofilm) with sterile phosphate buffer saline (PBS). Fresh TSB medium

or TSB with respective concentrations of antibiotics [(4 μg/ml, 2 μg/ml or 1

μg/ml) for gentamicin, or (1 μg/ml, 0.5 μg/ml or 0.25 μg/ml) for ciprofloxacin)]

were added into the wells containing 24 h old biofilms. The TSB only wells

were labelled as controls. The 24 h old biofilms in fresh medium were then re-

incubated at 37 °C for 24 h under shaking conditions at 200 rpm. After

24 h incubation, the absorbance of planktonic cells (represents the bacterial

growth) were measured using a microplate reader (Dynex MRX TC II, MTX

Labsystem, USA) prior to quantification of biofilms using crystal violet.

3.2.5 Crystal violet quantification of biofilm formation

Planktonic cells were removed from the incubated plate and the wells were

washed three times with 200 μl of PBS to remove the non-adherent cells. The

plate was then air-dried at 37 °C for 10 min and the biofilms were stained with

300 μl of 0.005 % of crystal violet (CV) (Pro-Lab Diagnostics, UK) and

incubated at room temperature for 30 min. The CV was discarded and the

plate was washed twice with 300 μl of PBS to remove the unbound CV. Three

hundred microlitres of ethanol (90 %) was added into the wells to solubilise the

stain and the plate was incubated for 15 min. The absorbance (OD590) was

measured using a microplate reader (Dynex MRX TC II, MTX Labsystem,

USA).

83 |

Figure 3-4 Simplified diagram of methodology used in this chapter.

84 |

3.3 Results

3.3.1 Optimisation of growth conditions and growth media for

biofilm formation assays

The main objective of this chapter was to investigate the effect of antibiotic

exposure on K. pneumoniae biofilm formation. To determine the most

suitable method to be used for the subsequent biofilm formation assays in

further investigations, several factors were optimised to ensure reproducibility

of the assays during the study. Among those factors were the optimal growth

conditions and a suitable medium to grow biofilms using flat-bottom 96-well

polystyrene tissue culture treated plates. To determine the optimum

conditions that could be used to grow biofilms and whether glucose had an

effect on biofilm formation of these six K. pneumoniae strains, preliminary

assays were conducted by inoculating 20 µl of mid-log phase cultures into

180 µl of TSB or TSB + 2 % glucose in 96-well microtitre plate and incubating

the plate at 37 °C for 24 h under static and shaking conditions. Subsequently,

the capacity to form biofilms in two different conditions and two different

media was determined and compared. For the purpose of explanation and

comparison, the graphs for optimisation of the growth conditions (shaking

and static) (Figure 3-5) and growth media (TSB or TSB + 2 % glucose)

(Figure 3-6) were plotted separately, even though, both of these experiments

were conducted at the same time.

85 |

Figure 3-5 shows the biofilm formation by K. pneumoniae after incubation at

37 °C for 24 h under shaking and static conditions in two different media, TSB

and TSB + 2 % glucose. Kpneu_1 was observed to form more biofilms in

TSB under shaking conditions compared to static conditions in TSB. While

Kpneu_110 was observed to form more biofilms under shaking conditions in

both media. However, no significant difference was observed for strains

Kpneu_63, Kpneu_64, Kpneu_65, and C3091 cultivated in TSB or TSB +

2 % glucose when incubated under either static and shaking conditions.

86 |

Figure 3-5 Biofilm formation by K. pneumoniae under different growth

conditions.

Biofilms were grown in A) TSB and B) TSB + 2 % glucose at 37 °C for

24 h in 96-well microtitre plates under static and shaking conditions. Bars show

average absorbance readings from three independent experiments. The data was

analysed by two-way ANOVA and Bonferroni’s post hoc test. Error bars indicate the

standard deviation of nine replicates. Statistical significance was indicated by

* represents p < 0.05, *** represents p < 0.001, and **** represents p < 0.0001.

87 |

Figure 3-6 shows that no significant difference was found between biofilm

formation in TSB compared to TSB + 2 % glucose under static or shaking

conditions. The presence of 2 % glucose in TSB was shown to have no effect

on biofilm formation by any of the strains. Given this finding, all of the

subsequent biofilm assays were carried out under shaking conditions in TSB

only.

88 |

Figure 3-6 Biofilm formation by K. pneumoniae in two different media.

Biofilms were grown under A) static and B) shaking conditions in 96-well microtitre

plates at 37 °C for 24 h in two different media, TSB and TSB + 2 % glucose. Bars

show average absorbance readings for three independent experiments. The data

was analysed by two-way ANOVA and Bonferroni’s post hoc test. Error bars indicate

the standard deviation of nine replicates. No significant difference was observed for

all the strains between the biofilm formation in TSB and TSB + 2 % glucose.

89 |

3.3.2 Pre-exposure of cultures to gentamicin or ciprofloxacin

prior to biofilm formation

In order to determine the effect of antibiotic pre-exposure on

biofilm formation by K. pneumoniae, the K. pneumoniae culture was

pre-exposed to antibiotics: gentamicin or ciprofloxacin during 2 h incubation

until it reached mid-log phase. Figure 3-7 shows the result of pre-exposure to

A) gentamicin and B) ciprofloxacin on biofilm formation by K. pneumoniae.

Strains Kpneu_1, Kpneu_64, and Kpneu_110 showed a significant reduction

of biofilm formation upon exposure to 1 µg/ml and 2 µg/ml gentamicin

compared to the control group. Whereas, no significant difference was seen

upon pre-exposure of strains Kpneu_63 and Kpneu_65 to gentamicin. Strain

C3091 was not included in this pre-exposure to gentamicin biofilm assays

since it was susceptible to gentamicin and addition of gentamicin in the

growth medium for incubation until mid-log phase killed the cells. Figure 3-7

shows that strain Kpneu_65 had a significant reduction in biofilm formation

upon pre-exposure to 0.25 µg/ml, 0.5 µg/ml and 1 µg/ml of ciprofloxacin while

Kpneu_110 showed a significant reduction upon pre-exposure to 0.5 µg/ml

and 1 µg/ml of ciprofloxacin compared to controls. Two susceptible strains:

strains Kpneu_1 and C3091 were not included in this pre-exposure to

ciprofloxacin biofilm assays since they were susceptible to ciprofloxacin and

did not survive during the incubation until mid-log phase with addition of of

ciprofloxacin in growth medium. The graph in Figure 3-7 shows that pre-

exposure of K. pneumoniae cultures during 2 h of incubation until they

reached mid-log to sub-MICs of gentamicin or ciprofloxacin did not increase

the biofilm mass of all the strains used in this study.

90 |

Figure 3-7 Effect of pre-exposure of cultures to gentamicin or ciprofloxacin

prior to biofilm formation by K. pneumoniae.

The effect of pre-exposure of cultures to A) gentamicin or B) ciprofloxacin during 2 h

incubation until it reached mid-log phase, on biofilm formation by

K. pneumoniae. The biofilms were grown in 96-well microtitre plates at 37 °C for

24 h in TSB under shaking conditions. Bars show average absorbance readings from

three independent experiments. The data was analysed by one-way ANOVA and

Dunnett’s post-hoc test. Error bars indicate standard deviation of nine replicates.

Significant reduction was observed after pre-exposure of strains Kpneu_1, Kpneu_64

and Kpneu_110 to 2 μg/ml gentamicin compared to controls. Whilst strain Kpneu_65

showed significant reduction after pre-exposed to 0.25 μg/ml ciprofloxacin and both

Kpneu_65 and Kpneu_110 showed reduction upon pre-exposed to 0.5 μg/ml and

1.0 μg/ml ciprofloxacin compared to controls. Statistical significance is indicated by

* represents p < 0.05, ** represents p < 0.01, *** represents p < 0.001 and ****

represents p < 0.0001.

91 |

3.3.3 Exposure of mid-log phase cultures to gentamicin or

ciprofloxacin during 24 h biofilm formation

The effect of antibiotic exposure during 24 h biofilm formation was examined

by inoculating 20 µl of mid-log phase cultures into growth medium (with and

without antibiotics) and incubating at 37 °C for 24 h under shaking conditions.

The data in Figure 3-8 shows that gentamicin susceptible strain, C3091 (MIC

0.38 µg/ml) showed a significant reduction in biofilm formation upon exposure

to 2 µg/ml, 4 µg/ml or 8 µg/ml of gentamicin in comparison to control group.

Apart from C3091, all strains showed no significant difference in biofilm

formation compared to controls upon incubation for 24 h with gentamicin.

Kpneu_1 and C3091 exhibited a highly significant reduction in biofilm

formation upon exposure to ciprofloxacin compared to controls and this could

be because both of these strains were susceptible to ciprofloxacin as

previously mentioned (MIC 0.016 µg/ml and 0.012 µg/ml respectively). Strain

Kpneu_64 had a significant reduction in biofilm formation upon exposure to

0.25 µg/ml and 0.5 µg/ml of ciprofloxacin, while Kpneu_63 had a significant

reduction in biofilm formation upon exposure to 0.25 µg/ml, 0.5 µg/ml, and

1 µg/ml of ciprofloxacin compared to controls. Taken together, exposure of

mid-log phase culture to sub-MICs of gentamicin or ciprofloxacin during 24 h

biofilm formation did not increase the biofilm formation by any of the

K. pneumoniae strains used in this study.

92 |

Figure 3-8 Effect of antibiotic exposure during 24 h incubation on biofilm

formation by K. pneumoniae.

The effect of A) gentamicin or B) ciprofloxacin exposure during 24 h biofilm formation

on biofilm formation by K. pneumoniae. The biofilms were grown in 96-well microtitre

plates at 37 °C for 24 h in TSB under shaking conditions. Bars show average

absorbance readings from three independent experiments. The data was analysed

by one-way ANOVA and Dunnett’s post-hoc test. Error bars indicate the standard

deviation of nine replicates. C3091 showed a significant reduction in biofilm formation

upon exposure to 1 μg/ml, 2 μg/ml, or 4 μg/ml of gentamicin in comparison to controls.

Highly significant reduction can be seen after exposure to sub-MICs of ciprofloxacin

on biofilm of Kpneu_1 and C3091 during 24 h biofilm formation. Strain Kpneu_63

showed a significant reduction in biofilm, after exposure to 0.25 μg/ml, 0.5 μg/ml, and

1 μg/ml of ciprofloxacin whilst strain Kpneu_64 had a significant reduction in biofilm

formation upon exposure to0.25 μg/ml and 0.5 μg/ml of ciprofloxacin. Statistical

significance is indicated by * represents p < 0.05, ** represents p < 0.01,

*** represents p < 0.001 and **** represents p < 0.0001.

93 |

3.3.4 Exposure of mature biofilms to antibiotics

The effects of exposure to antibiotics on the growth of K. pneumoniae are

shown in Figure 3-9 while the effects of gentamicin or ciprofloxacin exposure

to mature K. pneumoniae biofilms are presented in Figure 3-10. To evaluate

the effect of antibiotics exposure to pre-formed K. pneumoniae biofilms, the

biofilms were grown as previously mentioned at 37 °C for 24 h under shaking

conditions. The 24 h old biofilms were then exposed to sub-MICs of antibiotics.

Figure 3-10 shows the absorbance of planktonic bacteria upon re-incubation

of the mature biofilms in a fresh TSB in the presence of gentamicin or

ciprofloxacin at sub-MICs. The growth of strain C3091, however, was

observed to be reduced upon exposure to 2 µg/ml of gentamicin compared to

control. Planktonic cells of ciprofloxacin susceptible strains, Kpneu_1 and

C3091 were also observed to be significantly reduced upon incubation in TSB

in the presence of 0.25 µg/ml, 0.5 µg/ml and 1 µg/ml of ciprofloxacin compared

to controls. Apart from these two strains, a significant reduction in growth was

also seen for Kpneu_64 upon exposure to 1 µg/ml of ciprofloxacin compared

to control.

94 |

Figure 3-9 Effect of antibiotic exposure on growth of K. pneumoniae.

The effect of A) gentamicin and B) ciprofloxacin exposure on growth of

K. pneumoniae (derived from the mature biofilm experimental plate) incubated at

37 °C for 24 h in TSB under shaking conditions. Bars show average absorbance

readings from three independent experiments. The data was analysed by one-way

ANOVA and Dunnett’s post hoc test. Error bars indicate the standard deviation of

nine replicates. The growth of gentamicin-susceptible strain C3091 showed a

significant reduction upon exposure to 1 μg/ml, 2 μg/ml and 4 μg/ml of gentamicin in

comparison to control upon 24 h incubation. Whilst, the growth of ciprofloxacin-

susceptible strains, strain Kpneu_1 and C3091 also showed a significant reduction

upon exposure to 0.25 µg/ml, 0.5 µg/ml and 1 µg/ml of ciprofloxacin compared to the

control groups. Statistical significance was indicated by ** represents p < 0.01,

*** represents p < 0.001 and, **** represents p < 0.0001.

95 |

Figure 3-10 shows the effect of antibiotic exposure on mature biofilms of

K. pneumoniae. No significant difference was found after exposure of mature

biofilm to gentamicin and ciprofloxacin for all of the strains apart from C3091.

Strain C3091 was a gentamicin and ciprofloxacin-susceptible strain and the

growth of this strain in TSB was disturbed with addition of these antibiotics

compared to the control (see Figure 3-9).

Mature biofilms of C3091 were slightly reduced upon exposure to 2 µg/ml of

gentamicin and upon exposure to 0.25 µg/ml, 0.5 µg/ml and 1 µg/ml of

ciprofloxacin compared to controls. No significant differences in the amount

of mature biofilm were seen for ciprofloxacin-susceptible strain Kpneu_1, upon

exposure to ciprofloxacin compared to controls, even though it was clearly

shown that the growth of this strain was completely inhibited with the addition

of ciprofloxacin in the growth medium. No reduction of mature biofilms was

seen for Kpneu_64 although its growth was significantly reduced upon

exposure to 1 µg/ml of ciprofloxacin compared to controls.

96 |

Figure 3-10 Effect of antibiotic exposure on mature biofilms of K. pneumoniae.

The effect of exposure of mature biofilms of K. pneumoniae to A) gentamicin and B)

ciprofloxacin upon cultivation in 96-well microtitre plates at 37 °C for 24 h in TSB under

shaking conditions. Bars show average absorbance readings from three independent

experiments. The data was analysed by one-way ANOVA and Dunnett’s post hoc

test. Error bars indicate the standard deviation of nine replicates. Strain C3091

showed a significant reduction upon exposure to 2 μg/ml of gentamicin and also upon

exposure to 0.25 μg/ml, 0.5 μg/ml and 1.0 μg/ml ciprofloxacin in comparison to

controls. No significant difference was seen for all the other strains upon exposure of

mature biofilm to gentamicin and ciprofloxacin. Statistical significance was indicated

by * represents p < 0.05, ** represents p < 0.01 and, *** represents p < 0.001.

97 |

3.4 Discussion

The initial part of this chapter described an optimisation of the biofilm

formation assays that would be used for subsequent experiments involving

any biofilm mass quantification. Biofilm formation assays in our study were

performed on 24 h biofilms. This was based on our data, that showed the

absorbance readings of crystal violet that was used for staining of biofilm

mass post 24 h and 48 h was almost similar (Figure 3-7 and Figure 3-8, and

Figure 3-10) for all of the strains. Hence, all of the biofilm assays were

conducted on 24 h old biofilms for all subsequent experiments.

The first optimisation was cultivation of K. pneumoniae under static and

shaking conditions. K. pneumoniae was observed to form biofilms more

efficiently under shaking conditions compared to static conditions for strains

Kpneu_1 and strain Kpneu_110. Shaking conditions provide shear stress

(Moreira, et al. , 2013); by means of the friction force between liquid in wells

and wall surfaces (Liu, et al. , 2002; Saur, et al. , 2017). Our findings showed

that formation of biofilm was significantly higher for strain Kpneu_1 and

Kpneu_110 under shaking conditions (Figure 3-5).

The ability of bacteria to form more biofilms under shaking conditions in

microtitre plates might be because of the existence of shear stress they

experience in the wells. Also shaking conditions could bring more bacterial

cells into contact with the well surfaces (Stoodley, et al. , 1998) and this

results in increased cell attachment on the wells. Shaking conditions could

also result in higher cell density (Paris, et al. , 2007) provided by the higher

98 |

oxygen and substrate mass transfer from the medium to biofilm cells upon

incubation under shaking conditions (Moreira, et al., 2013).

The mechanical force which is caused by the hydrodynamic shear force has

been shown to enhance the biological adhesive bonds or catch bonds by

pulling the ligand-receptor complex apart. Bacterial surface allosteric protein,

FimH, is known to be involved in the catch bond mechanism of adhesion by

mediating the shear-enhanced adhesion where the binding is increased

under high shear stress condition. In addition, bacterial cells are also proven

to adhere strongly to the surface in high shear environments compared to low

shear environments (Sokurenko, et al. , 2008; Thomas, 2008). A previous

publication by Liu, et al. (2002) has also shown that higher hydrodynamic

shear force can lead to stronger biofilms, and biofilms tend to develop a

relatively fragile structure when the shear force is too weak.

Shear stress in shaking conditions has also been suggested to provide higher

chances for adhesion by inducing some conformational changes in the O8-

antigen of E. coli structure (Kumar, et al. , 2015). However, the role of

different O-serotype of K. pneumoniae in biofilm formation upon cultivation

under shaking condition is still unknown and it needs further investigations.

A highly significant increase in biofilm formation was seen upon cultivation of

the hypermucoviscous Kpneu_110 (see Figure 1-2, Chapter 1) strain under

shaking conditions in comparison to static conditions in both of the media with

and without 2 % glucose. Our observation was also in line with the study by

Kong, et al. (2012) who showed that the biofilm formation by

99 |

hypermucoviscous K. pneumoniae strain hvKP1 was greater upon incubation

in LB under shaking conditions compared to static conditions.

A suitable medium to be used for biofilm formation was then determined by

cultivating biofilms in TSB and TSB + 2 % glucose. The addition of glucose

is based on a previous study which mentioned that supplementation of

1 % glucose to TSB increased the carbon to nitrogen ratio in growth medium,

which was associated with enhancement of capsule production in

K. pneumoniae (Domenico, et al. , 1999; Thornton, et al. , 2012). Addition of

glucose in LB broth was also shown to restore capsule production in

a K. pneumoniae trehalose-6-phosphate hydrolase (treC) mutant strain and

subsequently increased biofilm formation of this strain (Wu, et al. , 2011).

Capsule production is suggested to be involved in biofilm formation of

K. pneumoniae, and it is also important for construction of mature biofilm

architecture (Balestrino, et al., 2008). Decreased production of capsule is

thought to be responsible for decreased biofilm formation in K. pneumoniae

(Boddicker, et al., 2006). Supplementation of 2 % glucose in growth medium

such as modified TSB, TSB, and BHI was also shown to enhance biofilm

formation by other bacterial species, for example, E. faecalis. Addition of

glucose was shown to increase the multi-layered biofilm structural and

cellular viability of E. faecalis (Seneviratne, et al. , 2013). Our result,

however, showed no significant difference in biofilm formation by

K. pneumoniae in microtiter plates upon cultivation in TSB or TSB with 2 %

of glucose (Figure 3-6). For that reason, all of the subsequent biofilm assays

were conducted using TSB without any glucose.

100 |

Sub-inhibitory concentrations of antibiotics are known to induce stress in

several bacterial species (Kaplan, 2011). The aim of this research was to

determine whether sub-inhibitory concentrations of common antibiotics used

to treat K. pneumoniae infections may enhance biofilm formation. We tested

the hypothesis by exposing the K. pneumoniae strains to two different

antibiotics which have different modes of action since very few studies have

been carried out to investigate the effect of antibiotic exposure on biofilm

formation by K. pneumoniae. A study by Singla, et al. (2013) which

specifically tested the susceptibility of different phases of biofilm formation,

such as planktonic cells at mid-log phase, planktonic cells at stationary

phase, adherent monolayers and mature biofilms, to three different antibiotics

from different families (piperacillin, amikacin and ciprofloxacin) were different

from our study since they investigated the susceptibility of K. pneumoniae

upon exposure to far higher concentrations of antibiotics, which ranged from

0.062 µg/ml to 1024 µg/ml.

A publication by Romero, et al. (2011), suggests that antibiotics are signalling

molecules. They have been shown to be involved in quorum-sensing and

also interfere with global transcriptional responses at sub-inhibitory

concentrations, for example, in AI-2/LuxS autoinducer signalling. A study by

Ahmed, et al. (2009), showed that 0.04 µg/ml ampicillin, 0.1 µg/ml

ciprofloxacin, and 0.1 µg/ml tetracycline could induce biofilm formation by

S. intermedius wild type strain but not the biofilm formation by a luxS mutant

strain. These antibiotics were shown to increase luxS expression which

encodes for an autoinducer synthase (LuxS). LuxS is known to be involved

101 |

in this AI-2/LuxS signalling system and associated with several bacterial

global regulatory and genetic networks.

The relevance of pre-exposure of cultures to antibiotics prior to biofilm

formation was to investigate if there were any short-term exposure effects of

antibiotics on gene expression that could affect biofilm formation by

K. pneumoniae. Our findings, suggest that there is no significant increase in

biofilm formation by K. pneumoniae after pre-exposure to sub-MICs of

gentamicin and ciprofloxacin compared to controls. We also investigated the

effect of antibiotic exposure on biofilm formation during 24 h incubation. Data

from our study showed that exposure of cultures during 24 h biofilm formation

to some commonly used antibiotics for treatment of infection and

antimicrobial prophylaxis of K. pneumoniae also did not induce biofilm

formation. In addition, we also conducted experiments to examine the effect

of antibiotic exposure on mature biofilms of K. pneumoniae and no increase

in biofilm mass was observed.

Even though, several studies have demonstrated that exposure of bacterial

to sub-MICs of antibiotics could induce biofilm formation of other bacterial

species such as P. aeruginosa (Aka, et al. , 2015), and S. enterica ser.

Infantis (Tezel, et al. , 2016), our findings, however, showed that gentamicin

and ciprofloxacin exposure leads to decreased biofilm formation by several

K. pneumoniae strains, and no increase in biofilm formation was found upon

exposure of different states of growth to different classes of antibiotics.

Owing to the fact that the modes of action of antibiotics and their target are

very specific, it is worth investigating the effect of other antibiotics apart from

102 |

gentamicin and ciprofloxacin on biofilm formation by K. pneumoniae since

very little is known about the mechanisms of antibiotic-induced biofilm

formation in this pathogen.

3.5 Conclusion

In summary, exposure of i) cultures during bacterial growth until mid-log

phase, ii) cultures during biofilm formation, and iii) mature biofilms of

K. pneumoniae cultivated in TSB to sub-MICs of gentamicin and ciprofloxacin

in microtitre plates under shaking conditions does not significantly increase

biofilms of the K. pneumoniae strains used in this study.

103 |

Chapter 4.0

Investigation into extracellular

polymeric substances (EPS) of

Klebsiella pneumoniae

biofilms

104 |

4.0 Investigation into extracellular polymeric

substances (EPS) of K. pneumoniae biofilms

4.1 Introduction

Based on the data in Chapter 3, we concluded that exposure of K. pneumoniae

to two different group of antibiotics did not increase the biofilm formation by

the tested strains. We were then interested in obtaining some insights into

K. pneumoniae extracellular polymeric substances (EPS) compositions since

this pathogen is well-known as one of the species responsible for antibiotic

resistant and biofilm-associated infections. Despite the numerous findings

regarding biofilm-mediated resistance (Bandeira, et al. , 2017; Singla, et al.,

2013; Vuotto, et al., 2014; Vuotto, et al. , 2017) and virulence mechanisms

(Clegg, et al., 2016) of K. pneumoniae, very little is known about the main

constituents of K. pneumoniae biofilm EPS.

4.1.1 The biofilm matrix

In general, the biofilm matrix also known as EPS is composed of proteins,

extracellular DNA (eDNA), exopolysaccharides, and lipids (Flemming, et al.,

2010). Extracellular polysaccharides are regularly identified as the most

prominent components in biofilms (Limoli, et al. , 2015). The exact

composition of EPS differs greatly depending on bacterial species and even

between isolates. Other factors that can affect EPS composition include

growth environments, such as growth media (e.g. nutritious and nutrient-poor

media) and growth conditions such as temperature and external forces (e.g.

shear and hydrodynamic forces).

105 |

4.1.1.1 Proteins

Proteins are suggested to play a significant role in biofilm formation by

providing three-dimensional structural integrity such as amyloid in E. coli and

Pseudomonas spp. (Dueholm, et al. , 2013; Hobley, et al. , 2015; Taglialegna,

et al. , 2016). Proteins in biofilm matrix can be classified as secreted

extracellular proteins and cell surface proteins (Fong, et al. , 2015). Some

secreted proteins act as degrading enzymes or dispersion enzymes that can

enzymatically degrade polysaccharides, proteins, and eDNA (Fong, et al.,

2015). For instance, the glycosyl hydrolase dispersin B (DspB), a soluble N-

acetylglucosaminidase that has been shown to cause detachment of

Actinobacillus actinomycetemcomitans biofilms (Kaplan, et al. , 2003). While,

cell surface proteins can be categorised as non-fimbrial cell surface adhesins

such as Ag43 in E. coli (Berne, et al. , 2015; Hasman, et al. , 1999) and fimbrial

adhesins such as flagella, curli and fimbriae (Latasa, et al. , 2006). These

proteinaceous components are known to be involved in cell-surface

attachment, to interact synergistically with other EPS, such as

exopolysaccharides and eDNA components for stabilisation of the biofilm

matrix, and also to contribute in three-dimensional biofilm architecture (Fong,

et al., 2015).

4.1.1.2 Extracellular DNA (eDNA)

Apart from proteins, extracellular DNA (eDNA) also plays an important role in

the early stage of structural integrity (Das, et al. , 2013) and stability (Song, et

al. , 2016) of biofilms by binding with other matrix components such as

polysaccharides and proteins (Das, et al., 2013). The release of bacterial

106 |

chromosomal DNA into the environment as eDNA is normally caused by cell

autolysis (Montanaro, et al. , 2011). eDNA is also found to be secreted actively

by bacteria, for example, by Neisseria gonorrhoeae (Zweig, et al. , 2014). It is

believed that eDNA could attach to the bacterial surfaces and forms an

extension loop that facilitates the adhesion of the cells to substratum (Song, et

al., 2016). In addition, eDNA also is shown to bind with other biopolymers in

biofilm EPS, for instance, DNABII, the histone like proteins in E. coli

(Goodman, et al. , 2011) and N-acetylglucosamine in Listeria monocytogenes

(Harmsen, et al. , 2010), and these binding mechanisms are important for

bacterial adhesion and improve biofilm structural stability. The role of eDNA

in K. pneumoniae biofilm structure has not been determined, but eDNA is

established to act as a component of biofilms in some other Gram-negative

bacterial species, for example, P. aeruginosa (Whitchurch, et al. , 2002).

4.1.1.3 Polysaccharides

4.1.1.3.1 Cellulose

Other than proteins and eDNA, polysaccharides are also known to play an

important role in biofilm EPS, for example, cellulose and PNAG. Cellulose is

a linear polymer of β-(1-4)-D-glucose produced by a variety of Gram-negative

bacteria as a component of the biofilm matrix, for example, in

S. enterica ser. Typhimurium and E. coli (Zogaj, et al. , 2001). As an example,

cellulose is co-produced with thin aggregative fimbriae by E. coli and

Salmonella spp. to form inert and hydrophobic EPS that mediates cell-cell

interconnections (Zogaj, et al., 2001). Cellulose production in K. pneumoniae

is directed by two bcs operons, cellulose synthase operon I and cellulose

107 |

synthase operon II, and one divergently oriented operon, bEFG operon. The

existence of two bacterial cellulose synthase (bcs) operons in the

K. pneumoniae genome potentially provides this bacterial species with higher

flexibility in term of cellulose production and biofilm formation, and it is

suggested to be due to the lateral gene transfer with extra bcs genes (Römling,

et al. , 2015). The details regarding regulatory mechanisms of biosynthesis of

cellulose in K. pneumoniae are still unknown.

4.1.1.3.2 Poly-N-acetylglucosamine (PNAG)

Poly-N-acetylglucosamine (PNAG: also known as polysaccharide intercellular

adhesin [PIA] and poly-β-1,6-N-acetyl-D-glucosamine [PGA]) is established as

one of many polysaccharides that are important in the biofilm architecture.

This bacterial polysaccharide is known to be involved in the intercellular

cohesion and cell-surface adhesion. PNAG is reported to mediate biofilm

formation by S. epidermidis (Mack, et al. , 1996) and E. coli (Wang, et al. ,

2004), and identified as a major component of B. subtilis biofilm matrix (Roux,

et al. , 2015). Additionally, it is also known to play a role in host bacteria-

interactions, for example, it is crucial for protection of S. epidermidis from the

killing by human innate immunity (Vuong, et al. , 2004) and essential for

bacterial virulence, for instance, in S. aureus (Kropec, et al. , 2005). PNAG is

synthesised by four proteins encoded by a four-gene operon, termed

pgaABCD (polyglucosamine) in E. coli. The pga operon of K. pneumoniae has

a very high similarity with the E. coli pga operon (Chen, et al., 2014) and since

the pgaABCD locus is conserved amongst most K. pneumoniae strains, it is

postulated that K. pneumoniae will have PNAG as a component in its biofilms

(Chen, et al., 2014).

108 |

4.1.2 Extracellular polymeric substances (EPS) produced by

K. pneumoniae during biofilm formation

4.1.2.1 Biofilm EPS of K. pneumoniae upon cultivation in

nutrient-rich media

Little is known about the biofilm EPS of K. pneumoniae. A study by Bandeira,

et al. (2017) found that EPS composition of three K. pneumoniae strains were

richer in exopolysaccharides than eDNA upon cultivation in MH broth in

96-well polystyrene plates. The protein content, however, differed between

each strain where one of them showed a particularly high percentage of protein

content around 65 % in its biofilm in comparison to PNAG and eDNA.

A publication by Flemming, et al. (2010) suggested that cellulose could be an

important biofilm EPS of K. pneumoniae This is supported by Wang, et al.

(2016) who showed that cellulose was a major biofilm component when

K. pneumoniae was grown under a simulated microgravity environment (SMG)

in LB medium compared to controls. A carbapenem-resistant K. pneumoniae

strain ATCC BAA-1705 was grown aerobically in LB broth in two different

conditions. The test group (SMG) was conducted by growing the bacterial

cells in high-aspect ratio vessels (HARV) bioreactors and rotating the

bioreactor with its axis perpendicular to gravity, while, the control (NG) was

conducted by growing the bacteria in HARV reactors with its axis parallel to

gravity.

109 |

4.1.3 Artificial urine medium (AUM)

Since previous studies have shown that cultivation in different rich media has

an effect on the composition of EPS of K. pneumoniae biofilm, we wondered

what the effect of cultivation in a nutrient-poor medium such as human urine

would be on the composition of EPS. Since K. pneumoniae is often associated

with urinary tract infections, the usage of artificial urine medium (AUM) has

clinical relevance since it is chemically more representative of real human

urine than nutrient-rich media like MH or LB. Whilst usage of real human urine

could be considered to be more relevant than using AUM for cultivation, it does

have some pitfalls. The chemical composition of human urine obtained from

the donors is highly dependent on the donors’ diet, fluid intake, and life style.

By eating more carbohydrate-rich foods, monosaccharide levels (e.g. glucose)

will be affected and consumption of more sodium-rich foods will affect the

sodium level in urine. The amount of fluid intake will also affect the

concentrations of nutrients in urine. These variations could make it very

difficult to reproduce experiments. By using an AUM, the composition of the

nutrients can be kept constant, making experiments more reproducible. In the

study described in this chapter, we used an AUM based on that designed by

Brooks, et al. (1997) with some additions to mimic human urine.

110 |

4.1.4. Aims of this chapter

In this chapter, we investigated the capacity of seven K. pneumoniae strains

to form biofilms in nutritious medium, tryptic soy broth (TSB) and nutrient-poor

medium, artificial urine medium (AUM). It is still unclear what EPS are

produced by K. pneumoniae during biofilm formation, since very few studies

have been published regarding this matter. The purpose of the work described

in this chapter was to determine the components of EPS produced by

K. pneumoniae that contributed to biofilms using enzymatic treatment

(proteinase K, DNase I and cellulase) and a mild oxidising agent (sodium

metaperiodate, NaIO4) upon cultivation in nutrient-rich and nutrient-poor

medium.

111 |

4.2 Materials and methods

4.2.1 Bioinformatic analysis

Using a bioinformatic approach, the operons involved in the production of

polysaccharides such as cellulose and PNAG were identified in the draft whole

genome sequence of the strains used in this study. Comparative analysis of

the DNA sequences was conducted to compare the operons amongst the

strains using BLASTn, BLASTp (Altschul, et al., 1990), EMBOSS-transeq

(Rice, et al., 2000), Clustal Omega (Sievers, et al., 2011), and Genedoc. The

details of the software were described in Chapter 2, Table 2-8.

4.2.2 Media

Media used in this study were tryptic soy broth, TSB (Sigma, T8907) and

artificial urine medium (AUM). TSB was prepared according to the

manufacturer’s protocol (Sigma), while AUM was prepared according to the

method published by Brooks, et al. (1997) with some modifications. The

components of the media are described in the table below (Table 4-1). The

TSB was autoclaved at 121 °C for 15 min. The AUM was prepared by

dissolving all the components in autoclaved ultrapure Milli-Q water (MQ H2O)

(Millipore, UK). The final pH for the AUM was adjusted to pH 6.5 and the

medium was then filter sterilised with a 0.2 µM PES filter (Nalgene, UK) and

kept at 4 °C.

112 |

Table 4-1 Media used in this study

Media

Tryptic Soy Broth (TSB) (Sigma, T8907)

Compositions

Quantity (g/L)

Casein (pancreatic digest)

Soya peptone (papaic digest)

Sodium chloride

Dipotassium phosphate

Dextrose

17

3

5

2.5

2.5

Artificial urine medium (AUM)

Compositions

Quantity (g/L)

Molarity (mmol/l)

Peptone L37 (Oxoid, LP0037)

Yeast extract (Fluka, 70161)

D-Lactic acid (Sigma, L1893)

Citric acid (Sigma, C8532)

Sodium bicarbonate

(Sigma, S6014)

Urea (Sigma, 15604)

Uric acid (Sigma, U0881)

Creatinine (Sigma, C4255)

Calcium chloride.2H2O

(Analar, 1007044)

Sodium chloride (Sigma, S9625)

Iron II sulphate.7H2O

(Sigma, F7002)

Magnesium sulphate .7H2O

(Analar, 1015144)

Sodium sulphate (Analar, 10264)

1

0.005

0.1

0.4

2.1

10

0.07

0.8

0.37

5.2

0.0012

0.49

3.2

1.1

2

25

170

0.4

7

2.5

90

0.005

2

22.5

113 |

Potassium dihydrogen

orthophosphate (BDH, 29608)

Di-potassium hydrogen

phosphate (Merck,

AM04300204448/1.05104.1)

Ammonium chloride

(Sigma, A5666)

0.95

1.2

1.3

7

7

25

Modified supplement

Quantity (g/L)

Molarity (mmol/l)

D-Glucose anhydrous

(Analar, 10117744)

Sucrose (Sigma, 16104)

D- Lactose monohydrate

(Sigma, L1768)

MEM amino acid solution

(50X) - (Sigma, M550)

RPMI-1640 vitamin solution

(100X) - (Sigma, R7256)

L- Arginine (Sigma, A5006)

Glycine (Sigma, G8790)

L-Histidine (Sigma, H8000)

Nicotinic acid (Sigma, N4126)

Sterile MQ H2O

0.02

0.02

0.02

20 ml/L

10 ml/L

1

1

1

1

To 1 liter

0.1

0.06

0.06

5.8

13.3

6.4

8.1

Abbreviations:

* MEM – Minimum essential media

* RPMI – Roswell Park Memorial Institute

114 |

4.2.3 Growth kinetics of examined K. pneumoniae strains in

TSB and AUM

A single colony of each K. pneumoniae strain was separately inoculated into

5 ml of TSB or AUM and the culture was incubated at 37 °C for 16 h on a rotary

shaker at 200 rpm. The 16 h culture was subsequently diluted to an optical

density at 600 nm (OD600) of 0.05 and inoculated into a 50 ml tube containing

10 ml of fresh TSB or AUM. The culture was further incubated at 37 °C on the

rotary shaker and the absorbance (OD600) was determined using a WPA

CO8000 cell density meter (Biochrom Ltd, UK) at hourly time intervals.

4.2.3.1 Determination of relationship between absorbance

(OD600) and colony forming unit per ml (CFU/ml)

Three different strains of K. pneumoniae were chosen to determine the

correlation between absorbance (OD600) and colony forming unit per ml

(CFU/ml) upon growth in suitable media. They were Kpneu_1, a clinical strain,

Kpneu_110, a hypermucoviscous strain, and C3091, a laboratory strain. The

correlation was determined by measuring the absorbance (OD600) at hourly

time intervals, and subsequently making a series of serial dilutions of the

corresponding suspensions.

The 16 h culture was diluted to OD600 of 0.05 and inoculated into a 50 ml tube

containing 10 ml of fresh TSB. The culture was incubated at 37 °C on the

rotary shaker and the absorbance (OD600) was determined. The suspension

was diluted accordingly once the OD600 reached the range of OD600 ~1.0

and the exact OD600 was determined by multiplying the absorbance reading

obtained with the dilution factor. The series of diluted suspensions were

115 |

immediately plated on TSA for viable count at hourly time intervals to

determine the colony forming unit per ml (CFU/ml). The plates were incubated

at 37 °C for 16 h and the enumeration was then conducted. The graph of the

relationship between absorbance and the CFU/ml was plotted.

4.2.3.2 Determination of bacterial growth rate

The growth rate constant (µ) of the strains was determined by calculation from

the following equation (Hall, et al. , 2014).

ln OD2 – ln OD1 = µ (t- t0)

OD1 – absorbance (OD600 nm) or CFU1/ml at t0

OD2 – absorbance (OD600 nm) or CFU2/ml at t

t –time at the late-log phase

t0 –time at the early-log phase

The above equation was converted as follows:

log10 (OD2 or CFU2/ml) – log10 (OD1 or CFU1/ml) = (µ/2.303) (t- t0)

or alternatively is

µ= (log10 OD2 or CFU2/ml – log10 OD1 or CFU1/ml) x 2.303/ (t- t0)

and the doubling time (td) can be determined by

µ= ln 2/ td

116 |

Example of the calculation for:

Growth rate (µ) and doubling time (td) calculation of Kpneu_1 upon cultivation

in TSB (using OD600 as an example).

Equation 1 (to calculate the growth rate constant)

µ = (log10 OD2 – log10 OD1) x 2.303/ (t- t0)

= (log10 3.3 – log10 0.3) x 2.303/ (3 – 1)

= 0.5186 – (-0.523) x 2.303/ 2

= 2.399/2

= 1.199 h-1

Equation 2 (to calculate the doubling time)

td = ln 2/ µ

= 0.693/ 1.199

= 0.578

= 0.578 x 60 min

= 34.67

= 34.7 min

117 |

4.2.4 Biofilm assays in TSB and AUM

The biofilms were grown in TSB and AUM in flat-bottom 96-well polystyrene

tissue culture treated plates (Sarstedt, Germany) and incubated at 37 °C for

24 h according to the previously mentioned method in Chapter 3, section 3.2.1.

4.2.5 Biofilm inhibition and disruption assays

4.2.5.1 Biofilm inhibition assays

To determine the composition of EPS within biofilms of K. pneumoniae in TSB

and AUM, biofilm inhibition and disruption assays in the presence of

i) proteinase K ([EC 3.4.21.64] [≥30 units/mg] [Sigma, P2308]),

ii) deoxyribonuclease I from bovine pancreas ([DNase I] [EC 3.1.21.1] [≥400

KUnitz units/mg] [Sigma, DN25]), or iii) cellulase ([EC 3.2.1.4] [≥700 units/g]

[Sigma, 2730]) were conducted. For the inhibition assays, the biofilm-forming

assays was carried out in the presence of 100 µg/ml proteinase K, 100 µg/ml

DNase, or 0.1 % cellulase in the medium, and the strains were allowed to form

biofilms at 37 °C for 24 h under shaking condition at 200 rpm on an orbital

shaker (Luckham, UK). No inhibition assays in the presence of NaIO4 was

conducted since NaIO4 is established to inhibit bacterial growth when added

into the growth medium. The biofilms grown in media without any

proteinase K, DNase, and cellulose acted as controls. The media without any

inoculation were used as a sterility control. The biofilm mass was quantified

using the CV staining assays detailed in the Chapter 3 section 3.2.5.

118 |

4.2.5.2 Biofilm disruption assays

Biofilms were grown at 37 °C for 24 h and the planktonic culture from the wells

containing 24 h old biofilms were removed, and the wells were washed

carefully with 200 µl PBS. The pre-formed biofilms were treated with

proteinase K, DNase, NaIO4, or cellulase. Three hundred microlitres of

i) 100 µg/ml proteinase K in 20 mM Tris-HCl (pH 7.5) and 100 mM NaCl,

ii) 100 µg/ml DNase in 150 mM NaCl, iii) NaIO4 (Sigma, S1878) in MQ H2O, or

iv) 0.1 % cellulase (v/v) in 0.05 M citrate buffer (pH 4.8) were added into the

wells. The treated biofilms were further incubated at 37 °C for 3 h for

proteinase K, at 37 °C for 24 h for DNase, at 4 °C for 24 h for NaIO4, and at

45 °C for 24 h for cellulase. The biofilms treated with corresponding buffers

only, acted as controls. The biofilms were then quantified using the CV staining

assay.

119 |

4.3 Results

4.3.1 Bioinformatic analysis

Bioinformatic analysis was conducted to determine if operons identified in

other bacteria as being involved in biofilm formation may be present in

K. pneumoniae before conducting any further experiments. Cellulose and

PNAG operons (Figure 4-1 and Figure 4-6) were chosen to be investigated

since these polysaccharides were speculated to contribute to biofilm

composition of K. pneumoniae (Chen, et al., 2014; Huertas, et al., 2014). The

bcs and pga operons of the strains were identified in the draft whole genome

sequences data. Comparative analysis was conducted between all the strains

for both bcs and pga operons, and also between Kpneu_1 and E. coli MG1655

for pga operon only. These analyses were conducted to identify any

differences in the genes in these operons. We also wanted to compare the

amino acid sequences and identify any amino acid variations between the

strains.

4.3.1.1 Comparative analysis of bacterial cellulose

synthase (bcs) operon

We performed the comparative analysis to confirm the existence of a complete

bcs operon and to identify if there were any defects in the genes that could

result in a truncated protein in the genome of all the strains used in this study.

Comparative analysis was conducted by comparing the genes in bcs operons

between all of the strains (Figure 4-1). Our results revealed differences in

several genes including bcsO (deletion of 1 bp) in CG43S3 which would result

in a truncated BcsO (Figure 4-2), bcsQ_I (deletion of 1 bp) in Kpneu_63 and

120 |

Kpneu_64 which would result in a truncated BcsQ (Figure 4-3), bcsB_I

(insertion of 5 bp) and bcsC_I (deletion of 52 bp) in Kpneu_1 which would

result in a truncated BcsB_I and BcsC_I proteins (Figure 4-4 and Figure 4-5)

compared to the rest of the strains. Several single nucleotide polymorphisms

(SNPs) in other genes were also identified among the strains. The SNPs

resulted in amino acid variations were shown in Table 4-2.

121 |

Colour coding labels:

122 |

Figure 4-1 Schematic presentation of bcs operons of K. pneumoniae.

The bcs cluster of K. pneumoniae can be divided into three operons, bcsI, bcsII, and the divergent bEFG operon. The first operon, bcsI consists

of seven genes encoding for: BcsO, BscQ; four core cellulose synthesis subunits: BcsA, BcsB, BcsC, and BcsD, and also the regulator for

endoglucanase, CelY. While operon II is comprised of six genes encoding for: BscR, BcsQ, BcsA, BcsB, BcsZ, and BcsC. Whereas the divergent

operon consists of three genes, bcsE, bcsF, and bcsG. A) the bcs cluster of Kpneu_65, Kpneu_110, and C3091 contains three complete bcsI,

bcsII and bEFG operons, B) the bcs cluster of CG43S3 contains a defective bcsO, C) the bcs cluster of Kpneu_63 and Kpneu_64 contain defective

bcsQ genes, and D) the bcs cluster of Kpneu_1 contains defective bcsB_I and bcsC_I genes. The name of the genes is indicated by the small

colour-coding circles. The figure is drawn and adapted from Huertas, et al. (2014) and Jahn, et al. (2011).

123 |

Figure 4-2 Partial multiple sequence alignment of bcsO and BcsO.

The figure shows A) A partial sequence alignment of bcsO. Multiple sequence alignment of genes in the bcs operons revealed a single nucleotide

deletion at location 328 in the bcsO of CG43S3 (boxed in black) that resulted in a frameshift mutation (the starting of the reading frame shift is

marked with a black line under the AGC sequence which encodes for the amino acid serine, S). B) A partial sequence alignment of BcsO. This

reading frame shift leads to the incorrect protein product (starts from S which was shown by hollow arrow) and a premature stop codon that leads

to the truncation of the predicted protein (showed by filled arrow). The length of the truncated BcsO protein in CG43S3 is 153 aa and the length

of the intact BcsO in the other strains is 179 aa. The amino acid variation for the other strains which resulted from the SNPs is presented in Table

4-2.

124 |

Figure 4-3 Partial multiple sequence alignment of bcsQ and BcsQ.

The figure shows A) A partial sequence alignment of bcsQ. Multiple sequence alignment of genes in the bcs operons revealed a single nucleotide

deletion at location 61 in the bcsQ of Kpneu_63 and Kpneu_64 (boxed in black) that resulted in a frameshift mutation (the starting of the reading

frame shift is marked with a black line under the ATT which encodes for amino acid isoleucine, I. The adjacent TAG (marked with the black line)

encodes for the stop codon. B) A partial sequence alignment of BcsQ. This reading frame shift resulted in a premature stop codon (showed by

filled arrow) and a truncation of the predicted protein. The length of the truncated BcsQ protein in Kpneu_63 and Kpneu_64 is 21 aa and the

length of the intact BcsQ in the other strains is 267 aa. The amino acid variation for the other strains which resulted from the SNPs is presented

in Table 4-2.

125 |

Figure 4-4 Partial multiple sequence alignment of bcsB_I and BcsB_I.

The figure shows A) A partial sequence alignment of bcsB_I. Multiple sequence alignment of genes in the bcs operons revealed a 5 bp insertion

starting at location 189 in bcsB_I of Kpneu_1 (showed by black line and black arrow) that resulted in a frameshift mutation (the starting of the

reading frame shift is marked with a black line on top of CGC and CCC that encodes for the amino acid proline, P). B) A partial sequence

alignment of BcsB_I. This reading frame shift leads to the incorrect protein product (starts from P which was shown by hollow arrow) and a

premature stop codon that leads to the truncation of the predicted protein (showed by filled arrow). The length of the trucated BcsB_I protein in

Kpneu_1 is 86 aa and the length of the intact BcsB_I in the other strains is 811 aa. The amino acid variation for the other strains which resulted

from the SNPs is presented in Table 4-2.

126 |

Figure 4-5 Partial multiple sequence alignment of bcsC_I and BcsC_I.

The figure shows A) A partial sequence alignment of bcsC_I. Multiple sequence alignment of genes in the bcs operons revealed a 52 bp deletion

starting at location 1125 in bcsC_I of Kpneu_1 (showed by black line and black arrow) that resulted in a frameshift mutation (the starting of the

reading frame shift is marked with a black line on top of ATG that encodes for the amino acid methionine, M). B) A partial sequence alignment of

Bcs_I. This reading frame shift leads to the incorrect protein product (starts from M which was shown by hollow arrow) and a premature stop

codon that leads to the truncation of the predicted protein (showed by filled arrow). The length of the truncated BcsC_I protein in Kpneu_1 is 394

aa and the length of the intact BcsC_I in the other strains is 1350 aa. The amino acid variation for the other strains which resulted from the SNPs

is presented in Table 4-2.

127 |

Table 4-2 Amino acid variation in proteins encoded by the bcs operon.

128 |

4.3.1.2 Comparative analysis of pga operon

To verify the existence of a complete pga operon in all of the genomes of the

examined strains, comparative analyses of amino acid sequences were

conducted by comparing the deduced amino acids of PgaA, PgaB, PgaC and

PgaD among the strains. The amino acid variation between strains are shown

in Table 4-3. In addition, a pairwise amino acid sequence alignment of PgaA,

PgaB, PgaC, and PgaD of E. coli MG1655 and K. pneumoniae strain Kpneu_1

was also conducted to identify the amino acid sequence identity between

E. coli and K. pneumoniae. (Figure 4-6). Comparison between deduced amino

acid sequences of PgaA, PgaB, PgaC, and PgaD of K. pneumoniae showed

high level of amino acid identity percentage with Pga proteins of E. coli

MG1655.

Figure 4-6 shows the schematic diagram of the pga operon. Both E. coli and

K. pneumoniae pga operons consist of four genes encoding for proteins

associated with the synthesis of PNAG. PgaA of Kpneu_1 K. pneumoniae

(829 aa) had 41 % amino acid identity with PgaA of E. coli (807 aa) (Figure

4.7A); PgaB of Kpneu_1 (671 aa) had 53 % amino acid identity with PgaB of

E. coli (672 aa) (Figure 4.7B). While PgaC of Kpneu_1 (442 aa) had about

65 % amino acid identity with PgaC of E. coli (441 aa) (Figure 4.7C), and

PgaD of Kpneu_1 (149 aa) had 25 % amino acid identity with PgaD of E. coli

(137 aa) (Figure 4.7D). The pairwise amino acid sequence alignment of Pga

of E. coli MG1655 and Kpneu_1 is presented in Figure 4-7.

129 |

Table 4-3 Amino acid variation in proteins encoded by pga operon.

Figure 4-6 Schematic presentation of pgaABCD operon.

Schematic diagram of the pgaABCD gene cluster of Kpneu_1 and E. coli MG1655

are shown for comparison. The percentage amino acid sequence identity of E. coli

with K. pneumoniae homologs was determined with Clustal Omega. The figure is

drawn and adapted from Chen, et al. (2014). pgaA is indicated as a yellow arrow,

pgaB as a green arrow, pgaC as a blue arrow, and pgaD as a pink arrow.

130 |

Figure 4.7A) Pairwise sequence alignment between PgaA of E. coli MG1655 and

PgaA of Kpneu_1. PgaA of E. coli consists of 807 aa while PgaA of Kpneu_1 consists

of 829 aa.

131 |

Figure 4.7B) Pairwise sequence alignment between PgaB of E. coli MG1655 and

PgaB of Kpneu_1. PgaB of E. coli consists of 672 aa and PgaB of Kpneu_1 consists

of 671 aa.

132 |

Figure 4.7C) Pairwise sequence alignment between PgaC of E. coli MG1655 and

PgaC of Kpneu_1. PgaC of E. coli consists of 441 aa and PgaC of Kpneu_1 consists

of 442 aa.

133 |

Figure 4.7D) Pairwise sequence alignment of PgaD of E. coli MG1655 with PgaD of

Kpneu_1. PgaD of E. coli consists of 137 aa and PgaD of Kpneu_1 consists of 149

aa.

Indicators:

Figure 4-7 Pairwise sequence alignment of Pga amino acid.

The figures show amino acid pairwise sequence alignments of A) PgaA, B) PgaB,

C) PgaC, and D) PgaD of E. coli MG1655 and Kpneu_1. The alignments were

conducted using Clustal Omega (Sievers, et al., 2011).

134 |

4.3.2 Determination of the relationship between absorbance

and colony forming unit per ml (CFU/ml)

The relationship between absorbance and CFU/ml was determined by

measuring the absorbance at hourly time intervals and immediately plating the

diluted suspensions on the TSA. The graphs in Figure 4.8 show the growth

curve and the CFU/ml of the different isolates, Kpneu_1, Kpneu_110, and

C3091 upon growth in TSB. From the calculation based on the equation in

Chapter 4 section 4.2.3.2, it was found that, the growth rate constant and

doubling times obtained from the calculation of the two parameters, the

absorbance and CFU/ml of these three strains were considered no difference

since no significant difference was found upon statistical analysis. All the

strains had approximately 30 min doubling time (Table 4-4). The growth rate

(µ) is the number of generations per hour, while the doubling time (td) is the

average time required for a particular bacterial population to double. All of the

graphs obtained in these experiments also showed that OD600 ~1.0 is equal

to ~109 CFU/ml after 2 h incubation in TSB for all three strains. Since we could

relate the absorbance value directly to the number of CFU/ml, absorbance

readings were used for the determination of the growth rates and doubling

times of the other strains in the subsequent experiments.

135 |

136 |

Figure 4-8 The relationship between absorbance and CFU/ml of K.pneumoniae

upon cultivation in TSB.

The figure shows the growth curve and CFU/ml of A) Kpneu_1, B) Kpneu_110, and

C) C3091 upon cultivation in TSB. Data presented are the mean of three replicates

and error bars indicate the standard deviations. All of the strains showed almost

similar growth upon cultivation in TSB. The OD1/t0 or CFU1/ml/t0 and OD2/t or

CFU2/ml/t used for calculation of the growth rate constant and doubling time are as

labelled. The graph lines indicate as follow: blue: Kpneu_1, orange: Kpneu_110, red:

C3091 and fuchsia: CFU/ml for all the strains. The OD1 and OD2 are the absorbance

values, CFU1/ml and CFU2/ml are the colony forming unit per ml and t0 and t are the

time, at early and late log phase used for calculation based on the equation in Chapter

4 section 4.2.3.2.

137 |

Table 4-4 Growth rate constant (µ) and doubling time (td) of K. pneumoniae strains upon cultivation in TSB using OD600 and CFU/ml

as parameters

The data is based on the growth curve in Figure 4.8. No significant differences were found between the growth rates constant and doubling times

obtained from the calculation using OD600 and CFU/ml upon statistical analysis using Mann-Whitney non-parametric test.

Strain

Parameters

Equation 1

(h-1)

Equation 2

(min)

µ= (log10 OD2 – log10 OD1) 2.303/ (t- t0) or

µ= (log10 CFU2/ml – log10 CFU1/ml) 2.303/ (t- t0)

td = ln 2/ µ

Kpneu_1

OD600 1.41 29.5

CFU/ml 1.63 25.4

Kpneu_110

OD600 1.25 33.3

CFU/ml 1.58 26.3

C3091

OD600 1.4 29.8

CFU/ml 1.85 22.4

138 |

4.3.3 Growth kinetics of K. pneumoniae strains in TSB and

AUM

From the data obtained in the previous sub-chapter, the growth kinetics of

K. pneumoniae was then determined in TSB and AUM. Growth curves were

performed in TSB and AUM for all examined strains and showed that all the

strains grew faster in TSB compared to AUM. This can be seen by the duration

for the strains to reach mid-log phase and mean generation time or doubling

time. The growth rate (µ) is the number of generations per hour while the

doubling time (td) is the average time required for those particular bacterial

population to double. The growth constant rate and the doubling time were

determined by the calculations (Equation 1 and Equation 2) explained in

section 4.3.2. All strains reached mid-log phase after 2 h upon cultivation in

TSB (Figure 4-9Ai) while they approached the mid-log phase approximately

after 4 h upon cultivation in AUM (Figure 4-9Bi).

139 |

Figure 4.9A) Growth curves of K. pneumoniae in TSB in two different scales: i) linear

scale and ii) logarithmic scale. The linear scale graph was plotted to determine the

early log phase and late log phase for calculation of the growth rate constant and

doubling time, while the logarithmic scale was plotted to visualise the exponential

phase clearly (that might not be very obvious on the linear scale graph). Mid-log

phase (indicated by the red arrow) of all the strains was observed after 2 h upon

cultivation in TSB.

140 |

Figure 4.9B) Growth curves of K. pneumoniae in AUM in two different scales: i) linear

scale and ii) logarithmic scale. The linear scale graph was plotted to determine the

early log phase and late log phase for calculation of the growth rate constant and

doubling time, while the logarithmic scale was plotted to visualise the exponential

phase clearly (that might not be very obvious on the linear scale graph). A different

pattern of growth curve can be seen upon cultivation of tested K. pneumoniae strains

in AUM. Mid-log phase (indicated by the red arrow) of all the strains was observed

after 4 h upon cultivation in AUM.

141 |

Figure 4-9 Growth of K. pneumoniae in two different growth media.

The figures show growth kinetics of K. pneumoniae upon cultivation in A) TSB and B)

AUM. Data presented are the mean of two independent experiments and error bars

indicate the standard deviation. The graph lines are indicated as follow: blue:

Kpneu_1, light green: Kpneu_63, black: Kpneu_64, magenta: Kpneu_65, orange:

Kpneu_110, red: C3091 and light blue: CG43S3.

An example of the calculation is shown in section 4.2.3.1, and the growth rate

constant (µ) and doubling time (td) of all the strains tested in this study are

presented in Table 4-5. Based on the calculations, C3091 had longer doubling

time (41 min) compared to the rest of the strains, which was about 34-39 min

upon cultivation in TSB (Figure 4-9A). Whilst, strain Kpneu_1, and C3091 had

the longest doubling time, which was about 114 min (113 -115 min) upon

cultivation in AUM (Figure 4-9B). This was followed by strain Kpneu_64 and

Kpneu_110 which was about 101 min (100 – 102 min) and strain Kpneu_63

which was 92 min. Strain CG43S3 had a shortest doubling time upon growth

in AUM which was about 71 min.

142 |

Table 4-5 Growth rate constant (µ) and doubling time (td) of K. pneumoniae strains upon cultivation in TSB and AUM.

Strain

Media

Equation 1

(h-1)

Equation 2

(min)

µ= (log10 OD2 – log10 OD1) 2.303/ (t- t0) td = ln 2/ µ

Kpneu_1 TSB 1.19 35

AUM 0.36 115

Kpneu_63 TSB 1.22 34

AUM 0.45 92

Kpneu_64 TSB 1.07 39

AUM 0.41 102

Kpneu_65 TSB 1.17 36

AUM 0.49 85

Kpneu_110 TSB 1.07 38.5

AUM 0.42 100

C3091 TSB 1.00 41

AUM 0.37 113

G43S3 TSB 1.11 38

AUM 0.58 71

143 |

4.3.4 Quantification of K. pneumoniae biofilm upon cultivation

in TSB and AUM

An early objective of this study was to devise the assays to quantify the biofilms

mass (bacterial cell and EPS) of K. pneumoniae upon cultivation in two

different media, highly nutritious medium, TSB, and a nutrient-poor medium,

AUM. Figure 4-10 shows that significant increases were observed for biofilm

formation by Kpneu_1, Kpneu_64, and Kpneu_110 cultivated in TSB

compared to biofilm formation in AUM. Kpneu_110 was observed to have a

very low capacity to form biofilm in AUM. While strain Kpneu_65, C3091, and

CG43S3 were observed to form more biofilm in AUM compared to TSB.

144 |

Figure 4-10 Biofilm formation by K. pneumoniae in TSB and AUM.

The figure shows biofilm formation by K. pneumoniae in TSB and AUM upon

incubation at 37 °C for 24 h under shaking conditions. Bars show the average

absorbance obtained using the CV staining assays from three independent

experiments. The data was analysed by a Mann-Whitney non-parametric test. Error

bars indicate standard deviation of n = 15. Significant increases were observed for

biofilm formation by strain Kpneu_1, Kpneu_64, and Kpneu_110, cultivated in TSB

compared to AUM, while strain Kpneu_65, C3091, and CG43S3 were observed to

form more biofilms in AUM compared to TSB. There was no significant difference

found for strain Kpneu_63 upon cultivation in both media. Statistical significance is

indicated by * represents p < 0.05, ** represents p < 0.01, *** represents p < 0.001

and **** represents p < 0.0001.

145 |

4.3.5 Biofilm inhibition and disruption assays

In order to study the components of K. pneumoniae biofilm EPS upon

cultivation in two different media, biofilm inhibition and biofilm disruption

assays were conducted. Proteinase K, DNase, and cellulase were used in

both of the assays, while NaIO4 was only used in the disruption assays.

Proteinase K catalyses the hydrolysis of peptide bonds in peptide or protein

substrates (Saenger, 2013). Cellulase catalyses the breakdown of cellulose

into glucose by endohydrolysis of the β-(1-4)-D glucosidic linkages in cellulose

(Liao, et al. , 2011). While, DNase I catalyses the endonucleolytic cleavage of

substrate to yield 5’-phosphodinucleotide and 5’-phosphooligonucleotide end

products (Kishi, et al. , 2001).

The biofilm inhibition assays were conducted by incubating the culture in

control medium (without any enzymes) or in medium with these particular

enzymes in 96-well microtitre plates. The culture was then incubated at 37 °C

for 24 h. The biofilm disruption assays were conducted by treating the 24 h

pre-formed biofilms with proteinase K, DNase, cellulase or NaIO4. The treated

biofilms were then incubated at 37 °C for 3 h or 1 h for proteinase K and DNase,

respectively, at 48 °C for 24 h for cellulase, and at 4 °C for 24 h for NaIO4.

Pre-formed biofilms incubated in buffer only (without any enzymes or NaIO4)

acted as controls.

146 |

4.3.5.1 Effect of 100 µg/ml proteinase K on biofilm

formation and on 24 h pre-formed biofilms

The biofilm inhibition assays (Figure 4-11A and Figure 4-11B) in the presence

of 100 µg/ml of proteinase K showed that biofilm formation by Kpneu_1

cultivated in TSB and biofilm formation by Kpneu_63, Kpneu_65, C3091, and

CG43S3 cultivated in AUM was significantly lower in the presence of

proteinase K (100 µg/ml) compared to controls. Whilst, biofilm formation by

Kpneu_110 was significantly increased upon cultivation in TSB in the presence

of proteinase K (100 µg/ml). The biofilm disruption assays (Figure 4-11C and

Figure 4-11D) showed that apart from Kpneu_110 and CG43S3, pre-formed

biofilms of Kpneu_1, Kpneu_63, Kpneu_64, Kpneu_65, and C3091 in TSB

were significantly disturbed upon treatment with proteinase K (100 µg/ml).

While upon cultivation in AUM, pre-formed biofilms of Kpneu_65 and

Kpneu_110 were observed to be disrupted by proteinase K.

147 |

148 |

Figure 4-11 Biofilm inhibition and disruption assays in the presence of 100 µg/ml proteinase K.

The graph shows biofilm inhibition assays of K. pneumoniae cultivated in A) TSB and B) AUM in the presence of proteinase K (100 g/ml)

incubated at 37 °C for 24 h, and disruption of pre-formed biofilms cultivated in C) TSB and D) AUM upon treatment with proteinase K (100 g/ml)

at 37 °C for 3 h. Bars show average absorbance readings using the CV assays from three independent experiments. The data was analysed by

a Mann-Whitney non-parametric test. Error bars indicate standard deviation of nine replicates. Kpneu_1 was observed to have significantly lower

biofilm formation and Kpneu_110 showed significantly higher biofilm formation upon cultivation in TSB in the presence of proteinase K (100 g/ml)

compared to control. While Kpneu_63, Kpneu_65, C3091 and CG43S3 had significantly lower biofilms upon cultivation in AUM in the presence

of proteinase K (100 g/ml) in comparison to controls. The biofilm disruption assays demonstrated that pre-formed biofilms of Kpneu_1,

Kpneu_63, Kpneu_64, Kpneu_65, and C3091 in TSB were significantly reduced as well as pre-formed biofilm of Kpneu_65 and Kpneu_110 in

AUM upon treatment with proteinase K (100 g/ml) compared to controls. Statistical significance is indicated by * represents p < 0.05,

** represents p < 0.01, *** represents p < 0.001 and **** represents p < 0.0001.

149 |

4.3.5.2 Effect of 100 µg/ml of DNase on biofilm formation

and on 24 h pre-formed biofilms

Since it is established that eDNA has an important role in stabilising the biofilm

structure in other bacterial species, e.g. P. aeruginosa (Whitchurch, et al.,

2002), we also investigated its role in biofilm formation and on pre-formed

biofilms of K. pneumoniae. Biofilm formation by Kpneu_63 and C3091 was

significantly increased upon cultivation in TSB in the presence of

100 µg/ml of DNase (Figure 4-12A), while biofilm formation by Kpneu_65 was

significantly reduced upon cultivation in AUM in the presence of 100 µg/ml of

DNase compared to controls (Figure 4-12B). The treatment of pre-formed

Kpneu_110 biofilms with 100 µg/ml of DNase resulted in a significantly higher

CV absorbance readings compared to controls, whereas, treatment of pre-

formed C3091 biofilms with 100 µg/ml of DNase exhibited a significant

reduction of the pre-formed biofilm in comparison to the control group (Figure

4-12C and Figure 4-12D). Our result suggested that eDNA possibly play a

significant role in biofilm formation by Kpneu_65 upon cultivation in AUM and

it is also suggested that eDNA possibly acts as a major composition in pre-

formed biofilms of C3091 upon cultivation in AUM.

150 |

151 |

Figure 4-12 Biofilm inhibition and disruption assays in the presence of 100 µg/ml DNase.

The graph shows biofilm inhibition assays of K. pneumoniae cultivated in A) TSB and B) AUM in the presence of DNase (100 µg/ml) incubated at

37 °C for 24 h, and disruption of pre-formed biofilms cultivated in C) TSB and D) AUM upon treatment with DNase (100 µg/ml) at 37 °C for 24 h.

Bars show average absorbance readings using the CV assays from three independent experiments. The data was analysed by a Mann-Whitney

non-parametric test. Error bars indicate standard deviation of nine replicates. A significant difference can be seen for biofilm formation of

Kpneu_63 and C3091 upon cultivation in TSB in the presence of 100 µg/ml of DNase and additionally, biofilms of Kpneu_65 was shown to have

a significant reduction upon cultivation in AUM in the presence of 100 µg/ml of DNase compared to controls. The biofilm disruption assays

demonstrated that CV absorbance readings of pre-formed biofilms of Kpneu_110 in TSB was significantly higher and pre-formed biofilms of C3091

in AUM was significantly disrupted upon treatment with 100 µg/ml of DNase compared to controls. Statistical significance is indicated by

* represents p < 0.05 and ** represents p < 0.01.

152 |

4.3.5.3 Effect of NaIO4 on 24 h pre-formed biofilms

We then assayed the sensitivity of pre-formed biofilms produced by the

K. pneumoniae strains to NaIO4, which is known to cause disruption of

polysaccharides in biofilm by oxidising the carbon to carbon bond at vicinal

diols in carbohydrate rings (Meile, et al. , 2014). The results showed that the

CV absorbance of the pre-formed biofilms of all seven strains cultivated in TSB

were significantly higher upon treatment with NaIO4 (Figure 4-13A) compared

to controls. Whereas, only five of the seven strains, CV absorbance of the pre-

formed biofilm was significantly higher upon treatment of pre-formed biofilm

cultivated in AUM with NaIO4 (Figure 4-13B). Our result shows that treatment

of pre-formed biofilm with NaIO4 upon cultivation in TSB and AUM could

possibly effect the polysaccharides in the biofilms of K. pneumoniae, even

though, no reduction of pre-formed biofilm was observed upon the treatment.

This is clearly shown by the highly significant higher CV absorbance value

post-treatment of pre-formed biofilm with NaIO4 for all the strains upon

cultivation in TSB as well as upon cultivation in AUM except for strain Kpneu_1

and Kpneu_110. Further explanation is discussed in the discussion.

153 |

Figure 4-13 Biofilm disruption assays in the presence of 10 mM NaIO4.

The graph shows biofilm disruption assays of pre-formed biofilms cultivated in A) TSB and B) AUM and treated with 10 mM NaIO4 at 4 °C for

24 h. Bars show average absorbance readings using the CV assays from three independent experiments. The data was analysed by a Mann-

Whitney non-parametric test. Error bars indicate standard deviation of nine replicates. A highly significant increase of CV absorbance readings

can be seen upon treatment of pre-formed biofilm of all the strains cultivated in TSB with 10 mM NaIO4 in comparison to controls. Upon cultivation

in AUM, the CV absorbance readings of pre-formed biofilms of Kpneu_63, Kpneu_64, Kpneu_65, C3091, and CG43S3 was found to be

significantly increased compared to controls. Statistical significance is indicated by * represents p < 0.05, ** represents p < 0.01, *** represents

p < 0.001 and **** represents p < 0.0001.

154 |

4.3.5.4 Effect of cellulase on biofilm formation and on 24 h

pre-formed biofilms

Since our data shows that NaIO4 could possibly affect polysaccharides in

biofilms of K. pneumoniae, we then tried to identify the polysaccharides.

Cellulase was tested to determine the effect on biofilm formation by

K. pneumoniae and on pre-formed biofilms. Biofilm inhibition assays in the

presence of 0.1 % cellulase and treatment of pre-formed biofilms with

0.1 % cellulase were conducted. The biofilm inhibition assays in the presence

of 0.1 % cellulase (Figure 4-14A and Figure 4-14B) showed that biofilm

formation by C3091 upon cultivation in TSB and biofilm formation by Kpneu_63

and Kpneu_110 upon cultivation in AUM were significantly increased in the

presence of 0.1 % cellulase compared to controls. No significant reduction in

biofilm formation was observed upon growth in both media in the presence of

0.1 % cellulase. The treatment of pre-formed biofilms with 0.1 % cellulase was

observed to significantly disrupt the pre-formed biofilm of Kpneu_64 and

Kpneu_65 upon cultivation in TSB, and pre-formed biofilm of Kpneu_63 and

Kpneu_64 upon cultivation in AUM in comparison to the controls (Figure 4-14C

and Figure 4-14D).

155 |

156 |

Figure 4-14 Biofilm inhibition and disruption assays in the presence of 0.1 % cellulase.

The graph shows biofilm inhibition assays of K. pneumoniae cultivated in A) TSB and B) AUM in the presence of 0.1 % cellulase incubated at

37 °C for 24 h, and disruption of pre-formed biofilms cultivated in C) TSB and D) AUM upon treatment with 0.1 % cellulase at 45 °C for 24 h. Bars

show average absorbance readings using the CV assays from three independent experiments. The data was analysed by a Mann-Whitney non-

parametric test. Error bars indicate standard deviation of nine replicates. Biofilm formation of C3091 upon cultivation in TSB and biofilm formation

of Kpneu_63 and Kpneu_110 upon cultivation in AUM were observed to be significantly increased in the presence of 0.1 % cellulase compared

to controls. The biofilm disruption assays demonstrated that pre-formed biofilms of Kpneu_64 and Kpneu_65 in TSB were significantly disrupted

as well as pre-formed biofilm of Kpneu_63 and Kpneu_64 in AUM upon treatment with 0.1 % cellulase in comparison to the controls. Statistical

significance is indicated by * represents p < 0.05, ** represents p < 0.01 and *** represents p < 0.001.

157 |

4.4 Discussion

In the study described in this chapter, the components of K. pneumoniae

biofilms EPS were investigated upon cultivation in two different media, TSB

and AUM in 96-well microtitre plates. The planktonic growth kinetics of all the

strains were determined prior to biofilm assays. The mid-log phase in TSB

was observed about 2 h into growth and was equal to ~109 CFU/ml. Our data

is in accordance with several previous findings that measure the growth rate

of K. pneumoniae upon cultivation in TSB. For example, studies by Dulek, et

al. (2014) and Tsao, et al. (2015) mentioned that K. pneumoniae reached mid-

log phase after 2 h incubation at 37 °C in TSB and that an OD600 of 1 was

equal to 109 CFU/ml, respectively. In addition, less growth of the

K. pneumoniae strains was observed upon cultivation in AUM compared to

TSB and the duration for the bacteria to reach the mid-log phase was longer

in AUM than in TSB.

The AUM was developed by Brooks, et al. (1997). This AUM was designed to

correspond to the median values for electrolytes found in diluted human urine.

We did a modification by supplementing a very low concentration of glucose

and other sugar components such as sucrose and lactose, some additional

amino acids, and vitamins.

Since K. pneumoniae is a urease-producing species, growth limitation also

could probably be due to the catabolism of urea that causes a rise in pH and

slows down the growth of the K. pneumoniae strains used in this study (Tsuji,

et al. , 1982). Since all of the tested strains harbour a ure operon (data not

158 |

shown), they were predicted to catalyse the hydrolysis of urea (H2NCONH2) to

ammonia (NH3)2 by producing urease. Further production of ammonium

(NH4)2 ions would increase the pH of the AUM. At alkaline pH, trivalent

phosphate ions (PO4)3- are formed in urine, which then combine with water

(H2O), ammonium ions (NH4)+, and magnesium ions (Mg+) to form magnesium

ammonium phosphate (MgNH4PO4.H2O) or struvite. The trivalent phosphate

ions (PO4)3- could also combine with calcium ions (Ca2+) and CO32- to produce

carbonate apatite (Ca10(PO4)6CO3. Both of these compounds result in the

formation of calculi in catheters (Thomas, et al. , 2008).

In this study, we hypothesised that cellulose and PNAG might be amongst the

polysaccharides that play a role in K. pneumoniae biofilms (Chen, et al., 2014;

Wang, et al., 2016). In order to confirm the existence of the genes that encode

for the proteins that are involved in the synthesis of these polysaccharides in

the genome of all tested strains, we analysed and conducted a comparative

analysis of these polysaccharide biosynthetic gene clusters. The operons

examined were the bcs cluster which contains three operons and the pga

operon. The proteins encoded by the genes in bcs operon are established to

contribute to cellulose biosynthesis (Römling, et al., 2015) while the proteins

encoded by the genes in pga operon are known to be involved in synthesis of

PNAG. Cellulose biosynthetic gene clusters consist of three operons, bcsI,

bcsII and bEFG, while the PNAG biosynthetic gene cluster consists of one

pgaABCD operon. Apart from K. pneumoniae, cellulose was also shown to be

involved in biofilm formation of other bacterial species, e.g. E. coli. A previous

study showed that an E. coli 1094bcsC mutant lost its capacity to form biofilm

in microfermentors (Da Re, et al. , 2006).

159 |

K. pneumoniae has two distinct bcs operons (Römling, et al., 2015) and one

divergent operon that are important for cellulose biosynthesis. BcsO which is

encoded by bcsO is a proline-rich protein with unknown function. bcsQ

encodes for BcsQ protein which is involved in cellular localisation and might

play an important role in cellulose biosynthesis (Le Quéré, et al. , 2009). The

first two genes of bcs operons, bcsA and bcsB encode for BcsA (cellulose

synthase catalytic subunit A [located in the inner membrane]) and BcsB

(cellulose synthase subunit B [spanning the periplasm and a part of inner

membrane]) (Morgan, et al. , 2014). These proteins were shown to play an

important role for Bcs machinery and exist in all of the other bacterial bcs

operons (Solano, et al. , 2002). BcsA is an inner membrane protein, which

binds to the UDP-glucose. The other two proteins, BcsC (cellulose synthase

subunit C [located in the outer membrane and apart of inner membrane]) and

BcsD (cellulose synthase subunit D [located in the periplasm]) are involved in

exporting glucan molecules and arrangement of the Bcs complex prior to

exportation. In addition, bcsZ encodes for endo-β-1,4-glucanase (located in

periplasmic), which participates in the hydrolysis of β-D-glucan rather than

synthesis and its role in this cellulose biosynthesis is still speculative. Apart

from its role in hydrolysis, a study by Ahmad, et al. (2016) also showed that

BcsZ, could block cellulose production and subsequently reduce biofilm

formation and increased virulence of S. enterica ser. Typhimirium. The

divergent operon, bcsEFG is also postulated to be involved in the production

of cellulose. bcsE encodes for BcsE that was shown to bind to c-di-GMP and

contribute to the increased level of cellulase production (Fang, et al. , 2014)

and bcsG encodes for putative endoglucanase (Feng, et al. , 2013). While,

160 |

the function of the protein encoded by bcsF is still unknown (Ahmad, et al.,

2016).

Bioinformatic analysis revealed that one of the weak biofilm formers, Kpneu_1

had defects in bcsB_I and bcsC_I. These defects were caused by a 5 bp

insertion and a deletion of 52 bp in bcsB_I and bcsC_I respectively

(Figure 4-4 and Figure 4-5). These would result in a truncated BcsB_I and

BcsC_I proteins. Due to the fact that BcsB and BcsC proteins play an

important role in cellulose synthesis, the differences in the bcs operon I of

Kpneu_1 compared to the other strains could be why it formed less biofilms.

Other than a truncated BcsO in CG43S3, and truncated BcsB_I and BcsC_I in

Kpneu_1, we also identified truncated BcsQ in Kpneu_63 and Kpneu_64. We,

however, were unable to confirm the relationship between truncated BcsQ and

cellulose biosynthesis in K. pneumoniae since very little information is known

regarding mechanisms of cellulose synthesis in K. pneumoniae and this needs

further experiment to confirm its function.

Since Kpneu_110 is a UTI isolate, it is unclear why Kpneu_110 had less

capacity to form biofilm but could grow very well, at about the same rate as

Kpneu_64 upon cultivation in AUM (Figure 4-9). Another isolate, C3091, which

was originally isolated from a UTI patient, however, was seen to have more

biofilm upon cultivation in AUM compared to TSB. We hypothesised that AUM

might not be suitable for Kpneu_110 to form biofilms compared to in human

urine. Human urine is known to have some natural constituents including

hormones, iron chelators, pyrophosphates (Ipe, et al. , 2016) and proteins

which could promote the growth of K. pneumoniae and facilitate biofilm

161 |

formation by this pathogen in the urinary tract, which are lacking in AUM. As

human urine has its own natural proteins that do not exist in AUM, this could

be one of the reasons that resulted in less capacity of several strains such as

Kpneu_1 and Kpneu_110 to form biofilms under laboratory conditions in this

artificial environment. For example, Tamm-Horsfall protein (THP) also known

as uromodulin, is the most abundant protein in urine. THP is involved in natural

resistance of humans to uropathogenic bacteria. THP has been shown to bind

to > 90 % of E. coli isolates from UTI patients (Wu, et al. , 1996), and it can

bind specifically to type 1 fimbriated E. coli (Pak, et al. , 2001). THP is a

polymeric glycoprotein which is produced in the thick ascending limb of the

Loop of Henle in renal tissue (Serafini-Cessi, et al. , 2003). Uropathogenic

bacteria have been shown to bind to THP and as a result are flushed from the

human urinary tract by urination (Dou, et al. , 2005). A previous study has also

shown that THP enhances the adherence of E. coli on silicone and latex

catheters (Raffi, et al. , 2012). Lack of THP in AUM could be the reason why

the biofilm formation by Kpneu_1 and Kpneu_110 was found to be low in AUM.

Furthermore, since UTIs cases are not always associated with biofilms and

catheters, it is not necessarily for this strain to have a high capacity to form

biofilms in AUM. This could be another reason why the biofilm formation by

Kpneu_110 was found to be very low in AUM, even though it was isolated from

UTI patient.

Little is known regarding the biofilm matrix composition of K. pneumoniae in

different media. We investigated some of the components of K. pneumoniae

biofilms upon cultivation in TSB and AUM by conducting a biofilm inhibition

and disruption assays. Our data showed that 100 µg/ml of proteinase K

162 |

inhibited biofilm formation by Kpneu_1 upon cultivation in TSB and biofilm

formation by Kpneu_63, Kpneu_65, C3091, and CG43S4 upon cultivation in

AUM. Intriguingly, biofilm formation by Kpneu_110 was significantly increased

after cultivation in TSB in the presence of proteinase K. The supportive effect

of proteinase K on Kpneu_110 biofilms was very significant and to the best of

our knowledge, no publications have shown such an effect with

K. pneumoniae. Why this occurred is unknown but we speculate that it could

be that the proteinase K possibly cleaves self-secreted extracellular dispersion

enzyme or signalling molecules resulting in increased biofilm formation by

Kpneu_110. Several previous publications have described a number of self-

secreted extracellular dispersion enzymes by various pathogens, for example,

β-N-acetylglucosaminidase or dispersin B (DspB) that is produced by

A. actinomycetemcomitans. This enzyme was shown to degrade PNAG

structure that facilitates the cell-cell cohesion in biofilms (Kaplan, et al., 2003).

A study by Gilan, et al. (2013) also showed the supportive effect of

proteinase K on biofilm formation. 1 mg/ml of proteinase K was shown to

enhance biofilm formation by Rhodococcus ruber upon growth in minimal salts

medium, increased multi-layered structure of R. ruber biofilms.

We also treated the 24 h pre-formed biofilm with proteinase K to determine

whether protein is a major constituent of K. pneumoniae biofilms. Apart from

Kpneu_110 and CG43S3, treatment of pre-formed biofilm with

100 µg/ml of proteinase K was found to disrupt the pre-formed biofilms of the

rest of the strains upon cultivation in TSB, and also strain Kpneu_65 and

Kpneu_110 upon cultivation in AUM. Our findings are in agreement with a

previous finding which suggested that protein content in EPS of

163 |

K. pneumoniae strain Kp703 biofilms was shown to be relatively high at about

65 %, in comparison to the other EPS components such as eDNA and PNAG

upon cultivation in MH broth (Bandeira, et al., 2017).

Based on these results, proteins could possibly play a role in initial adhesion

or biofilm structure of K. pneumoniae in both media. The biofilms formation

was shown to be inhibited and the mature biofilms of most strains were

disrupted upon treatment with proteinase K. We hypothesised that 100 µg/ml

of proteinase K might degrade the proteinaceous appendages on

K. pneumoniae surface or other matrix components that act as adhesins to

attach on the surfaces.

Besides proteins, extracellular DNA (eDNA) has also been established to be

structural components of the biofilm matrix of many Gram-negative and Gram-

positive bacteria (Okshevsky, et al. , 2014). A study by Whitchurch, et al.

(2002), for example, demonstrated that addition of DNase I in M63 medium

supplemented with 1 mM MgCl2 and 0.5 % casamino acid prevented biofilm

formation by P. aeruginosa in microtiter plates. The results presented in this

chapter show that eDNA involvement in biofilm formation by K. pneumoniae is

strain dependent. On one hand, DNase treatment resulted in increased biofilm

formation by Kpneu_63 and C3091 upon cultivation in TSB. On the other

hand, DNase was shown to inhibit biofilm formation by Kpneu_65 upon

cultivation in AUM.

We also investigated the effect of DNase on 24 h old biofilms of

K. pneumoniae. The data showed that eDNA involvement in biofilm EPS of

164 |

K. pneumoniae is also strain dependent. Pre-formed biofilms of C3091 were

disrupted upon treatment with DNase compared to the untreated groups. This

observation is in line with a previous publication by Tetz, et al. (2009) that

showed that treatment of 24 h pre-formed biofilm of K. pneumoniae with

5 µg/ml of DNase I disrupted the biofilm structure and increased the

penetration of antibiotics to the cells. Strain Kpneu_110 biofilms, however,

showed a slightly higher but statistically significant biofilm level upon CV

staining compared to untreated controls upon cultivation in TSB and the

reason for this is unclear.

Other than proteins and eDNA, we also investigated the role of

polysaccharides in K. pneumoniae biofilms. Upon treatment of pre-formed

biofilms grown in TSB with NaIO4, a highly significant increase of CV

absorbance readings was observed in comparison to the controls for all the

strains used in this study. Apart from Kpneu_1 and Kpneu_110, a highly

significant increase was also observed upon treatment of pre-formed biofilms

by NaIO4 for the rest of strains compared to controls upon cultivation in AUM.

We speculated the reason for the higher readings of CV absorbance for the

treated group could possibly be because of the chemical reaction of cleaved

polysaccharide complexes with CV. NaIO4 reacts by oxidising the carbons

bearing vicinal OH group. For examples, NaIO4 cleaves the C3-C4 bonds of

N-acetyl-glucosamine (Chaignon, et al. , 2007) and C2-C3 bonds of cellulose

structure (Figure 4-15) (D. Chen, et al. , 2016). This reaction results in a

modified chain and open ring with two aldehyde groups in the polysaccharides.

The details for the mechanism of periodate oxidation is shown in Appendix 1.

165 |

Figure 4-15 Periodate oxidation of cellulose monomer.

The figure shows an example of the periodate oxidation reaction of cellulose by NaIO4

on its glucose monomer. The dialdehyde groups are hypothesised to react with CV

and presumably there was increased binding of CV to the dialdehyde groups because

of this reaction. The figure was drawn using ChemDraw Professional 16.0.1.

CV molecules could possibly react with the aldehydes groups increasing CV

binding compared to non-oxidised polysaccharides. The possible reaction

mechanism between CV molecules and the aldehyde groups of the oxidised

polysaccharides is shown in Appendix 2. Since Kpneu_1 and Kpneu_110

were shown to have a very low capacity to form biofilms in AUM, they were not

affected as much as other strains. This is possibly because less

polysaccharides were produced by these two strains in AUM and there was

less substrate for NaIO4 to cleave, resulting in less chemical reaction between

the open rings of polysaccharides with CV. We also speculated that both of

these strains produced less polysaccharides in AUM compared to in TSB,

since we observed a very significant difference of the treated group in

comparison with the control group upon cultivation in TSB. However, no

differences could be seen between those groups upon cultivation in AUM. In

general, our findings showed that polysaccharides possibly play a role in

pre-formed biofilms of examined K. pneumoniae strains.

166 |

Since we could observe an increased in the CV absorbance value upon

treatment of pre-formed biofilms with NaIO4, we speculated that NaIO4 could

possibly oxidise the polysaccharides in biofilms of K. pneumoniae. To test the

hypothesis that polysaccharide could play a role in pre-formed biofilms by

K. pneumoniae, we investigated a possible role for one of the polysaccharides

that is suggested to be involved in biofilm formation by K. pneumoniae

(Huertas, et al., 2014). Cellulose was chosen as one of the potential

polysaccharides that could be produced by K. pneumoniae since four of our

strains were shown to have the complete bcs operons (Figure 4-1).

Cultivating the tested strains in TSB in the presence of 0.1 % cellulase showed

that cellulase induced biofilm formation by C3091. Two of the strains

(Kpneu_63 and Kpneu_110) also showed a significantly higher biofilm

formation upon the cultivation in AUM in the presence of 0.1 % cellulase. Even

though the rest of the five strains did not show any significant differences, but

the biofilm formation of these five strains was found to be relatively higher upon

cultivation in AUM in the presence of 0.1 % cellulase compared to control

groups. In addition, the growth of all of the strains used in this study was also

increased upon cultivation in AUM in the presence of 0.1 % cellulase upon

incubation for 24 h (data not shown).

We speculated that small amount of sorbitol which acted as preservative in the

cellulase solution could act as an extra carbon source (Bren, et al. , 2016) in

the AUM other than small amount of glucose, sucrose and lactose in AUM

itself. This hypothesis was made due to the fact that K. pneumoniae could

metabolise D-sorbitol (Brisse, et al., 2006). Furthermore, all of our

167 |

K. pneumoniae strains were shown to have a complete gutAEBDMRQ operon

which is important for D-sorbitol metabolism, upon analysis using Rapid

Annotation using Subsystem Technology (RAST) software (Aziz, et al. , 2008).

Figure 4-7 revealed that all of the strains had a complete predicted pga operon

which consists of four genes, pgaA, pgaB, pgaB, and pgaD. The pga operon

of the K. pneumoniae strains used in this study had similarity with the pga

operon of E. coli MG1655 (Chen, et al., 2014) and the genes in pga operon of

E. coli are known to contribute to biofilm formation by E. coli (Wang, et al.,

2004). Although we investigated the role of proteins, eDNA, and

polysaccharides such as cellulose in biofilm formation by K. pneumoniae, we,

however, have not confirmed the role of PNAG as one of the components of

K. pneumoniae biofilm EPS. Having the data that all the strains have the

complete pgaABCD operon, it suggested that K. pneumoniae could possibly

produce PNAG. Since PNAG has been shown to be involved in biofilm

formation by K. pneumoniae in the presence of bile salts (Chen, et al., 2014),

it would be beneficial to determine if PNAG has a role in biofilms in the

nutritious and nutrient-poor media. The role of PNAG in biofilm formation and

pre-formed biofilms can be tested by using the enzyme

β-acetylglucosaminidase, also known as dispersin B, and this should be

considered as a future direction for this research. This enzyme cleaves β-1,

6-glycosidic bond in N-acetylglucosamine polymer (Ramasubbu, et al. , 2005).

It should be noted that the findings presented in this chapter are based on

experiments, which were conducted in 96-well microtitre plates, and it is

possible that the EPS composition of the biofilm varies in different

168 |

environments. Taken together, although there were some common EPS

components produced by multiple strains of K. pneumoniae, our data

suggested that EPS composition and K. pneumoniae biofilms are strain

dependent. It not only varies between the bacterial strains but is also

dependent on the growth media.

4.5 Conclusion

All of the strains used in this study had a complete pga operon but several

strains were shown to have defects in genes in the bcsI operon. All of the

examined strains had the capacity to form biofilms in TSB. Apart from strain

Kpneu_110, all the strains were also found to have the capacity to form

biofilms in AUM. However, the amount of biofilm formed was strain specific

and Kpneu_110 had the lowest capacity to form biofilms in AUM. Proteins

were found to play a role in biofilm formation and were a component of pre-

formed biofilms of most of the strains used in this study. The role of eDNA in

biofilm formation and pre-formed biofilms was found to be strain dependent.

Treatment of pre-formed biofilms with NaIO4 may have oxidised some of the

polysaccharides in the biofilm matrix of K. pneumoniae but the specific

polysaccharides involved in biofilms of K. pneumoniae are still undetermined.

We also showed that cellulase in a liquid form with sorbitol as a preservative

is not suitable for biofilm inhibition assays in AUM.

169 |

Chapter 5.0

Investigation into the

orientation of the

Klebsiella pneumoniae

fim-switch which controls

expression of type 1 fimbriae

170 |

5.0 Investigation into the orientation of the

K. pneumoniae fim-switch which controls

expression of type 1 fimbriae

5.1 Introduction

In earlier work described in this thesis in Chapter 4, proteins were identified as

one of the components of K. pneumoniae biofilm EPS. There are quite a

number of proteins, which are speculated to be involved in biofilm formation

by K. pneumoniae and among them are fimbriae. Type 1 fimbriae (T1F) and

type 3 fimbriae (T3F) are known as the major fimbriae of K. pneumoniae. We

decided to investigate the role of T1F in biofilm formation by K. pneumoniae

since a role for T3F in biofilm formation by K. pneumoniae has already been

determined (Jagnow, et al. , 2003; Johnson, et al. , 2011). We hypothesised

that T1F could also be involved in biofilm formation by K. pneumoniae upon

cultivation in TSB and AUM. Cleavage of these surface-exposed proteins by

proteinase K would probably contribute to the disruption of pre-formed biofilm

of K. pneumoniae as described in chapter 4.

5.1.1 Major fimbriae of K. pneumoniae

T1F and T3F belong to the chaperone-usher class fimbriae family (Stahlhut,

et al., 2012). The chaperon-usher (CU) pathway of pilus biogenesis involves

the membrane-embedded assembly and secretion of the adhesive tips. This

CU secretion system is commonly clustered into an operon containing a

minimum of three genes encoding for essential proteins: a major fimbrial

subunit, a chaperone, and an usher protein. These operons also regularly

171 |

consist of additional genes for structural (i.e. minor fimbrial subunits),

assembly and regulatory proteins (Nuccio, et al. , 2007).

Figure 5-1 shows the biogenesis of T1F as an example. During CU fimbriae

assembly, fimbria subunit (FimA, FimF, FimG, and FimH) enter the periplasm

in an unfolded conformation. They subsequently form 1:1 complexes with

FimC which assists the proteins to fold and prevents premature assembly of

subunits. The chaperone-subunit complexes dissociate upon interaction with

the outer membrane anchored-FimD usher protein which serves as a scaffold

by forming a pore allowing the passage of individual FimA subunits (Figure

5-1). These subunits cross the OM and become incorporated to produce fully

functional fimbriae. The three-dimensional right handed helical structure of the

fimbrial backbones are then made of ~1000 copies of FimA with a short flexible

fibrillum on top, consisting of single copies of FimF, FimG, and FimH adhesin

(Figure 5-1 and Figure 5-2) subunits, which mediates attachment to the host

cells (Busch, et al. , 2012; Vetsch, et al. , 2004).

172 |

Figure 5-1 The chaperon-usher (CU) pathway of T1F biogenesis.

A schematic drawing of the T1F chaperone usher pathway as an example. IM: inner

membrane, OM: outer membrane. The figure is drawn and adapted from Vetsch, et

al. (2004) and Korea, et al. (2011).

173 |

5.1.1.1 Type 1 fimbriae (T1F)

Type 1 fimbriae (T1F) are thin, rigid and thread-like protrusions on the cell

surface, they are approximately 7 nm in diameter and 1 to 2 µm in length

(Korea, et al., 2011) (see Figure 5-1). T1F are suggested to be one of the

important virulence factors of K. pneumoniae UTIs and they are wide spread

among the members of Enterobacteriaceae (Struve, et al., 2008). These

mannose-sensitive surface organelles are the major adhesive fimbriae

structures in K. pneumoniae and mediate binding of this pathogen to

D-mannosylated containing receptors (Murphy, et al., 2013), for example, on

bladder urothelial cells (Jorgensen, et al. , 2012) and Tamm-Horsfall

glycoproteins via the T1F adhesin, FimH (Rosen, Pinkner, Walker, et al. ,

2008). The expression of these surface appendages has been shown to

mediate the initial adhesion of K. pneumoniae, allowing it to colonise the

uroepithelium and eventually cause UTIs (Stahlhut, et al., 2012; Struve, et al.,

2008). T1F has been shown to play a pivotal role in the pathogenesis of UTIs

and it has been experimentally demonstrated in a murine bladder infection

model (Struve, et al., 2008).

174 |

Figure 5-2 The structure of T1F mannose-sensitive fimbriae.

FimA is the major subunit, and FimH is the adhesin that binds to mannosylated

glycoprotein receptors (Paczosa, et al., 2016). The figure is re-published with

permission.

175 |

The T1F operon (fim operon) consists of genes encoding for FimB and FimE

recombinases, FimA; the major fimbrial component, FimI; pilus biogenesis

protein, FimC; chaperone protein, FimD; a fimbrial usher protein, and other

genes encoding for accessory and structural FimF and FimG subunits, and for

the minor subunit FimH adhesin. This operon also includes fimK which is

located downstream of the fimH in the K. pneumoniae genome. It is predicted

to encode a protein with a DNA binding motif, and conserved EAL domain

which is associated with cyclic dimeric guanosine monophosphate (c-di-GMP)

putative phosphodiesterase (Struve, et al., 2008; Wang, et al., 2013). The T1F

gene cluster of K. pneumoniae is homologous to that of E. coli with regard to

genetic composition and regulation (Struve, et al., 2008), apart from the

existence of a unique gene, fimK. T1F are regulated in the same manner as

T1F of E. coli which is by a phase variation mechanism (Rosen, Pinkner,

Jones, et al. , 2008). This mechanism is controlled by inversion of an invertible

fim-switch (fimS), which controls the major subunit of T1F, fimA expression

(Schembri, et al., 2005; Struve, et al., 2008).

176 |

Figure 5-3 Schematic diagram of T1F operon.

The T1F operon consists of two genes that encode recombinases FimE and FimB, which control the orientation of fim-switch (fimS), seven genes

(fimA, fimI, fimC, fimD, fimF, fimG, and fimH) that encode proteins involved in assembly and secretion of T1F, fimK which encodes the regulatory

protein, FimK. FimK is composed of a putative DNA binding domain and a phosphodiesterase domain at the N-terminal and C-terminal,

respectively (Figure 5-7). The putative promoter is indicated by the angled arrows. The inverted repeat is labelled as IR.

177 |

5.1.1.1.1 Fim-switch (fimS) orientation

The regulation of T1F in K. pneumoniae is similar to that in E. coli. It is

controlled by a phase variation mechanism via inversion of an invertible DNA

element, the fim-switch (fimS). fimS is flanked by 9 bp inverted repeats (IR).

The inversion of fimS is mediated by the site-specific tyrosine-integrases, FimE

and FimB (McCusker, et al. , 2008). FimE catalyses the inversion of fimS from

the ‘ON-to-OFF’ orientation and FimB catalyse the switch of fimS to both of the

‘ON-to-OFF’ and ‘OFF-to-ON’ orientations. Inversion of the switch results in a

fimbriated cell state when it is in the ‘ON’ orientation and an afimbriated cell

state when it is in the ‘OFF’ orientation (Dorman, et al. , 2009).

The invertible DNA element, fimS contains a promoter for fimA, the major

subunit of T1F. The orientation of fimS can be determined by amplifying the

switch region using the forward and reverse primers (Figure 5-4), and digesting

the amplified PCR product with HinfI. Due to the asymmetric location of HinfI

restriction sites within fimS, it will result in different digestion patterns

depending on the orientation of fimS. The length of amplified/undigested fimS

is 817 bp. The ‘ON’ orientation of fimS results in 212 bp and 605 bp fragments

on digestion of the PCR product, while the ‘OFF’ orientation of fimS results in

496 bp and 321 bp fragments (Figure 5-4).

178 |

Figure 5-4 Schematic diagram of fimE, fimS and fimA.

Figure shows the invertible DNA element of fimS located in between fimE and fimA in

the T1F operon. The inverted repeat is labelled as IR. The promoter is indicated by

the angled arrow. The direction of the primers is as labelled. The figure is adapted

from Struve, et al. (2008).

179 |

5.1.1.2 Type 3 fimbriae (T3F)

Type 3 fimbriae (T3F) are mannose-resistant fimbriae (Alcántar-Curiel, et al.,

2013), and they are thinner than T1F, being 2 to 4 nm wide and 0.5 to 2 uM

long (Figure 5-5) (Allen, et al. , 1991; Gerlach, et al. , 1988). T3F have been

shown to play an important role in initial adhesion of K. pneumoniae to

extracellular matrix and collagen-coated surfaces (Jagnow, et al., 2003). They

attach to the basolateral surfaces of tracheal epithelial cells, human-derived

extracellular matrix proteins (ECMPs) (Jagnow, et al., 2003), and human lung

tissue membranes, and this subsequently leads to respiratory tract infections.

The T3F appear to be highly conserved in K. pneumoniae since a study by

Old, et al. (1985) showed that antisera raised against T3F from a single

K. pneumoniae isolate recognised most of the fimbriated Klebsiella.

T3F are encoded by the mrkABCDF gene cluster (Johnson, et al., 2011). mrkA

encodes the MrkA major fimbrial subunit, mrkD encodes the adhesive unit on

the tip, mrkB encodes the chaperon protein, mrkC encodes the usher protein,

and mrkF encodes the scaffolding protein. MrkB, MrkC, and MrkF proteins are

important in fimbrial assembly, structuring and stabilising the fimbriae on the

cell surfaces respectively. The expression of mrk is regulated by a complex

regulatory system, which involves a combination of a phosphodiesterase

encoded by mrkJ and a PilZ domain protein encoded by mrkH, together with

a DNA binding protein encoded by mrkI located downstream and adjacent to

the mrk operon (Murphy, et al. , 2012).

180 |

Figure 5-5 The structure of T3F mannose-resistant fimbriae.

MrkA is the major fimbrial subunit and MrkD is the mannose-resistant adhesin

(Paczosa, et al., 2016). The figure is re-published with permission.

181 |

5.1.2 Role of type 1 fimbriae (T1F) in biofilm formation by

K. pneumoniae

The T1F gene cluster of K. pneumoniae has an overall high degree of

structural resemblance with the well-studied T1F gene cluster of E. coli

(Struve, et al., 2008). As previously mentioned, there is only one difference

that can be identified between these two T1F operons; the existence of a gene

that encodes for regulatory protein, fimK, which is exclusive to K. pneumoniae

(Struve, et al., 2008) located downstream of the fimH. T1F are known to be

involved in biofilm formation of uropathogenic E. coli (UPEC) strains. As

previously mentioned, the T1F operon of E. coli has a high similarity in terms

of genetic organisation with T1F operon of K. pneumoniae except for fimK.

Since the role of T3F in biofilm formation by K. pneumoniae is well established

and T1F also was reported to play a key role as crucial as T3F in biofilm

formation by K. pneumoniae upon cultivation in human pooled urine (Stahlhut,

et al., 2012), we postulated that T1F of K. pneumoniae could also be involved

in biofilm formation by K. pneumoniae in our experimental settings.

Some studies have revealed that T1F are involved in biofilm formation by

K. pneumoniae but at the same time, several studies have proven that T1F do

not contribute to the biofilm formation by K. pneumoniae (Paczosa, et al.,

2016). For instance, T1F were shown to play a complementary role with T3F

in biofilm formation by K. pneumoniae TOP52, an acute cystitis clinical isolate

(Johnson, et al. , 2014), and contribute to bacterial adherence in a murine

catheter-associated urinary tract infection model (Murphy, et al., 2013). A

study by Stahlhut, et al. (2012) also concluded that K. pneumoniae requires

182 |

either T1F or T3F to form biofilms on urinary catheters. Absence both of the

fimbriae significantly reduced the colonisation and biofilm formation by

K. pneumoniae C3091 on catheter surfaces and complementation of the T1F

or T3F mutant strains with either one fimbrial type restores the phenotype of

the wild type. Both of the fimbriae also independently promote biofilm

formation on urinary catheters compared to nonfimbriated E. coli upon

transformation and expression of T1F or T3F gene cluster. In contrast,

isogenic mutants of K. pneumoniae C3091 (∆fim), which was developed by

Schroll, et al. (2010) suggested that T1F did not have a significant role in the

biofilm formation by K. pneumoniae upon growth in modified fastidious

anaerobe broth in flow chambers.

These observations could be because all of the experiments were conducted

in different experimental settings, models and media which could contribute to

these differences. Therefore, the exact contribution of T1F to biofilm formation

by K. pneumoniae is still unclear. In the study described in this chapter, we

investigated further, the role of T1F in biofilm formation by determining the

orientation of fimS in two different growth states, planktonic and biofilm cells

upon cultivation in two different media, highly nutritious medium, TSB and

nutrient-poor medium, AUM in 96-well microtitre plates.

183 |

Table 5-1 Involvement of fimbriae in biofilm formation by Enterobacteriaceae.

Bacterial

species

Contribution to biofilm

formation

Description Model/system Media References

E. coli 2K1056

T1F are proposed to

establish stable attachment

on abiotic surfaces, such as

polyvinyl chloride (PVC)

possibly via the interaction

between FimH and surfaces

T1F has been shown to be

essential for the initial

attachment of E. coli.

Microscopic analysis showed

that a fimH mutant strain was

dramatically defective in initial

attachment and had no cells

attached to PVC surfaces in

comparison to wild type strain,

and this subsequently

hampered biofilm formation.

PVC microtiter

plate and a tab of

PVC plastic

LB

(Pratt, et al. ,

1998)

E. coli MG1655

T1F are not essential for

initial adhesion on glass

surfaces, however, they are

The fimA mutant strain showed

less capacity to form biofilms

in both of the media tested in

96-well microtitre

plate

M63 and LB

media

supplemented

with different

(Rodrigues, et al. ,

2009)

184 |

Bacterial

species

Contribution to biofilm

formation

Description Model/system Media References

crucial for development of

mature biofilm

comparison the wild type and

complementation strains.

concentration

of mannose

E. coli UTI89

The UTI89 cells harvested

from a 20 h old biofilm grown

on silicone surfaces were

found to positively expressed

T1F. Expression of T1F led

to increase initial adhesion

and essential for early biofilm

stage

The bacteria were inoculated

onto silicone surfaces and

cultivated in a flow of pooled

human urine

Flow chamber

mounted with

silicone sheets

Pooled human

urine

(Staerk, et al. ,

2016)

E. coli

MG1655

Absence of T1F attenuated

the colonisation of E. coli on

catheters during cultivation in

a catheterised bladder model

to mimic CAUTI in vitro. This

showed that T1F are required

The wild type and fim mutant

strains were inoculated in

separate catheterised bladder

models. The model system

inoculated with the mutant

strain had less cell numbers

Catheterised

bladder model

Pooled human

urine

(Reisner, et al. ,

2014)

185 |

Bacterial

species

Contribution to biofilm

formation

Description Model/system Media References

for catheter colonisation in

vitro.

In addition, biofilm cells of

E. coli isolated from the

catheterised patient

predominantly had cells with

fimS in the ‘ON’ orientation.

recovered from the catheter tip

and internal catheter surface

compared to the model system

that was inoculated with the

wild type strain.

K. pneumoniae

C3091

T1F and T3F were shown to

be coordinately controlled

during the biofilm formation of

K. pneumoniae on catheter

pieces.

Biofilm cells on catheter were

shown to have fimS in the ‘ON

and OFF’ orientation

compared to fimS in planktonic

cells which was in the ‘OFF’

orientation

Catheterised

bladder model

Pooled human

urine

(Stahlhut, et al.,

2012)

K. pneumoniae

C3091

T3F but not T1F were shown

to be involved in the biofilm

The biofilm of wild type strain

were found to have cells with

Flow chamber

Fastidious

anaerobe broth

(Schroll, et al.,

2010)

186 |

Bacterial

species

Contribution to biofilm

formation

Description Model/system Media References

formation on flow chamber fimS in the ‘OFF’ orientation

demonstrating that T1F are

down-regulated in biofilm cells.

T1F mutant also formed as

much biofilms as wild type

strain

with glucose

and IPTG

K. pneumoniae

TOP52

Defined T1F and T3F

mutants were constructed

and were then tested on the

capacity to form biofilm in a

murine CAUTI model.

Both T1F and T3F are

experimentally proven to be

involved in biofilm formation by

K. pneumoniae and bacterial

adherence in a murine CAUTI

model.

Murine

(Murphy, et al.,

2013)

187 |

5.2 Materials and Methods

5.2.1 DNA isolation from planktonic and biofilm cells

Biofilms were grown in microtitre plates according to the protocol in

Chapter 3, section 3.2.1. After 24 h incubation at 37 °C under shaking

conditions, planktonic cells were removed from 96-well microtitre plates

and transferred into a microcentrifuge tube. The cells were centrifuged at

15700 x g (Eppendorf 5415D) for 15 min and the supernatant was

decanted. The pellet was resuspended in 100 µl of molecular grade water.

These samples were labelled as planktonic cells. For the isolation of DNA

from biofilm cells, the wells (without planktonic cells) were washed

carefully with molecular grade water to remove the non-adherent cells.

Biofilm cells were scraped from the bottom and wall of the 96-well

microtitre plates using pipette tips and resuspended in 50 µl molecular

grade water in a microcentrifuge tube. The suspension was then

centrifuged at 15 700 x g (Eppendorf 5415D) for 15 min and the

supernatant was decanted and resuspended in 30 µl molecular grade

water. These samples were labelled as biofilm cells. The resuspended

pellets of planktonic cells and biofilm cells were boiled for 5 min in a

heating block, and quickly transferred into ice. The suspension was

centrifuged at 15 700 x g (Eppendorf 5415D) for 5 min. The supernatant

was then used as a DNA template for amplification of the fimS region.

188 |

5.2.2 Determination of the invertible element (fimS) orientation

The orientation of the K. pneumoniae invertible DNA element containing the

fimA promoter was determined using PCR-based assays. Extracted DNA was

used as a template following a previously described method by

Struve, et al. (2008) with some modifications. Primer fimF

(GGGACAGATACGCGTTTGAT) and fimR (GGCCTAACTGAACGGTTTGA)

were used to amplify 817 bp of the fimS region. PCR was performed according

to the protocol mentioned in Chapter 2, section 2.4.3. The PCR product was

separated on a 1 % agarose gel by electrophoresis. The orientation of fimS

was determined by digesting the PCR product using HinfI (NEB, UK) at 37 °C

in a heating block for one hour and deactivated at 80 °C for 20 min. The

digested product was then separated on a 1.2 % agarose gel by

electrophoresis.

5.2.3 Bioinformatic analysis of the T1F operon

The DNA sequences of the genes in the T1F operon for all isolates were

obtained from their corresponding draft genome sequences. The fim operon

alignment was done using GeneDoc version 2.7 (Pittsburgh, USA) and Clustal

Omega. Deduced amino acid sequences for all the T1F genes were obtained

using EMBOSS-Transeq (Rice, et al., 2000).

189 |

5.3 Results

5.3.1 Determination of fimS orientation

To investigate the influence of growth state and growth environment on T1F

phase switching, the orientation of the invertible DNA element containing the

promoter of fimA, was compared between the cells harvested from planktonic

and biofilm cells upon cultivation in TSB and AUM under shaking conditions.

Digestion of the PCR amplified fimS region, using HinfI, harvested from the

planktonic cells of strain Kpneu_1, Kpneu_63, Kpneu_64, Kpneu_65,

Kpneu_110, C3091, and CG43S3 grown in TSB resulted in two fragments that

were of expected sizes corresponding to the fimS being in the ‘OFF’ position.

The orientation of the fimS obtained from the biofilm cells of strain Kpneu_1,

Kpneu_65, C3091, and CG43S3 was in the ‘ON and OFF’ orientation in TSB.

The biofilms of the other three strains, Kpneu_63, Kpneu_64, and Kpneu_110,

however, showed the ‘OFF’ orientation for fimS when cultured in TSB (Table

5-2).

We also determined the orientation of fimS by digestion of the PCR amplified

fimS region obtained from planktonic and biofilm cells of K. pneumoniae

cultivated in AUM. The orientation of fimS in all strains was in the ‘OFF’

orientation in both planktonic and biofilm cells (Table 5-2). The differences of

HinfI digestion pattern of amplified fimS region in biofilm cells cultivated in TSB

and AUM can be seen in Figure 5-6.

190 |

Table 5-2 Orientation of fimS in planktonic and biofilm cells of K. pneumoniae.

The orientation of fimS in planktonic cells of all strains used in this study was in the

‘OFF’ orientation upon cultivation in both of TSB and AUM. Upon cultivation in TSB,

the orientation of fimS in biofilm cells of Kpneu_1, Kpneu_65, C3091, and CG43S3

was in the ‘ON and OFF’ orientation, while the orientation of fimS in biofilm cells of

Kpneu_63, Kpneu_64, and Kpneu_110 was in the ‘OFF’ orientation. However, all of

the strains were found to have biofilm cells with fimS being in the ‘OFF’ orientation

upon growth in AUM. The orientation of fimS in biofilm cells is boxed in red.

Strain Upon cultivation in TSB Upon cultivation in AUM

Planktonic

cells

Biofilm cells Planktonic

cells

Biofilm cells

Kpneu_1 OFF ON and OFF OFF OFF

Kpneu_63 OFF OFF OFF OFF

Kpneu_64 OFF OFF OFF OFF

Kpneu_65 OFF ON and OFF OFF OFF

Kpneu_110 OFF OFF OFF OFF

C3091 OFF ON and OFF OFF OFF

CG43S3 OFF ON and OFF OFF OFF

191 |

Figure 5-6 Determination of fimS orientation in biofilm cells.

Detection of fimS orientation in biofilm cells upon cultivation in A) TSB and B) AUM

was conducted by digestion of fimS amplified region by HinfI. Kpneu_1, Kpneu_65,

C3091, and CG43S3 showed four different digestion bands that corresponding to

fimS being in the ‘ON and OFF’ orientation upon cultivation in TSB. While, Kpneu_63,

Kpneu_64, and Kpneu_110 showed a different digestion pattern which corresponded

to fimS being in the ‘OFF’ orientation upon cultivation in TSB (Figure 5-6A). The

biofilms of all strains were observed to have a digestion pattern that corresponded to

fimS being in the ‘OFF’ orientation (Figure 5-6B) upon cultivation in AUM.

M: HyperLadder I (Bioline, UK), U: undigested fimS amplified regions and D: digested

fimS amplified region by HinfI. The strains with ‘ON and OFF’ orientation of fimS in

their biofilm cells are indicated with red boxes. The fimS region is indicated with

yellow arrows, the ‘OFF’ orientation is labelled with red arrows, and the ‘ON’

orientation is labelled with blue arrows.

A)

B)

192 |

5.3.2 Analysis of the T1F operon sequences

Since the orientation of the fimS was in the ‘ON and OFF’ orientation in biofilm

cells of strains Kpneu_1, Kpneu_65, C3091, and CG43S3 and in the ‘OFF’

orientation in biofilm cells of strains Kpneu_63, Kpneu_64, and Kpneu_110,

we hypothesised that a difference in the T1F operons between strains may

account for this finding. For the purpose of explanation, Kpneu_1, Kpneu_65,

C3091, and CG43S3 will be mentioned as the ‘ON and OFF’ group while

Kpneu_63, Kpneu_64, and Kpneu_110 will be mentioned as the ‘OFF’ group

in the further description and discussion.

The sequences of the T1F operon of all strains were obtained from their

corresponding draft whole genome sequences. The length of this operon was

about 10,255 bp. Sequences were analysed using GeneDoc to obtain multiple

sequence alignments and EMBOSS-Transeq for amino acid translation.

Based on the multiple sequence alignments, various SNPs were identified in

several genes of the fim operon, and they resulted in amino acid variations

(Appendix 3). Among all of these amino acid variations, some of them were

specific for the ‘ON’ and OFF’ group (Kpneu_1, Kpneu_65, C3091, and

CG43S3), and some for the ‘OFF’ group (Kpneu_64, Kpneu_64, and

Kpneu_110). These specific SNPs were located in fimF, fimI and fimK and

these SNPs resulted in the amino acid substitutions of FimF, FimI and FimK.

(Table 5-3). The SNPs in fimF resulted in the amino acid substitutions at

position A38T and A106T (Appendix 10). This substitution is a conservative

substitution. Other than fimF, three SNPs were identified in fimI (Appendix

11). This SNPs resulted in two conservative amino acid substitutions at

193 |

position R44H and A136T, and one non-conservative amino acid substitution

at position P146R. The most remarkable SNP between the ‘ON and OFF’

group (Kpneu_1, Kpneu_65, C3091, CG43S3) and the ‘OFF’ group

(Kpneu_63, Kpneu_64, and Kpneu_110) was in fimK which resulted in a

premature termination codon at position 441 (Q441*) in the ‘OFF’ group

(Figure 5-7). This substitution was conserved in the ‘OFF’ group, Kpneu_63,

Kpneu_64, and Kpneu_110. Several SNPs were also identified in the

intergenic regions between fimB and fimE (Appendix 4, Appendix 5, and

Appendix 6) and fimS (Appendix 7, Appendix 8, and Appendix 9) between

these two groups. Six SNPs were identified in the intergenic region between

fimB and fimE at position T-61C, A-63T, A-204G, G-247A, A-257G, and

T-265G for the ‘OFF’ group. These positions are based on the start codon of

fimE, ATG (Appendix 6). While two SNPs were identified in the fimS region at

position of A-321C, and A-204C in the upstream of fimA for the ‘OFF’ group.

The position is based on the fimA start codon GTG (Appendix 9). Analysis of

the intergenic regions indicates that the intergenic SNPs probably do not affect

any putative promoter or rho-independent terminator sites in these regions.

194 |

Table 5-3 Amino acid substitutions identified in FimF, FimI, and FimK.

The amino acid substitutions were caused by SNPs identified in the ‘ON and OFF’ (Kpneu_1, Kpneu_65, C3091, and CG43S3) and the ‘OFF’

(Kpneu_63, Kpneu_64, and Kpneu_110) groups.

Indicators:

Alanine to threonine (A→T): Neutral substitution, aliphatic non-polar to polar residue

Arginine to histidine (R→H): No changes in terms of charge. Both are positively charge residues.

Proline to arginine (P→R): Aliphatic non-polar residue to positively charged residue

Glutamine to STOP (Q→*): Polar residue to stop codon

Amino acid

change

Kpneu_1

Kpneu_63

Kpneu_64

Kpneu_65

Kpneu_110

C3091

CG43S3

FimF A38T A T T A T A A

A106T A T T A T A A

FimI R44H R H H R H R R

A136T A T T A T A A

P146R P R R P R P P

FimK Q441* Q * * Q * Q Q

195 |

196 |

Figure 5-7 Multiple sequence alignment of FimK.

The multiple amino acid sequence alignment of FimK was conducted using GeneDoc. The amino acid substitutions resulted from the SNPs are

boxed in red. The predicted helix-turn-helix (HTH) DNA-binding domain is boxed in blue. The HTH domain region was predicted using an HTH

prediction software (https://npsa-prabi.ibcp.fr) (Combet, et al. , 2000). The predicted EIL/EAL domain, which is underlined with purple. This

domain is proposed to act as a diguanylate phosphodiesterase. The domain is predicted using PROSITE (http://prosite.expasy.org). The ‘ON

and OFF’ and ‘OFF’ groups are as indicated.

197 |

5.4 Discussion

Initial adhesion and biofilm formation of K. pneumoniae on host cell surfaces,

for example, on urothelial cells are commonly associated with two major

fimbrial adhesins, the mannose-sensitive T1F and mannose-resistant T3F

(Alcántar-Curiel, et al., 2013). Besides these two major surface appendages,

Kpc fimbriae (Wu, et al., 2010) and a homolog of the E. coli common pilus

(Ecp) (Saldaña, et al. , 2014) are also present in K. pneumoniae. Besides

these fimbriae, there is also a less investigated fimbrial adhesin of

K. pneumoniae called KPF-28 and an afimbrial adhesin named CF29K that is

encoded by genes on the R-plasmid of K. pneumoniae strain CF914-1 (Brisse,

et al. , 2009; Di Martino, et al. , 1995; Di Martino, et al. , 1996). The vast

majority of K. pneumoniae strains, specifically clinical isolates have the

capacity to produce T1F and T3F fimbriae suggesting that they are important

surface appendages in K. pneumoniae (Podschun, et al., 1998; Schroll, et al.,

2010). The interaction between T1F and terminally exposed mannose

residues in N-linked oligosaccharides that are normally located on cell

surfaces is facilitated by the FimH, adhesin subunit of T1F (Figure 5-1 and

Figure 5-2) (Stahlhut, et al. , 2009).

Our findings reveal that regardless of the growth media used for cultivation, all

the examined strains had planktonic cells with fimS being in the ‘OFF’

orientation, which is normally associated with no expression of T1F. The

observation of fimS being in the ‘OFF’ orientation in planktonic cells upon

cultivation in AUM is in line with the previous study by Stahlhut, et al. (2012).

Stahlhut, et al. (2012) also showed that orientation of fimS in planktonic cells

198 |

of strain C3091 was also in the ‘OFF’ orientation upon cultivation in pooled

human urine under hydrodynamic conditions. Moreover, our results showed

that four of the strains used in this study, had biofilm cells with fimS in the ‘ON

and OFF’ orientation and three strains had biofilm cells with fimS in the ‘OFF’

orientation upon cultivation in TSB. Surprisingly, all the strains had the biofilm

cells with fimS in the ‘OFF’ orientation upon cultivation in AUM.

Our observation suggested the existence of a mixed population of fimbriated

and nonfimbriated cells in the biofilm community upon cultivation in TSB. This

subsequently resulted in the detection of fimS being in the ‘ON and OFF’

orientation of biofilm cells of strain Kpneu_1, Kpneu_65, C3091, and CG43S3.

The biofilms community might consist of i) adherent cells on the polystyrene

surface which are most likely to have cells with fimS being in the ‘ON’

orientation and ii) cells with fimS being in the ‘OFF’ orientation that are just

about to disperse from the adhered community.

Our investigation of the DNA sequences of the genes in the T1F operon in the

examined strains revealed various SNPs at multiple locations. SNPs that

specifically related to the ‘ON and OFF’ group (Kpneu_1, Kpneu_65, C3091,

and CG43S3) or the ‘OFF’ group (Kpneu_63, Kpneu_64, and Kpneu_110)

were identified in fimI, fimF and fimK (Table 5-3). All of these SNPs resulted

in amino acid substitutions in FimI (Appendix 10), FimF (Appendix 11) and

FimK (Figure 5-7). There were also a number of SNPs in the intergenic region

of fimB and fimE (Appendix 4, 5, and 6) and also in the fimS located in the

intergenic region of fimE or fimA (Appendix 7, 8, and 9). Intergenic regions

may carry various important elements such as promoter sites, terminator sites,

199 |

transcription factors binding sites, integrase binding sites, regulatory protein

binding sites, pseudogenes, and inverted repeats (Sharples, et al. , 1990).

However, the effect of these SNPs on the orientation of fimS in K. pneumoniae

is still unknown. Because of this, it is worth analysing the effect of these

intergenic SNPs on the orientation of fimS in future studies as this may lead to

a better understanding of their impact on the expression of T1F in

K. pneumoniae.

Both SNPs in fimF and two SNPs in fimI resulted in conservative substitutions.

These substitutions are known to have no or smaller effects on protein function

than non-conservative substitutions (Prakash, 2008). In addition, one SNPs

in fimI resulted in non-conservative substitution. The effect of the non-

conservative substitution on fimS orientation, however, is still unknown. We

hypothesised the SNP in fimK, could probably has a significant effect since it

results in a premature termination codon which would result in a truncated

FimK for strains Kpneu_63, Kpneu_64, and Kpneu_110. Since T1F are

secreted through the chaperone-usher pathway, defects in one of the proteins

may influence the biogenesis of T1F (Figure 5-1). We speculate that the

truncated FimK protein could possibly account for fimS being in the ‘OFF’

orientation in Kpneu_63, Kpneu_64, and Kpneu_110.

Given the importance of fimK as a gene that encodes for a regulatory protein,

it might affect the orientation of fimS and interrupt the expression of fimA and

T1F. FimK consists of helix-turn-helix (HTH) putative DNA-binding domain at

the N-terminus and putative EAL/EIL domain, at C-terminus (Wang, et al.,

2013). The EAL/EIL domain is associated with phosphodiesterases which

200 |

have been implicated in hydrolysis of c-di-GMP (Schmidt, et al. , 2005). Wang,

et al. (2013) also previously demonstrated that expression of T1F in

CG43S3∆mrk∆fimK was reduced and the putative DNA-binding region at the

N-terminus of FimK was proposed to positively affect the transcription of fimA,

but not the C-terminal domain. Our findings showed a premature termination

codon was located in the C-terminal region which we speculate could affect

the orientation of fimS. The length of the intact FimK protein is predicted to be

470 aa while the length of truncated protein caused by the premature

termination codon is predicted to be 430 aa (Figure 5-7). Whilst, the mutant

strain with a deleted EIL domain (CG43S3∆mrkA∆EILfimK) in the

aforementioned study by Wang, et al. (2013) had a different length of truncated

FimK (275 aa) compared to the truncated FimK identified in our study (430 aa).

These differences in size of truncated proteins could possibly act differently on

the orientation of fimS.

Previous work by Rosen, et al. (2008) was also not in line with our hypothesis.

FimK was shown to reduce the expression of T1F in

K. pneumoniae cystitis isolate, strain TOP52 and also suggested to act as an

inhibitory factor for biofilm formation of this strain. The isogenic mutant of

TOP52, TOP52∆fimK was observed to predominantly have fimS being in the

‘ON’ orientation and higher expression levels of T1F, in comparison to its wild

type, which generally had cells with fimS being in the ‘OFF’ orientation. The

TOP52∆fimK also had the capacity to form biofilm after 48 h incubation at room

temperature but its wild type strain failed to form biofilm in such conditions.

They concluded that loss of FimK increased the c-di-GMP signalling, which

has been correlated with formation of biofilms by regulating the transition

201 |

between the sessile and motile bacterial states (Römling, 2012; Song, et al.,

2016). Since the C-terminal of FimK contains the EIL/EAL domain, which is

normally implicated with phosphodiesterases that cleave c-di-GMP, this would

be an important point to study in future.

Our findings show that all of the strains had biofilm cells with fimS being in the

‘OFF’ orientation upon cultivation in AUM. We speculated that the ‘OFF’

orientation for ‘Kpneu_63, Kpneu_64, and Kpneu_110 could be because of the

SNP that resulted in the truncated FimK. This could be the reason why

Kpneu_63, Kpneu_64, and Kpneu_110 had the biofilm cells in the ‘OFF’

orientation upon cultivation in both of the media. Thus, further discussion on

the orientation of fimS is focused on strains Kpneu_1, Kpneu_65, C3091, and

CG43S3 only.

Our results showed that fimS orientation in biofilm cells of strain C3091 upon

cultivation in TSB in 96-well microtitre plates agree with the findings by

Stahlhut, et al. (2012) which showed that the same strain, C3091 had biofilm

cells with fimS being in the ‘ON and OFF’ upon cultivation of this strain in

pooled human urine and on catheter pieces under hydrodynamic conditions.

This suggests that the findings conducted in TSB used in the study described

in this thesis mimic the experimental results which obtained in pooled human

urine (Stahlhut, et al., 2012) rather than AUM even though the experimental

environments were different in both of the studies.

The discrepancy observed between the results conducted in AUM and pooled

human urine from the previous study by Stahlhut, et al. (2012) could be due to

202 |

differences in the experimental models that we used. We harvested the biofilm

cells from 96-well microtitre plates, while Stahlhut, et al. (2012) harvested the

cells from biofilms in the in vitro bladder model. Their model would probably

provide hydrodynamic conditions that mimic the real conditions in UTI patients.

In addition, their model also would supply more fresh-medium to bacteria by

providing constant flow rate of urine medium to the culture. Since phase

variable expression is highly dependent on the physical environment,

(Khandige, et al. , 2016) this could be another reason on the different

observation from our study.

203 |

5.5 Conclusion

Our results demonstrated that fimS in biofilm cells of Kpneu_1, Kpneu_65,

C3091, and CG43S3 was in the ‘ON’ and ‘OFF’ orientation upon cultivation in

TSB. However, the orientation of fimS in the biofilm cell of these four strains

was in the ‘OFF’ orientation upon cultivation in AUM. This suggests that T1F

possibly plays a role in biofilm formation of Kpneu_1, Kpneu_65, C3091, and

CG43S3 upon cultivation in TSB, but the role of T1F in biofilm formation

conducted in AUM is still call into question in our experimental settings. Our

data also shows that fimS is in the ‘OFF’ orientation in biofilm cells of strain

Kpneu_63, Kpneu_64, and Kpneu_110 regardless of the medium that we used

to grow them. This suggests that transcription of fimA and expression of T1F

might not be involved in the biofilm formation by those strains, but expression

studies are warranted to confirm this observation. Our data show that there

was no direct correlation between the orientation of fimS, which controls the

expression of fimA, and the capacity to form biofilms by K. pneumoniae upon

cultivation in AUM. The most remarkable SNP identified upon our analysis is

Q441* which resulted in a truncation in FimK. This substitution is conserved

in the OFF group (Kpneu_63, Kpneu_64, and Kpneu_110). This is hypothesed

to be a possible reason for fimS being in the ‘OFF’ orientation upon cultivation

in TSB for these three strains.

204 |

Chapter 6.0

Role of OmpK35 and OmpK36

porins in biofilm formation by

Klebsiella pneumoniae

205 |

6.0 Role of OmpK35 and OmpK36 porins in biofilm

formation by K. pneumoniae

6.1 Introduction

Gram-negative bacteria have distinct outer membrane porin (Omp) structures,

located at the outer membrane layer of Gram-negative cell wall (Nikaido,

1994). They are important for transportation of small hydrophilic solutes (600

Da) (Nikaido, 2003). It is notable that trimeric unique water channel-forming

Omps play an important role in uptake of nutrients to the periplasmic space

through the outer membrane layer, and secretion of toxic and waste molecules

from the cytoplasm (Schulz, 2004). These porins can be divided into three

different groups: i) general porins or porins for general diffusion e.g. OmpK35,

OmpK36, and PhoE, ii) specific porins or porins with an exclusive binding site

for specific solute, for example, maltodextrins which bind to the LamB

maltoporin (Klebba, 2002) and sucrose which binds to the ScrY porins (Sun,

et al. , 2016), and iii) porins that actively transport and consume energy such

as porins which are involved in the uptake of ferric-siderophore (Danelon, et

al. , 2004). These ferric-siderophore complexes require specific porins or

receptors to penetrate the outer membrane layer since these large complexes

exceed the molecular weight cut-off of general porins, for example, ferric

enterobactin receptor (FepA) (Buchanan, et al. , 1999; Krewulak, et al. , 2008).

There are two major non-specific porins in K. pneumoniae, which are known

as OmpK35 and OmpK36. These porins are homologs to OmpF and OmpC

of E. coli and S. enterica ser. Typhi with the percentage of amino acid identity

about 60 % and 80 % respectively (Balasubramaniam, et al. , 2012;

206 |

Hernández-Allés, et al. , 1999; Sugawara, et al., 2016). They have been

identified to play an important role in antibiotic resistance (Doménech-

Sánchez, et al. , 2003; Landman, et al. , 2009; Shakib, et al. , 2012), and

virulence (Chen, et al. , 2010) of K. pneumoniae (specifically for OmpK36). In

addition, OmpK35 has ~ 60 % amino acid homology to OmpK36 and ~ 58 %

amino acid identity to PhoE according to SWISS-MODEL prediction.

6.1.1 K. pneumoniae outer membrane porins (Omps)

K. pneumoniae expresses a variety of general porins, for example, OmpK35

and OmpK36. In addition, K. pneumoniae also might express several

additional porins. Beside these two major non-specific porins, some studies

have reported on another three porins, which are also important in

K. pneumoniae: OmpK26, OmpK34, and OmpK37 (Table 6-1).

A previous study by Doménech-Sánchez, et al. (1999) showed that expression

of K. pneumoniae OmpK37 porin, resulted in less susceptibility to β- lactam

antibiotics. However, more recent research by the other researchers,

demonstrated that OmpK37 played an insignificant role in antibiotic resistance

(Kaczmarek, et al. , 2006). Nevertheless, the functions of the other two porins,

OmpK26 and OmpK34, are currently not understood. As OmpK35 and

OmpK36 have been described to be involved in the penetration of antibiotics

into bacterial cells and virulence mechanisms of K. pneumoniae, we focused

on these two particular porins in this study.

207 |

Table 6-1 K. pneumoniae porins

Omp Description References

OmpK26

Oligogalacturonate-specific

porin

Confers resistance to

carbapenems when expressed

(García-Sureda, et al. , 2011)

OmpK34

Analogous to E. coli OmpA

Important in immune evasion in

vivo

(Hernández-Allés, et al. , 2000)

(Llobet, et al. , 2009)

OmpK35

Major nonspecific porin

Homologous to E. coli OmpF and

S. enterica ser. Typhi OmpF

(Chen, et al., 2010)

OmpK36

Major nonspecific porin

Homologous to E. coli OmpC and

S. enterica ser. Typhi OmpC

(Chen, et al., 2010)

OmpK37 Quiescent porin

Important for cellular functions in

OmpK35/36-deficient strains

(Doménech-Sánchez, et al.,

1999)

6.1.2 Major non-specific porins of K. pneumoniae

6.1.2.1 OmpK35

OmpK35 is a major trimeric porin (~40 kDa) which belongs to OmpF porin

group and possesses a large pore size compared to OmpK36 which belongs

to OmpC group (Pagès, et al., 2008). It has been reported that the diameter

of OmpF porin is 1.16 nm while the diameter of OmpC porin is 1.08 nm

(Seltmann, et al. , 2002). OmpK35 is upregulated in low-osmolarity

environments (Hernández-Allés, et al., 1999). The majority of the ESBL and

non-ESBL producing strains are suggested to express both OmpK35 and

OmpK36 (Hashemi, et al. , 2014). Many studies have also shown that loss of

208 |

OmpK35 and/or OmpK36 contribute to increased resistance of K. pneumoniae

to antibiotics specifically for ESBL-producing strains. For example, loss of

OmpK35 porin may result in antibiotic resistance in K. pneumoniae

(Kaczmarek, et al., 2006), since several studies have shown that absence of

OmpK35 is quite prominent in most ESBL-producing strains with increased

resistance to cephalosporins and carbapenems (Doménech-Sánchez, et al.,

2003; Jacoby, et al. , 2004). However, other studies have shown that OmpK35

and/or OmpK36 deficiency appeared to serve as a minor factor and does not

significantly alter the MICs of carbapenems to K. pneumoniae (Hernández-

Allés, et al., 2000; Jiang, et al. , 2009).

6.1.2.2 OmpK36

OmpK36 is a ~38 kDa trimeric porin which belongs to the OmpC porin group.

It has a smaller channel size in comparison to OmpK35, (Schembri, et al.,

2004), owing to the existence of two bulkier side chains (Trp80 and Tyr123)

(Albertí, et al. , 1995) at the pore constriction site (Dutzler, et al. , 1999).

OmpK36 is among the most abundant Omps in K. pneumoniae as well as

OmpA (Hernández-Allés, et al., 1999). This osmoporin is upregulated in high

osmolarity medium specifically in high salt or high sugar concentration

environments (Lugtenberg, et al. , 1983).

As suggested by Tsai, et al. (2011), OmpK36 might play a role in virulence of

K. pneumoniae. An ompK36 mutant of K. pneumoniae strain NVT2001 was

shown to have decreased virulence and bacterial survival, compared to wild

type strain and OmpK35 mutant in a murine model. This was possibly due to

the increase susceptibility of OmpK36 mutant to phagocytic clearance by

209 |

neutrophils (Tsai, et al., 2011). The reduced virulence has also speculated to

be due to the build up of internal toxic substances, and subsequently blocking

of the channels that are used for import of nutrients, even though, no

observable effect on OmpK36 mutant growth was noted (Chen, et al., 2010).

In addition, Omps are also suggested to be involved in host-pathogen

interactions (Alcántar-Curiel, et al., 2013; Galdiero, et al., 2008). As

demonstrated by Albertí, et al. (1995), OmpK36 might be involved in C1q

interaction since a binding site for C1q (Lys279 and Lys281 from β strand 13

and Asp282 of L7) (Dutzler, et al., 1999) was found in the crystal structure of

OmpK36. Further studies by this group also confirmed that C1q binds to

OmpK36 porin at its globular domain (Albertí, et al. , 1996) to activate the

complement classical pathway.

Tsai, et al. (2011) and Wang, et al. (2009) also showed that loss of both or one

of the porins may reduce susceptibility to cephalosporins and carbapenems

respectively. Synergistic mechanisms between loss of porins and production

of β-lactams is also suggested to contribute to the emergence of antibiotic

resistant K. pneumoniae strains. A study by Kaczmarek, et al. (2006) showed

that carbapenem-resistance is mediated by ESBL or AmpC β-lactamases in

combination with the loss of OmpK35 or/and OmpK36 or by carbapenemases

alone. Findings by Wang, et al. (2009) also showed that OmpK36 might play

a significant role in the reduced susceptibility to carbapenems in

K. pneumoniae producing AmpC and β-lactamases. However, there are no

in-depth studies regarding the specific functions of these particular porins in

K. pneumoniae adhesion and biofilm formation to abiotic surfaces.

210 |

6.1.2.3 Structure of non-specific porins

In general, the structure of non-specific porins, for example, OmpK36 consists

of three water-filled functional units making up a trimer. In each monomer,

there are 16 antiparallel strands which form a β-barrel to transverse the cell

membrane with short turns at the periplasmic side and large loops at the

outside of the cells (Schulz, 2004). The strands are tilted by 30 ° to 60 ° in

reference to the β-barrel axis and the diameter increases depending on the

degree of tilting. The third loop, L3, however, extends into the barrel (often

called ‘the eyelet’) and is not exposed at a cell surface, forming a constricted

region to determine the permeability properties towards certain substances

(Figure 6-1) (Danelon, et al., 2004; Nikaido, 2003). The specific porins are

thought to have a similar structure to non-specific porins, even though they

have 18-strands in their β-barrel structures, in contrast to the non-specific

porins which just have 16-strands (Schulz, 2004).

211 |

Figure 6-1 The structure of nonspecific porin.

The crystal structure is derived from OmpK36 of K. pneumoniae (PDB ID: 1OSM).

A) Top view of the trimeric structure. Loop 3 (known as ‘the eyelet’) which folds into

the barrel causes the pore restriction and narrowing the channel. L3 is coloured in

orange and as labelled. B) Side view of trimeric structure. The short turns and large

loops are as labelled (Dutzler, et al., 1999). The crystal structure was viewed using

NGL viewer (Rose, et al. , 2016; Rose, et al. , 2015).

212 |

6.1.3 Role of outer membrane porins (Omps) in biofilm

formation

Based on the results in Chapter 4, we found that proteins are likely to play an

important role in biofilm architecture and biofilm formation by K. pneumoniae.

We tested our hypothesis by investigating the role of type 1 fimbriae (T1F) in

biofilm formation by K. pneumoniae (Chapter 5), however, we could not really

establish a role for T1F in biofilm formation upon cultivation in both TSB and

AUM. We further hypothesised that another protein, outer membrane porins

(Omps), OmpK35 and OmpK36 could possibly have role in biofilm formation

by K. pneumoniae. This hypothesis was based on several previous

publications which have shown that ompC gene of E. coli, a homolog of

ompK36, has been found to be upregulated in cells growing within a biofilm of

E. coli (Prigent-Combaret, et al., 1999). In addition, OmpC also was

suggested to be involved in the first step of biofilm formation: initial adhesion

and also invasion of adherent-invasive E. coli (AIEC) to intestinal epithelial

cells, intestine-407 cells, under high osmolarity conditions (Rolhion, et al.,

2007).

Besides, the regulation of ompC is also shown to affect curli fimbriae

production (Prigent-Combaret, et al., 1999; Smith, et al. , 2017). Curli fimbriae

are essential surface components which are known to be involved in biofilm

formation by E. coli. The regulation of ompC and ompF is controlled under the

two-component EnvZ/OmpR system (detail is in section 6.14). This two-

component system is also implicated in the regulation of curli fimbriae

expression (Barnhart, et al. , 2006). Curli formation in E. coli are encoded by

two adjacent divergently transcribed csgBA and csgDEFG. csgB and csgA

213 |

encode for CsgB, the major structural subunit or curlin and CsgA, the nucleator

protein, respectively. While, csgD encodes for CsgD, a positive transcriptional

regulator for csgBA operon and also for csgEFG operon which encodes for

assembly components and transport factors (Barnhart, et al., 2006).

Increased phosphorylated outer membrane regulator (OmpR) is known to be

involved in increased transcriptional activation of ompC (Feng, et al. , 2003).

This subsequently affects the production of curli fimbriae since OmpR is known

to positively regulate curli fimbriae expression by activating csgD expression

(Prigent-Combaret, et al. , 2001). Since curli fimbrae assembly occurs at the

cell surfaces, genes involved in cell envelope and outer membrane biogenesis

are also thought to affect the production of curli fimbriae including ompC,

waaC, lpcA, and waaE (Smith, et al., 2017).

Although the correlation between regulation of ompC and curli fimbriae

expression in E. coli has been reported previously, there is still no evidence

regarding this kind of regulation in K. pneumoniae. Thus, it is important to

evaluate the contribution of Omps in biofilm formation by K. pneumoniae and

identify either the regulation of these porins would affect other surface proteins

(if any), such as fimbriae in K. pneumoniae that could eventually affect the

biofilm formation.

In addition, knowing the role of OmpC in the initial adhesion of E. coli on biotic

surfaces and also due to the fact that ompC is upregulated in E. coli biofilms,

we decided to explore the role of OmpK36 in biofilm formation by

K. pneumoniae on abiotic surfaces such as a silicone urinary catheter, as

214 |

almost 70-80 % of UTI (Nicolle, 2014) cases are attributed to the usage of

indwelling urinary catheters. In addition, it is clinically relevant since

K. pneumoniae is known as one of the most prevalent species causing

catheter-associated urinary tract infection (CAUTI) (Nicolle, 2014). Indwelling

devices also constitute a more attractive surface for bacterial adhesion and

biofilm formation since they do not have the protective mechanisms of mucosal

surfaces to bacterial infection, for example, exfoliation of bladder epithelial

cells (Ferrieres, et al. , 2007). Based on the previous observation, it is

suggested that exfoliation of epithelial cells and clearance of infected and

damaged bladder cells act as host defense mechanisms (Ferrieres, et al.,

2007; Mulvey, et al. , 2000).

Even though very few studies have looked for a relationship between Omps

and biofilm formation, there is still intriguing evidence from several studies,

which suggest that Omps might be involved in the formation of biofilms. OmpC

of S. enterica ser. Typhimurium was proposed to be involved in biofilm

formation of S. enterica ser. Typhimurium on cholesterol-coated surface

(Crawford, et al., 2010). In addion, a study by Prigent-Combaret, et al. (1999)

also suggested that ompC was upregulated in biofilm cells of E. coli. However,

to the best of our knowledge, there are no studies showing the relationship of

OmpK36, the homolog of these OmpC on biofilm formation by K. pneumoniae.

215 |

Table 6-2 Involvement of Omps in initial adhesion or biofilm formation on abiotic surfaces by Enterobacteriaceae.

Bacterial

species

Contribution to biofilm

formation

Description Model/system Media References

S. enterica ser.

Typhimurium

OmpC is suggested to

play an important role in

binding and biofilm

formation of S. enterica

ser. Typhimurium on

cholesterol-coated

surfaces in the presence

of bile.

Random transposons mutant

strains of S. enterica ser.

Typhimurium were screened

for the lack of an ability to form

biofilms by tube biofilm assay.

49 mutants were found to have

impaired adherence to and

biofilm formation on

cholesterol-coated tubes but

not on glass or plastic surfaces

and the transposon Tn10d was

identified in ompC and several

other genes in these mutants.

Cholesterol-coated

Eppendorf tubes or

24-well plastic

plates containing

Chromerge-

cleaned glass,

plastic, and

cholesterol-coated

coverslips

LB (with or

without 3%

crude ox bile)

(Crawford, et al.,

2010)

216 |

Bacterial

species

Contribution to biofilm

formation

Description Model/system Media References

E. coli K-12

ompC, the homolog of

ompK36 was found to be

upregulated in cells

growing within biofilm

communities.

Expression of four genes

known to be osmoregulated

was compared in attached

(biofilms) and free-living

(planktonic) bacteria in petri

dishes. ompC was among the

genes that were found to be

upregulated in biofilm cells

upon incubation in a high

osmolarity environment. This

was determined upon

observation on the kinetics of

osmoregulated lacZ fusion

expression by β-galactosidase

assay.

24-well microtitre

plates, petri dishes

or glass tubes

containing

Plexiglas strips

Minimal M63

(Prigent-Combaret,

et al., 1999)

217 |

6.1.4 Regulation of OmpK35 and OmpK36 expression

6.1.4.1 EnvZ-OmpR two-component system

The expression of OmpK35-OmpK36 (or OmpC-OmpF) is controlled by an

EnvZ-OmpR two-component system. This EnvZ-OmpR system is involved in

osmoregulation and responds to osmolarity changes in the environment

(Walthers, et al. , 2004). Expression of OmpK35 and OmpK36 is constitutive,

however, the expression levels of these porins individually is dependent on

the medium osmolarity. In low-osmolarity medium, OmpF (homolog of

OmpK35) is abundantly expressed, while in high-osmolarity medium,

expression of OmpF/OmpK35 is downregulated and expression of OmpC

(homolog of OmpK36) is upregulated (Feng, et al., 2003).

EnvZ is an osmosensor histidine kinase homodimer molecule, which

comprises of sensory histidine kinase domain (Feng, et al., 2003). This inner

membrane protein is comprised of short N-terminus tail in the cytoplasm, two

transmembrane regions, a periplasmic loop, and a large cytoplasmic domain

which contains histidine for EnvZ autophosphorylation (Wang, et al. , 2012).

OmpR is a cognate response regulator molecule; it consists of a receiver

domain and a DNA binding domains which binds to the OmpK35 or OmpK36

promoter and initiates the differential expression of these two porins

depending on the osmolality of the environment (Walthers, et al., 2004)

(Figure 6-2). The signaling cascade of this osmoregulation system can be

elucidated as i) the osmotic signal is detected by EnvZ and the sensor of this

histidine kinase is activated. Conformational changes in EnvZ are triggered

by variation of osmolarity in the environment, ii) EnvZ undergoes

218 |

autophosphorylation at histidine (His243) in the catalytic domain, iii) the

phosphate is transferred to OmpR on an aspartic acid (Asp50) residue and

designated as OmpR-P, iv) the OmpR-P binds to the ompF/ompK35 or

ompC/ompK36 promoters to regulate the expression of these two porins

depending on the osmolarity, v) ompF/ompK35 is expressed at low

osmolarity, while ompC/ompK36 is expressed and ompF/ompK35 is

suppressed under high osmolarity conditions and vi) the Omp-R is then

desphosphorylated by EnvZ which acts as a phosphatase.

Figure 6-2 Regulation of the porin genes by EnvZ/OmpR system.

EnvZ is an inner membrane sensor kinase. It is phosphorylated on a conserved His243

residue by cytoplasmic ATP. The phosphate group is then transferred to OmpR on a

conserved Asp50. OmpR-P complex subsequently binds to the promoter of porin

genes to regulate their expression. Figure is re-published with permission

(Feng, et al., 2003).

219 |

6.1.5 K. pneumoniae and urinary tract infections

Urinary tract infections (UTIs) are the most common bacterial infections that

can affect the upper urinary tract, which comprised of the kidneys

(pyelonephritis), and the lower urinary tract, which includes the ureter, bladder

(cystitis), and urethra (urethritis) (Amalaradjou, et al. , 2011). UTIs normally

start as a cystitis (bladder infection) but can spread by infecting the kidneys

(pyelonephritis) which can result in renal failure. The infection can spread to

the circulatory system if the pathogens which caused the UTIs cross the

tubular epithelial barrier in the kidneys (Flores-Mireles, et al. , 2015).

UTIs can be categorised into two classes: uncomplicated and complicated.

Uncomplicated UTIs normally affect healthy individuals who have no structural

or functional urinary tract abnormalities, are not pregnant and who have not

been fitted with indwelling devices such as urinary catheters (Foxman, 2010).

While complicated UTIs are defined as infections, which are associated with

several factors that may compromise the urinary tract or host defense such as

immunosuppression, renal failure, pregnancy, presence of foreign bodies such

as indwelling catheters and calculi. In addition, medical or surgical

comorbidities are also frequently associated with complicated UTIs (Flores-

Mireles, et al., 2015; Neal, 2008). UTIs can be caused by Gram-negative and

Gram-positive bacteria. The most prevalent species that causes UTIs is

uropathogenic E. coli (UPEC), and these are followed by K. pneumoniae,

E. faecalis, P. mirabilis, and S. saprophyticus (Flores-Mireles, et al., 2015).

220 |

6.1.5.1 Catheter-associated urinary tract infections

(CAUTIs)

Catheter-associated urinary tract infections (CAUTIs) are the most common

nosocomial infections prevalent in US hospitals and nursing homes. They are

associated with over than 1 million cases each year (Tambyah, et al. , 2000).

In England, 19.7 % of nosocomial infections are related with UTIs, and

between 43 % and 56 % of these cases are reported to be associated with

urinary catheter (Hopkins, et al. , 2012; Loveday, et al. , 2014; Smyth, et al. ,

2008). Urinary tract catheters are prescribed as soon as an individual has

impaired bladder function and reduced capacity to permit urinary drainage.

Other than this reason, there are several factors that contribute to prescription

of urinary catheters, for example, urinary retention, urinary incontinence,

bladder irrigation, and post-operative patients (Feneley, et al. , 2015). The

usage of indwelling urinary catheters is classified as short term if the catheters

are in situ for less than a month and long term when in situ for more than 30

days (Tambyah, et al., 2000).

6.1.5.2 Biofilms on catheters

Biofilm formation in catheters is a serious problem related to nosocomial

infections and it is noted that eradication of biofilms is very difficult.

K. pneumoniae is established as one of the important causes of CAUTIs right

after E. coli (Wilson, et al. , 2004). This is partly because of the capacity of

K. pneumoniae to produce urease (Nicolle, 2014), which catalyses the

hydrolysis of urea to carbon dioxide and ammonia. The production of urease

may facilitate the formation of a crystalline biofilm and this subsequently will

cause biofilm-associated infections. Biofilms in a catheter normally form in

221 |

the eyehole, in the catheter itself (Stickler, et al. , 2003), or in the drainage

bag (Nicolle, 2014).

6.1.6 Aims of this chapter

In our study, we generated ompK35 and ompK36 mutants in two different

strains of K. pneumoniae: Kpneu_1, a clinical strain isolated from blood and

C3091, a laboratory strain, which was originally isolated from UTI patient. The

molecular techniques used for mutant construction of ompK35 and ompK36

will be discussed in this chapter. The phenotypic assays comparing the MIC

of antibiotics and antiseptic for Kpneu_1 and C3091 parental strains and their

isogenic mutant strains will also be presented. Subsequently, the capacity of

these mutant strains to form biofilm on abiotic surfaces will be described.

222 |

6.2 Materials and methods

6.2.1 Bacterial strains and plasmid

E. coli -select (silver efficiency) was used for cloning of pGEM-T Easy

(Promega, UK) mutant constructs harbouring mutant cassettes (containing

flanking regions of the gene of interest with erm1(B) or aad9 selective marker

genes inserted in between of the flanking regions) or pUC19-tet(M)

harbouring putative complementation cassettes (containing ompK35 or

ompK36 with their putative promoters). Six strains: Kpneu_1, Kpneu_63,

Kpneu_64, Kpneu_65, Kpneu_110, and C3091 were used for ompK35 and

ompK36 screening. The mutant cassettes were constructed using the

ompK35 sequences from Kpneu_1 and ompK36 sequences from Kpneu_63,

Kpneu_65, and C3091. Strain Kpneu_1 and C3091 were used as the hosts

for development of ompK35 or ompK36 mutant strains. All strains were grown

on LB agar at 37 C for 16 h or with shaking at 200 rpm if cultivated in broth

with appropriate concentrations of antibiotic(s). All the constructs and strains

used in this study are summarised in Table 2-1 and Table 2-6 in Chapter 2.

6.2.2 Determination of genetic context of ompK35 and

ompK36 of K. pneumoniae

The genetic context of ompK35 and ompK36 was determined by first,

amplifying ~500 bp upstream and downstream regions of ompK35 and

ompK36 of K. pneumoniae. The upstream and downstream regions of

ompK35 and ompK36 were amplified using 35_KO1_2 and 35_KO4_3 or

36_KO1 and 36_KO4_4 primer sets, respectively. The PCR product was then

purified and sequenced. Multiple sequence alignment of upstream and

223 |

downstream regions of ompK35 and ompK36 were then conducted using

BioEdit and GeneDoc. The promoter and terminator sites for ompK35 and

ompK36, and also for the genes located upstream and downstream of ompK35

and ompK36 were predicted using BPROM (Softberry, Inc., NY) and ARNold

Finding Terminators (Naville, et al., 2011). The promoter site of rscD was

predicted based on the sequence obtained from Genbank (Accession no:

CP006648.1) since the amplified upstream region of ompK36 derived from the

strains used in this study could not cover the region with predicted promoter

site for rcsD.

6.2.3 Analysis of protein structure

We previously hypothesised that OmpK36 could possibly have an effect on

biofilm formation by K. pneumoniae since its homolog, OmpC was established

to be involved in the first step of biofilm formation, initial adhesion of E. coli on

epithelial cells. Before we further investigated the role of OmpK36 in biofilm

formation by K. pneumoniae, we firstly compared the protein structure of

OmpC (as a template) and OmpK36 (as a target). In addition, we also

compared the protein structure of OmpK36 (as a template) with predicted

structure of OmpK35 (as a target). Comparative protein structural analysis

was conducted by comparing the amino acid sequences and structural

alignment of the templates with the target sequence or structure. The

purposes of this protein structural analysis were; i) to align three-dimensional

(3D) structures of OmpC and OmpK36 and to obtain an idea regarding their

structural similarity and ii) to predict OmpK35 protein structure using homology

modelling using OmpK36 as a template. The dataset used for the templates

224 |

were obtained from Protein Data Bank (PDB). It included the amino acid

sequences and 3D structures of OmpC (PDB ID: 2J1N) and OmpK36 (PDB

ID: 1OSM) of E. coli and K. pneumoniae respectively. The 3D structures were

then generated using USCF Chimera 1.12rc (Pettersen, et al., 2004)

software.

The deduced amino acid sequence of OmpK35 derived from Kpneu_1 was

obtained by translating the nucleotide sequence using EMBOSS-Transeq

(Rice, et al., 2000). The in silico 3D structure prediction of OmpK35 was

conducted using template-based homology modelling. It was carried out by

searching the PDB for known protein structures using OmpK35 amino acid

sequences as the query, and the structure was then predicted using SWISS-

MODEL (Biasini, et al., 2014). The template structure was chosen based on

the highest sequence similarity to the query sequence. OmpK36 (PDB ID:

1OSM) was selected as a template for OmpK35 structure prediction and the

prediction was conducted by SWISS-MODEL (Biasini, et al., 2014). The

predicted model was illustrated using Chimera 1.12rc (Pettersen, et al.,

2004). The target-template alignment for comparative modelling was

established by superimposing the 3D structural alignment or superimposition

of OmpC and OmpK36, and also between OmpK36 and OmpK35. The

percentage of amino acid sequence identity and root-mean-square distance

(RMSD) between both structures were recorded. RMSD is the average

distance of atoms pairs in two macromolecules. This RMSD value is

calculated to determine the structural similarity between two 3D protein

structures, the reference protein structure and the superimpose protein

structure (Carugo, et al. , 2001). Two identical protein structures will have

225 |

RMSD value of 0 Å and the higher the RMSD the more dissimilar the protein

structures.

6.2.4 Development of ompK35 and ompK36 mutant strains

6.2.4.1 Screening of ompK35 and ompK36

All of the six strains were screened for ompK35 and ompK36 prior to mutant

development. The PCR was conducted using the following method mentioned

in Chapter 2 and list of primers are listed in Table 2-7 in Chapter 2.

6.2.4.2 Construction of mutant constructs using SOE PCR

Mutant constructs of ompK35 or ompK36 were generated by conducting

splicing by overlap and extension (SOE) PCR that amplifying a region

upstream (~500 bp) (5’) and a region downstream (~500 bp) (3’) from a

targeted gene, with an erythromycin resistance gene (erm1(B)). The erm1(B)

gene was chosen as a selective marker since none of the six K. pneumoniae

strains harboured the erm1(B) gene (data not shown). Two DNA fragments

(~500 bp for each upstream or downstream region) that flanked the regions

to be deleted (ompK35 or ompK36) were amplified using specific primers,

listed in Table 2-7 (in Chapter 2, section 2.4.2), from genomic DNA of

Kpneu_1, Kpneu_63, Kpneu_65, and C3091 strains. For ompK35 mutant

construction, the primer sets of 35_KO1 and 35_KO2 (to amplify Fragment

1_35 – the upstream region of ompK35) and 35_KO3 and 35_KO4 (to amplify

Fragment 2_35 – the downstream region of ompK35) were used. While the

flanking regions of ompK36 were amplified using primer sets of 36_KO1 and

36_KO2 (to amplify Fragment 1_36 – the upstream region of ompK36) and

36_KO3 and 36_KO4 for Kpneu_1, Kpneu_63, and Kpneu_65 or combination

226 |

between 36_KO3 and 36_KO4_2 for Kpneu_64, Kpneu_110, and C3091) (to

amplify Fragment 2_36 – the downstream region of ompK36) primer sets.

The upstream 40 bp primers of 35_KO2 and 36_KO2 (the reverse primers to

amplify Fragment_1) contained 20 bp with complementary to the upstream

region of the gene to be replaced and 20 bp of nucleotide sequence that was

homology to the first 20 bp of erm1(B) including the ribosome binding site

(RBS). Whilst, the downstream 40 bp primers of 35_KO3 and 36_KO3 (the

forward primers to amplify Fragment 2) contained 20 bp of nucleotide

sequence that complementary to the last 20 bp of erm1(B) gene and 20 bp

homology to the downstream region of the gene to be replaced. The erm1(B)

gene cassette was amplified from C. difficile 630 genomic DNA using 35_KO5

and 35_KO6 (to amplify the Fragment 3 with the 20 bp complementary to the

upstream and homology to downstream regions of ompK35) or 36_KO5 and

36_KO6 (to amplify the Fragment 3 with the 20 bp complementary to the

upstream and homology to the downstream regions of ompK36) primer sets.

The 40 bp primers of 35_KO5 and 35_KO6, and 36_KO5 and 36_KO6 primer

sets contained first 20 bp complementary and last 20 bp homology of the

erm1(B) gene and extension of 20 bp with complementary to the upstream

and homology to downstream region of ompK35 or ompK36.

These three purified PCR products (Fragment 1, Fragment 2, and Fragment

3) obtained from individual PCR were mixed at equimolar concentrations. This

was followed by an overlap and extension PCR using the 35_KO1_3 and

35_KO4_2 or the 36_KO1_2 and 36_KO4_3 primer sets to generate a mutant

cassette consisting of erm1(B), flanked by ~500 bp of the upstream and

downstream regions of the target gene to be replaced. The simplified method

227 |

is shown in Figure 6-3. The resulted ~1.8 kbp PCR product was purified by

gel extraction and ligated into the pGEMT-Easy vector (Promega, UK).

228 |

Figure 6-3 Schematic diagram of development of mutant constructs using

ompK35 as an example.

The blue rectangles represent upstream and downstream regions and red rectangle

or red arrow indicate the antibiotic selective marker. Primer position and direction are

as labelled.

229 |

Transformation by heat shock was conducted as described in Chapter 2,

section 2.4.11. The cells were subsequently plated onto LB agar plates

containing 100 µg/ml ampicillin (amp), 200 µg/ml erythromycin (erm),

50 µg/ml X-gal and 100 µM IPTG, and incubated at 37 °C for 48 h. The white

colonies were selected and grown at 37 °C for 16 h in LB broth containing

100 µg/ml amp and 200 µg/ml erm under 200 rpm shaking conditions. The

plasmid was then extracted and digested with EcoRI (NEB, UK). The digested

product was separated on a 1.0 % agarose gel by electrophoresis to

determine the size of the constructs and plasmids with the correct insert size

were sent for sequencing. A plasmid with correct insert and sequence was

designated as pGEMT_erm_∆ompK35 and pGEMT_erm_∆ompK36. The

plasmid was then used as a template for PCR to amplify the mutant cassettes

using high fidelity (HF) polymerase, Q5 polymerase (Chapter 2 section 2.4.3).

The amplified HF-PCR product was then purified by gel extraction and the

DNA was concentrated (SPD1010 Integrated SpeedVac, Thermo Scientific,

UK) prior to electroporation into electrocompetent K. pneumoniae cells.

Uncountable colonies were observed after incubation at 37 C for 16 h on

LB-erm 400 µg/ml. The colonies were then selected and screened for allelic

exchange using: 1) 35_KO1_2 and 35_KO4_3 primers for confirmation of

ompK35 deletion and 36_KO1 and 36_KO4_4 primers for confirmation of

ompK36 deletion, 2) specific target genes primer, OmpK35_scr_F and

OmpK35_scr_R, or OmpK36_scr_F and OmpK36_scr_R for ompK35 and

ompK36 respectively, and 3) selectable marker gene primer, ermBF and

ermBR to amplify erm1(B) region. However, no mutants were successfully

obtained. Suspecting that erm1(B) was not a suitable marker to be used for

230 |

mutant construct development, the mutant constructs were then re-developed

by replacing erm1(B) with a different selective marker, spectinomycin

resistance gene (aad9).

6.2.4.3 Construction of new mutant cassettes using aad9

as a selective marker

Since the development of mutant strains using erm1(B) as a selective marker

was unsuccessful, construction of new mutant cassettes was carried out by

replacing erm1(B) with add9 (Figure 6-4). aad9 was amplified using primer

Spect_HindIII_GC_F and Spect_HindIII_GC_R primer set and plasmid

pLR16T as a template. The pGEMT backbone together with upstream and

downstream regions of ompK35 or ompK36 was amplified using 35D_BB

_HindIII_GC_F and 35U_BB_HindIII_GC_R primer set for plasmid backbone

including ompK35 flanking regions and 36D_BB_HindIII_GC_F and

36U_BB_HindIII_GC_R primer set for plasmid backbone including ompK36

flanking regions, and pGEMT_erm_∆ompK35 and pGEMT_erm_∆ompK36

as template. The amplified PCR products, including the plasmid backbone

and aad9 were purified and digested using HindIII at 37 °C for 1 h. The

plasmid backbone was dephosphorylated and further purified together with

the digested aad9 fragment. Purified insert (aad9) and plasmid backbone

were ligated using T4 DNA ligase at 16 C for 16 h. The ligated product was

transformed into E. coli -select competent cells using the heat shock

method. Transformants were selected on LB agar plates containing

100 µg/ml amp, 100 µg/ml spectinomycin (spc), 50 µg/ml X-gal and 100 µM

IPTG. After 16 h incubation at 37 C, colonies were picked and grown in LB

broth containing 100 µg/ml amp and 100 µg/ml spc and incubated at 37 °C

231 |

for 16 h under 200 rpm shaking condition for plasmid extraction. The

extracted plasmids were digested with HindIII to confirm the correct size of

the insert. The plasmids were also used as a template for colony PCR

screening using insert primer, Spect_HindIII_F_GC and Spect_HindIII_R_GC

and mutant cassette primer sets, 35_KO1_3 and 35_KO4_2 or 36_KO1_2

and 36_KO4_3. The amplicons were separated on a 1.0 % agarose gel by

electrophoresis. The plasmids with the correct insert size were sent for

sequencing to confirm the correct insert. Plasmids with the correct insert and

sequences were designated as pGEMT_spc_∆ompK35 and

pGEMT_spc_∆ompK36. The plasmids obtained from this experiment were

listed in Table 2-6 in Chapter 2.

232 |

Figure 6-4 Schematic diagram of erm1(B) replacement with aad9 in ompK35

cassette as an example.

The replacement of erm1(B) with aad9 was conducted by amplifying plasmid

backbone (excluding the erm1(B)) by inverse PCR and the add9 by PCR. The

plasmid backbone and amplified aad9 product were then digested by HindIII. The

plasmid backbone was dephosporylated prior to ligation. The digested products

were purified and ligated with T4 DNA ligase. Plasmid backbone primer and insert

primer are denoted by green arrow.

233 |

6.2.4.4 Modification of the pKD46 helper plasmid by

replacing bla with aac(3)-IV

An attempt was taken to use a recombineering (homologous recombination-

mediated genetic engineering) method based on the Red recombinase

system (Wei, et al. , 2012) since previous attemps without pKD46 as a helper

plasmid did not result in any K. pneumoniae mutant strains. Temperature

sensitive pKD46 was used in this recombineering approach (provided

generously by Dr Sophie Halen, Imperial College) (Datsenko, et al., 2000).

This plasmid harbours exo, beta, and gam, which encode for Exo, Bet and

Gam proteins. Exo (red α), is an exonuclease (5’-3’) that digests the double-

stranded DNA and generates 3’ single-stranded DNA tails. While Bet (red β)

is a single strand DNA annealing protein that binds to the complementary

ssDNA strand, and Gam (red γ) protein protects linear extraneous DNA from

being degraded by host nucleases, RecBCD (Wei, et al. , 2013; Wei, et al.,

2012). The ampicillin resistance marker in pKD46, bla was replaced with

aac(3)-IV, apramycin resistance gene (the same approach as mentioned in

Figure 6-4) since all of the examined K. pneumoniae strains were resistant to

ampicillin. The aac(3)-IV gene was amplified using apra_HindIII_GC_F and

apra_HindIII_GC_R primer set and plasmid pSET152 as a template. Whilst,

the backbone of plasmid pKD46 (without the bla) was amplified using primer

pKD46_HindIII_GC_F and pKD46_HindIII_GC_R primer set and plasmid

pKD46 as a template. The amplified PCR products (pKD46 backbone and

aac(3)-IV insert) were purified, digested using HindIII, ligated (see Figure 6-

5) and transformed into E. coli -select competent cells as previously

mentioned. LB agar plates containing 50 µg/ml apramycin (apra) were used

234 |

as selective media. After incubation at 30 C for 24 h, colonies were picked

and grown in LB broth containing 100 µg/ml amp and 50 µg/ml apra, and

incubated at 30 °C for 16 h for plasmid extraction. Since pKD46 has a

temperature sensitive origin of replication, oriR101/repA101ts, cells

harbouring the plasmid must be grown at 30 C. The extracted plasmid was

digested with HindIII to confirm the insert. The plasmid was also used as a

template for PCR screening using apra_HindIII_GC_F and

apra_HindIII_GC_R primer set. The purified insert from the putative plasmids

were then sequenced and the confirmed plasmid was designated as pKD46-

apra. The plasmid obtained from this experiment was listed in Table 2-6 in

Chapter 2.

235 |

Figure 6-5 Schematic diagram of the replacement of bla with aac(3)-IV in

pKD46.

The replacement of bla with aac(3)-IV was conducted by amplifying the pKD46

plasmid backbone (excluding the bla) by inverse PCR and the aac(3)-IV by PCR.

The plasmid backbone and amplified aac(3)-IV product were then digested by

HindIII. The plasmid backbone was dephosporylated prior to ligation. The digested

products were purified and ligated with T4 DNA ligase. Plasmid backbone primer

and insert primer are denoted by purple arrow.

236 |

6.2.4.5 Preparation of electrocompetent K. pneumoniae

cells

A single colony of K. pneumoniae was grown in 5 ml of LB broth at 37 C for

16 h with 200 rpm shaking prior to preparation of electrocompetent cells. The

100 µl of 16 h culture was then used as a starter culture and inoculated in

10 ml LB broth (with the starting OD600 of 0.05) and incubated at 37 C with

200 rpm shaking until OD600 of 1.5. The mid-log phase culture was

centrifuged at 11952 x g for 30 min at 4 C (Sorvall RC5B plus, SS-34 Sorvall

rotor). The supernatant was decanted and the pellet was re-suspended with

1 ml of ice-cold molecular grade water. The suspension was then centrifuged

at 17950 x g for 15 min at 4 °C (Eppendorf 5417R). The supernatant was

discarded and the pellet was re-suspended in 1 ml of ice-cold molecular grade

water. This washing step was repeated three times. All the procedures were

carried out on ice. The electrocompetent cells were re-suspended in 1 ml

molecular grade water with 20 % glycerol and kept at -80 C until needed.

237 |

6.2.4.6 Electroporation of pKD46-apra into

electrocompetent K. pneumoniae cells

The electroporation was carried out using the method by Fournet-Fayard, et

al. (1995) with some modifications. The electrocompetent cells were thawed

on ice together with pKD46-apra plasmid DNA. Five microlitres of 200 ng/µl

plasmid DNA was added to 50 µl of thawed electrocompetent cells. The

mixture was mixed by flicking the bottom of the tube. The mixture was

incubated on ice for 5 min and transferred into a pre-chilled 1 mm

electrocuvette (Bio-Rad, US) using a pipette. The plasmid was electroporated

into the electrocompetent cells using a Gene Pulser II (Bio-Rad, UK)

apparatus set at 1.8 kV, 50 µF capacitance, 200 resistance. Following the

pulse, 950 µl of sterile BHI was added to the mixture and incubated at 30 C

for 2 h with 200 rpm shaking. After incubation, 200 µl of the culture was plated

onto LB agar supplemented with apra 50 µg/ml and incubated at 30 C for

24 h. Transformants obtained from the electroporation were then screened

for the aac(3)-IV insert and pKD46 plasmid backbone by PCR using

apra_HindIII_GC_F and apra_HindIII_GC_R, and pKD46_HindIII_GC_F and

pKD46_HindIII_GC_R primer set. The purified PCR product was

subsequently sent for sequencing. The confirmed strains were designated as

Kpneu_1/pKD46-apra or C3091/pKD46-apra. The strains obtained from this

experiment were listed in Table 2-1 in Chapter 2.

238 |

6.2.4.7 Preparation of electrocompetent K. pneumoniae

harbouring pKD46-apra cells

The preparation of electrocompetent Kpneu_1/pKD46-apra and

C3091/pKD46-apra cells was conducted according to the previous method

mentioned in 6.2.4.6 with some modifications. Single colony of

Kpneu_1/pKD46-apra or C3091/pKD46-apra was inoculated into 5 ml of LB

broth with 50 µg/ml of apramycin and incubated at 30 °C for 16 h with

200 rpm shaking. One hundred microlitres of 16 h culture was used as a

starter culture (with an OD600 of 0.05) and inoculated in 10 ml LB broth with

50 µg/ml of apramycin. The culture was grown until it reached OD600 of 0.1

and 100 mM of L-arabinose (Sigma) was added (to induce the PBAD promoter

of the araBAD (arabinose) operon) into the culture and further incubated at

30 °C until it reached OD600 of 0.4. The culture was chilled on ice for 5 min

and centrifuged at 4500 x g for 15 min at 4° C (Eppendorf 5804R). The

supernatant was discarded and the pellet was washed with 2 ml of ice-cold

molecular grade water. The suspension was aliquoted into microcentrifuge

tubes (1 ml for each tube) and centrifuge at 17950 x g for 15 min at 4° C

(Eppendorf 5417R). The supernatant was discarded and the pellet was

washed with 1 ml of ice-cold molecular grade water. This washing step was

repeated for three times. All the procedures were carried out on ice. The

electrocompetent cells were subsequently re-suspended in 1 ml molecular

grade water with 20 % glycerol and kept at -80 C until needed.

239 |

6.2.4.8 Electroporation of ompK35 and ompK36 mutant

cassettes into electrocompetent K. pneumoniae harbouring

pKD46-apra cells

The amplification of mutant cassettes was conducted by high-fidelity PCR

(HF-PCR) using pGEMT_spc_∆ompK35 or pGEMT_spc_∆ompK36 as a

template. The PCR product was separated on a 1 % agarose gel by

electrophoresis and the expected bands were excised and purified. Five

microlitres of the gel purified mutant cassettes DNA (100 ng/µl) was

subsequently electroporated into 50 µl of thawed electrocompetent

Kpneu_1/pKD46-apra or C3091/pKD46-apra cells as previously mentioned

with some modifications. The outgrowth incubation was conducted at 37 °C

for 1 h and 30 min instead of 30 °C for 2 h to cure the thermosensitive pKD46-

apra plasmid. The culture was plated on LB-spc 100 µg/ml and incubated at

37 C for 16 h. Several colonies were obtained upon incubation. The

occurrence of allelic exchange was confirmed by PCR using the 1) 35_KO1_2

and 35_KO4_3 primer for confirmation of ompK35 deletion and 36_KO1 and

36_KO4_4 primer for confirmation of ompK36 deletion, 2) specific target

genes primer, OmpK35_scr_R and OmpK35_scr_F or OmpK36_scr_F and

OmpK36_scr_R for ompK35 and ompK36 respectively, and 3) selectable

marker gene primer, Spect_HindIII_GC_F and Spect_HindIII_GC_R to

amplify the aad9 instead of the primers to amplify erm1(B). The purified PCR

product was then sent for sequencing for validation. The sequencing results

of the PCR products of the mutant strains were analysed by multiple

sequence alignments using BioEdit and GeneDoc.

240 |

Figure 6-6 Flow diagram of the gene replacement by double cross over

homologous recombination in K. pneumoniae.

The blue rectangles represent the flanking regions of ompK35 or ompK36. The

dashed line represents the chromosome of K. pneumoniae. Primer position and

direction are denoted by purple arrows.

241 |

6.2.5 Complementation of ompK35 and ompK36 strains

6.2.5.1 Preparation of ompK35 and ompK36

complementation constructs

pUC19-tet(M) (courtesy from Dr Supathep Tansirichaiya) was used to

develop complementation constructs. The ompK35 and ompK36 genes with

their predicted promoters were amplified using genomic DNA from Kpneu_1

and C3091 wild type strains as a template and acted as an insert. The purified

insert and pUC19-tet(M) plasmid were digested with SphI at 37 °C for 1 h.

Digested pUC19-tet(M) was dephosphorylated and purified together with the

digested insert. Both of the purified digested products were then ligated using

T4 DNA ligase. The ligated product was transformed into E. coli α-select

competent cells and plated on LB agar plates containing 15 µg/ml tetracycline

(tet), 50 µg/ml X-gal and 100 µM IPTG. The putative transformants were

screened using blue-white colony screening. The white colonies were grown

in LB broth containing 15 µg/ml tet at 37 °C for 16 h and the plasmid was then

extracted. The extracted plasmid was digested with SphI to determine the

insert size. The plasmid with correct insert size was then sent for sequencing

using M13F and M13R universal primers.

242 |

Figure 6-7 Construction of pUC19-tet(M)_ompK35 complementation construct.

The ompK35 was amplified and digested with restriction enzyme SphI. The digested

fragment was ligated to digested-dephosporylated pUC19-tet(M) vector backbone

prior to transformation into E. coli α-select competent cells. The ompK36

complementation construct was developed using the same method. The SphI sites

are boxed in red.

243 |

6.2.6 Epsilometer test (E-test) method

The wild type and their isogenic mutant bacterial stocks were streaked onto

LB or their selective medium and incubated at 37 °C for 16 h. A single colony

was grown in 5 ml MH broth at 37 °C for 16 h on a 200 rpm rotary shaker.

One hundred microlitres of a 16 h culture was adjusted to OD600 of 0.05 in

fresh MH medium and grown until it reached mid-log phase (data not shown).

The mid-log phase culture was diluted in saline solution (0.85 % NaCl) to

achieve a suspension turbidity equivalent to 0.5 McFarland Standard (OD600

of 0.08-0.1). The suspension was swabbed evenly on MH agar plates using

a cotton swab and allowed to dry for 10 min before applying an E-test strip.

E-test strips (gentamicin, ciprofloxacin, cefotaxime or tetracycline

(Biomerieux, UK)) were then placed carefully on agar plates with the MIC

scale facing upward with the maximum concentration nearest to the plate rim

using sterile forceps and incubated at 37 °C for 16 h. The zone of inhibition

(ellipse) was then observed and recorded.

6.2.7 Determination of MIC using agar and broth dilution

methods

Determination of MIC using agar and broth dilution method was conducted

following method by Wiegand, et al. (2008). Agar dilution method was used

for MIC determination of spectinomycin and apramycin for mutagenesis

purposes. Mid-log phase bacterial suspensions were adjusted to OD600 of

0.01. One microlitre of the diluted suspension was dropped on the MH agar

with different concentrations (12.5 µg/ml – 200 µg/ml for spectinomycin and

2 µg/ml – 256 µg/ml for apramycin) of antibiotics and incubated at 37 °C for

244 |

16 h. The MIC was then recorded based on the lowest concentration of

antibiotics that completely inhibited the growth of bacteria on plates.

The broth dilution method was used for MIC determination of cetylpyridinium

chloride (CPC) and cetyltrimethylammonium bromide (CTAB). Mid-log phase

bacterial suspensions were adjusted to OD600 of 0.01. Hundred microlitres

of MH broth with different dilutions of antiseptics (obtained by serial dilution

of the highest concentration of antiseptics) was added into wells of a 96-well

flat-bottom polystyrene non-treated microtitre plate (Falcon, BD), and

inoculated with 100 µl of diluted suspension in each well. The microtitre plate

was incubated at 37 °C for 16 h with gentle agitation (100 rpm) on a rotary

shaker (IKA ® Vibrax-VXR, Germany). The plate was observed for any

cloudiness and turbidity after incubation. The MIC of the antiseptics was

determined as the lowest concentration that inhibited visible growth.

6.2.8 Growth kinetics of wild type and mutant strains in TSB

and AUM

The growth curves were performed using the method described in Chapter 4,

section 4.2.3.

6.2.9 Biofilm formation assays in microtitre plate

The biofilm assays were performed using the method described in Chapter 3,

section 3.2.1.

245 |

6.2.10 Biofilm formation assays on catheter pieces

The capacity of K. pneumoniae to form biofilms on catheter pieces was

determined using strain Kpneu_1 and C3091, and their isogenic mutants. The

biofilm formation on catheter pieces for Kpneu_1 and its isogenic mutants was

conducted in TSB since our data in Chapter 4 showing that Kpneu_1 formed

more biofilm in TSB and had less capacity to form biofilms in AUM. In addition,

the fimS orientation assays in biofilm cells cultivated in TSB mirrored the

findings obtained by Stahlhut, et al. (2012). While, biofilm formation by C3091

and its isogenic mutants was conducted in AUM because this strain was

observed to form more biofilm upon cultivation in AUM compared to TSB (see

Figure 4-10).

Biofilm formation by strain Kpneu_1, C3091 and their isogenic mutants was

tested on sterile all-silicone Foley catheter pieces (Bard, UK, 165814UK) and

incubated under gentle agitation. Pieces of catheter, 5 cm in length were cut

and slit in half aseptically to facilitate accessibility to bacterial cells, and they

were pre-equilibrated in 15 ml of TSB or AUM at 37 °C for half an hour. A 1:10

dilution of mid-log phase bacterial culture (1.5 ml) was inoculated into the TSB

or AUM containing the catheter piece and the culture was incubated

horizontally at 37 °C for 24 h under gentle agitation at 60 rpm (R100, Luckham,

UK). The catheter pieces without any bacteria were labelled as blank group.

The experiment was conducted in triplicate for three independent biological

repeats.

246 |

Figure 6-8 Biofilm formation assay on catheter pieces

247 |

6.2.10.1 Staining and enumeration of biofilm cells on

catheter pieces

Cells attached to catheter pieces were stained using CV staining and the

viability of adhered bacteria was determined by serial dilution, plating and

counting viable bacteria. After the incubation, the catheter pieces were cut

into 3 cm section for CV staining and 1 cm fragment for viable-cell counting.

Three centimetres catheter pieces were washed in a new tube with 15 ml of

PBS (the tubes were inverted three times carefully to avoid disturbance to

biofilms). The PBS was discarded and the catheter pieces were subsequently

dried for 16 h before staining. The dried catheter pieces were transferred into

a new Bijou tube and stained with 5 ml of 0.1 % CV for 30 min (the tubes were

placed horizontally to ensure even distribution of CV). The stained catheter

pieces were transferred into new Bijou tubes and washed twice with 5 ml of

PBS. The washed catheter pieces were then placed on absorbent paper to

reduce any excessive CV residues, and placed in new Bijou tubes filled with

1 ml of 90 % ethanol. The catheter pieces were destained for 15 min

horizontally and vortexed vigorously to completely solubilize the CV. The CV

was measured spectrophotometrically at 590 nm.

The enumeration of biofilm cells was conducted by washing the catheter

pieces carefully for three times in 1 ml saline (0.85 % NaCl) to completely

remove the non-adherent cells. The washed catheter pieces were transferred

into a new microcentrifuge tube containing 1 ml saline (0.85 % NaCl) and the

adhered bacteria were detached by sonication using water-bath sonicator

(FB15046, Thermo Scientific) for 10 min and vortexed vigorously for additional

248 |

2 min. The suspension was serial diluted and plated on TSA. Based on the

colony counts, the CFU/cm2 were then calculated.

249 |

6.3 Results

6.3.1 Determination of genetic context of ompK35 and

ompK36

The genetic context of the ompK35 and ompK36 in K. pneumoniae strains

used in this study is shown in Figure 6-9. The ompK35 is flanked by

asparaginyl-tRNA synthetase (asnS) and tyrosine aminotransferase (tyrB)

genes. While the ompK36 is flanked by a gene coding for a putative phosphate

transfer intermediate protein (rcsD) and a gene coding for a thiamine

biosynthesis lipoprotein (apbE). The details of the upstream and downstream

regions of ompK35 and ompK36 is shown in Figure 6-9 respectively.

Figure 6-9 Genetic context of ompK35 and ompK36.

The figure shows the genetic context of the upstream and downstream regions of

A) ompK35 and B) ompK36. The predicted promoters are denoted by the arrows and

the predicted terminator is indicated by the hairpin loop structure. Promoters were

predicted by BPROM programme (Softberry, Inc., NY) and terminator site was

predicted by ARNold Finding Terminators software (Naville, et al., 2011).

250 |

Figure 6-10 and Figure 6-12 show the multiple sequence alignment of

upstream and downstream regions of ompK35 and ompK36 for all strains used

in this study. The sequences were obtained from the amplification of the

upstream and downstream regions of both of the genes using 35_KO1_2 and

35_KO4_3 or 36_KO1 and 36_KO4_4 primer sets. The DNA sequences of

the amplified product, however, could not cover the intergenic region with a

putative promoter site of rcsD in OmpK36. Thus, the putative promoter sites

of rcsD was predicted using the genome sequence of Klebsiella pneumoniae

CG43 (Accession no: CP006648.1) available in the GenBank (Figure 6-11).

251 |

Figure 6-10A) Multiple sequence alignment of upstream region of ompK35.

252 |

Figure 6-10B) Multiple sequence alignments of downstream region of ompK35.

253 |

Figure 6-10 Multiple sequence alignment of upstream and downstream regions of ompK35.

The figures show the multiple sequence alignments of A) upstream and B) downstream regions of ompK35. The start codon of ompK35 is

labelled with ATG and red arrow while the stop codon of ompK35 is labelled with TAA and asterisk. The putative start codon for tyrB is signposted

by thick red arrow. The predicted -35 and -10 sequences of putative promoters for A) ompK35 and B) tyrB are denoted by red boxes. The

putative terminator site is indicated by yellow line. The primers sites used for the development of mutant constructs are as labelled and the

direction of the primers are showed by red arrows. The nucleotide changes are indicated by blue box. The deduced amino acid sequences

encoded by tyrB are written in blue. The promoters sites were predicted by BPROM (Softberry, Inc., NY) and terminator site was predicted by

ARNold Finding Terminators software (Naville, et al., 2011). The multiple sequence alignments were conducted using GeneDoc.

254 |

Figure 6-11 The upstream region of ompK36 with a putative promoter site of rcsD.

The figure shows the upstream region of ompK36 with a putative promoter site of rcsD. The predicted -35 and -10 sequences of putative promoter

of rcsD is denoted by red box. The primer site is labelled and the direction of the primer is showed by red arrow. The putative start codon for

rcsD is signposted by thick red arrow. Promoter site was predicted by BPROM (Softberry, Inc., NY).

255 |

Figure 6-12A) Multiple sequence alignment of upstream region of ompK36

256 |

Figure 6-12 B) Multiple sequence alignment of downstream region of ompK36

257 |

Figure 6-12 Multiple sequence alignment of upstream and downstream regions of ompK36.

The figures show the multiple sequence alignments of A) upstream and B) downstream region of ompK36. The start codon of ompK36 is labelled

with ATG and red arrow while the stop codon of ompK36 is labelled with TAA and asterisk. The putative start codon for apbE is signposted by

thick red arrow. The predicted -35 and -10 sequences of putative promoters for A) ompK36 and B) apbE are denoted by red boxes. The putative

terminator site is indicated by the yellow line. The primers sites used for the development of mutant constructs are as labelled and the direction

of the primers are showed by red arrows. The nucleotide changes are indicated by blue box. The deduced amino acid sequences encoded by

apbE are written in blue. Promoter sites were predicted by BPROM (Softberry, Inc., NY) and terminator site was predicted by ARNold Finding

Terminators software (Naville, et al., 2011). The multiple sequence alignments were conducted using GeneDoc.

258 |

6.3.2 Analysis of OmpC, OmpK35 and OmpK36 protein

structure

In order to approximate a structural characterisation of these three porins

which could allow us to understand possible function of OmpK35 and

OmpK36, we performed a comparative analysis of 3D structures of OmpC of

E. coli and OmpK36 of K. pneumoniae, and between OmpK36 and putative

3D structure of OmpK35. It was previously reported that OmpC showed

~ 80 % identity to OmpK36 but there is no study comparing the 3D protein

structures.

The multiple sequence alignment of OmpC, OmpK36 and OmpK35 derived

from Kpneu_1 is presented in Figure 6-13. The multiple sequence alignments

were conducted using Clustal Omega and GeneDoc. The secondary

structures such as β-strands, loops, and short turns of these OmpC, OmpK36

and OmpK35 are as labelled in Figure 6-13. Based on the analysis, the

percentage of amino acid identity between OmpC and OmpK36 is 80.12 %,

while the percentage of amino acid identity between OmpK36 and OmpK35 is

59.35 %. We generated the image of 3D structures of OmpC, OmpK36 and

OmpK35 using Chimera 1.12rc. The 3D structures of templates, OmpC and

OmpK36 are shown in Appendix 14 and Appendix 15, respectively. While the

predicted structure of OmpK35 is shown in Appendix 16. Using Chimera

1.12rc, we analysed and compared the amino acid sequences and also

conducted the structural superimposition between these two porins, OmpC

and OmpK36. Based on the high percentage of amino acid identity, 80.12 %,

we hypothesised that these 3D structures would be highly similar. It is thought

that when the overall sequence identity is above 40 %, almost 90 % of the

259 |

main-chain atoms are likely to be modeled with a root-mean-square error

(RMSE) around 1 Å (Fiser, 2010; Sánchez, et al. , 1998).

The structural alignment results showed that the 3D structures of OmpC and

OmpK36 monomers were highly similar with a structural alignment RMSD

value of 0.894 Å as shown in Appendix 17. Besides, our analysis data

suggested that structural alignment of OmpK35 using OmpK36 as a template

also showed the integrity of the conserved domain since both of these porins

had a low RMSD value, 0.488 Å which means that the 3D structures were

highly similar Appendix 18.

Even though, OmpC and OmpK36 do have similar 3D structures, and high

percentage in amino acid sequence identity, we, however, cannot confirm that

both of these porins will have exactly the same functions. Similarity in 3D

structures and amino acid sequences does not necessarily lead to the same

function. The protein conformation and protein folding might be different and

to confirm the exact protein function between these two porins, it needs a

further investigation and qualification.

260 |

Figure 6-13 Amino acid multiple sequence alignment of OmpC, OmpK36 and OmpK35.

The figure shows the amino acid multiple sequence alignment of OmpC of E. coli, OmpK36 of K. pneumoniae, and OmpK35 derived from

Kpneu_1. 2J1N is a PDB ID for OmpC of E. coli, 1OSM is a PDB ID for OmpK36 of K. pneumoniae, and P_OmpK35 are deduced amino acid

sequences of OmpK35 derived from Kpneu_1. The analysis and alignment were done using Clustal Omega and GeneDoc. The secondary

structures were determined using Kabsch-Sander algorithm available in the dictionary of protein secondary structure (DSSP) (Kabsch, et al. ,

1983) in the Research Collaboratory for Structural Bioinformatics- Protein Data Bank (RCSB PDB [https://www.rcsb.org/pdb/]), and they were

also adapted from the study by Doménech-Sánchez, et al. (2003). The putative OmpK35 secondary structure was predicted based on the 1OSM

template using SWISS-MODEL and Chimera 1.12rc. The primary conserved region is highlighted in black and the secondary conserved

substitutions are highlighted in yellow. The black line indicates the β-strand, white boxes indicate the loop structure, orange boxes indicate short

turns and pink arrow indicates the β-strand 8.

261 |

6.3.3 Amplification of ompK35 and ompK36

As described in the General introduction, the aim of this study was to develop

the ompK35 or ompK36 mutant strains and compare their capacity to form

biofilm with their wild type strains. In order to knock out ompK35 or ompK36

in K. pneumoniae, all the strains were initially confirmed to harbour both genes

by PCR amplification. The ompK35 and ompK36 genes were amplified using

genomic DNA from those six strains of K. pneumoniae as a template. Primers

were designed to amplify the ompK35 and ompK36 with ~70 bp of flanking

regions. Figure 6-14 shows a PCR amplification of ompK35 and ompK36 for

all the strains. The amplicon size predicted for ompK35 was 1244 bp (Figure

6-14A, lane 1-6) and the amplicon size for ompK36 was 1276 bp for Kpneu_1,

1256 bp for Kpneu_63, Kpneu_64, and Kpneu_65, 1262 bp for Kpneu_110

and 1250 bp for C3091 (Figure 6-14B, lane 1-6). The amplicons were

subsequently purified and sequenced for confirmation.

262 |

Figure 6-14 Screening for ompK35 and ompK36.

The screening for ompK35 and ompK36 genes was conducted by PCR using

genomic DNA from six K. pneumoniae strains. Lanes 1-6 are the amplicons of the

ompK35 or ompK36 genes obtained from genomic DNA of Kpneu_1, Kpneu_63,

Kpneu_64, Kpneu_65, Kpneu_110, and C3091 respectively. Lane M is a DNA

marker, HyperLadder I (Bioline, UK), and lane 7 is the negative control (no template

control) for both of the amplifications. A) The ompK35 gene was amplified using the

OmpK35_FL_F and OmpK35_FL_R primer set and the amplicons were observed at

1244 bp. Nonspecific ~500 bp amplicons were seen faintly for all the strains and an

additional ~300 bp amplicon for strain Kpneu_1. This could be due to the nonspecific

binding of the primer to the DNA template. B) The ompK36 gene was amplified using

OmpK36_FL_F and OmpK36_FL_R primer set and the expected amplicons at

1276 bp for Kpneu_1, 1256 bp for Kpneu_63, Kpneu_64, Kpneu_65, 1262 bp for

Kpneu_110 and 1250 bp for C3091 were observed for all of the samples (Lanes 1-

6). Expected size for ompK35 and ompK36 amplicons is indicated by arrows.

263 |

6.3.4 Development of mutant constructs

Upon confirmation of ompK35 and ompK36 in all the strains used in this

chapter, we then continued to develop the mutant constructs for deletion of

ompK35 and ompK36 by a homologous recombination approach. The mutant

constructs were developed using SOE PCR method. The genomic DNA of

K. pneumoniae and C. difficile were extracted and used as template to amplify

individual ~ 500 bp upstream region and ~ 500 bp downstream region from

the gene to be deleted, and erm1(B) (790 bp) gene respectively. The

individual PCR fragments were ligated by SOE PCR. The SOE PCR of

ompK35 and ompK36 mutant cassettes with erm1(B) as a selective marker

was conducted using 35_KO1_3 and 35_KO4_2 or the 36_KO1_2 and

36_KO4_3 primer sets. The expected amplicon for the ompK35 mutant

cassette was 1654 bp. While the expected amplicon for the ompK36 mutant

cassettes were 1760 bp for Kpneu_63, 1762 bp for Kpneu_65 and 1761 bp

for C3091 (Figure 6-15).

Based on the multiple sequence alignments of the ompK35 and ompK36

flanking regions between the six strains, ten nucleotide differences in the

upstream and one nucleotide difference in the downstream region of ompK35

were identified and all of them were located in non-coding regions (see Figure

6-10). The mutant construct for ompK35 was then developed based on the

flanking regions, which was derived from Kpneu_1 strain since the nucleotide

differences did not result in any amino acid substitution and they were not

located on any regulatory sites such as predicted promoter. In addition, eight

nucleotide differences in the upstream region and three nucleotide differences

264 |

in the downstream region of ompK36 were identified from the multiple

sequence alignments of ompK36 flanking region of six strains (Figure 6-12).

All of the other differences were located in non-coding regions apart from the

nucleotide differences at position 173 (A→T). The nucleotide differences at

position 173 resulted in modification of protein sequence (Q58L) of apbE (see

Figure 6.12). For that reason, the mutant constructs of ompK36 for strain

Kpneu_1, Kpneu_65, and C3091 which had glutamine could be used

interchangeably between respective strains since they encoded the same

amino acid at position 56, as well as the constructs for Kpneu_63, Kpneu_64,

and Kpneu_110 which had the same amino acid, leucine at position 56 (see

Figure 6-12).

Figure 6-15 Amplification of ompK35 and ompK36 mutant cassettes.

The figure shows the amplification product of the mutant cassettes by PCR. The

pGEMT_ erm_ΔompK35 and pGEMT_erm_ΔompK36 constructs were used as a

template for amplification of mutant cassettes comprising erm1(B) and ~500 bp of

265 |

upstream and ~500 bp downstream regions of ompK35 or ompK36. Lane 1 and lane

6 are the negative controls (no template control). Lane 2 is the ΔompK35-1 cassette,

which was derived from Kpneu_1 genomic DNA. Lane M is a DNA marker,

HyperLadder I (Bioline, UK). Lane 3 is the ΔompK36-63 cassette, Lane 4 is the

ΔompK36-65 cassette and Lane 5 is the ΔompK36-C3091 cassette, which were

obtained from the Kpneu_63, Kpneu_65, and C3091 genomic DNA respectively.

6.3.5 Development of ΔompK35 or ΔompK36 strains

After construction of pGEMT_erm_ΔompK35 and pGEMT_erm_ΔompK36,

the mutant cassettes were amplified using HF-PCR, and the gel purified PCR

product was electroporated into the electocompetent K. pneumoniae strains.

After 16 h of incubation, uncountable colonies were obtained on LB agar

supplemented with erm (~ 400 µg/ml). However, no expected bands were

observed after PCR screening using 35_KO1_2 and 35_KO4_3 or 36_KO1

and 36_KO4_4 primer sets (Figure 6-16). The unsuccessful homologous

recombination was possibly because of the existence of RecBCD pathway in

K. pneumoniae, and the unsuitable selectable marker (erm1(B) that was used

in the mutant construct (this issue will be discussed later in the discussion).

Because of this reason, new mutant constructs were developed and a

recombineering method using modified helper plasmid pKD46-apra was

adapted in order to obtain the mutant strains (section 6.2.4.3 and 6.2.4.4).

The MIC of spectinomycin and apramycin for all tested strains were

determined to identify a selectable marker for reconstruction of the mutant

constructs and modification of the helper plasmid. Of six strains, only three

strains had MICs for spectinomycin which would allow for mutant

development: Kpneu_1, Kpneu_110, and C3091. Growth of these three

strains was inhibited at 100 µg/ml of spectinomycin (the concentration used

266 |

for selection) compared to the other three strains which could tolerate higher

concentration of spectinomycin (> 300 µg/ml). Kpneu_1, Kpneu_110, and

C3091 were therefore selected for mutant development. In addition, all of the

strains showed MICs of apramycin at 16 µg/ml, which was lower than the

concentration for selection (50 µg/ml).

Table 6-3 The MIC value of spectinomycin and apramycin against

K. pneumoniae.

Antibiotic MIC (µg/ml)

Strain Spectinomycin Apramycin

Kpneu_1

12.5

16

Kpneu_63

>300

16

Kpneu_64

>300

16

Kpneu_65

>300

16

Kpneu_110

12.5

16

C3091

12.5

16

Based on the MIC value, the spectinomycin resistance gene, aad9 was then

chosen to replace erm1(B) in pGEMT_erm_ΔompK35 or pGEMT_erm_

ΔompK36 mutant constructs, whilst, apramycin resistance gene, aac(3)-IV

was selected to replace bla in pKD46 helper plasmid (section 6.2.4.4) since

all the strains used in this study are resistant to ampicillin, and using bla as a

selective marker would result in no selection.

267 |

The pGEMT_spc_ΔompK35 or pGEMT_spc_ΔompK36 constructs were

developed by amplifying the aad9 and pGEMT backbone with the flanking

region by inverse PCR (see Figure 6-4). The insert and the backbone were

then digested and ligated before transforming into E. coli α-select competent

cells. The confirmed mutant constructs with aad9 as a selective marker were

used as a template for HF-PCR to amplify the mutant cassettes using primer

pairs 35_KO1_3 and 35_KO4_2 for ΔompK35 cassette and 36_KO1_2 and

36_KO4_3 for ΔompK36 cassette. The amplified PCR products were purified

using gel extraction kit prior to electroporation into electrocompetent

K. pneumoniae harbouring pKD46 cells.

Prior to electroporation of purified PCR product into electrocompetent cells

harbouring pKD46-apra cells, the helper plasmid was first modified. The

pKD46 plasmid was modified by replacing the bla in pKD46 with aac(3)-IV by

the same method as previously mentioned (Figure 6-4) and designated as

pKD46-apra. Next, pKD46-apra was electroporated into electrocompetent

Kpneu_1, Kpneu_110, and C3091 cells and designated as Kpneu_1/pKD46-

apra, Kpneu_110/pKD46-apra and C3091/pKD46-apra. Electrocompetent

Kpneu_1/pKD46-apra cells were then prepared using the method explained in

4.2. The gel extracted HF mutant cassettes were electroporated into

electrocompetent Kpneu_1/pKD46-apra cells and grown on LB-spc 100 µg/ml

at 37 C for 16 h. The colonies obtained from electroporation were screened

for the correct bands to confirm that allelic exchanged had occur by PCR. The

PCR was conducted using the primer sets that were designed based on the

chromosomal regions which are adjacent to the mutant cassettes (for the

purpose of explanation, these primers will be referred as an external primer

268 |

(see Figure 6-16), the specific aad9 primer set and the target gene ompK35 or

ompK36 primer sets.

Figure 6-16 Location of the external primers for confirmation of mutant

strains.

The figure shows the location of the external primers that are located at the

chromosomal regions adjacent to the mutant cassettes. The external primers were

used to distinguish between the mutant strains and wild type strains. These primers

were used to confirm occurrence of gene replacement by amplifying the mutant

cassettes including flanking regions of the gene of interest, aad9 selectable marker

gene (for mutant strains) or ompK35/ompK36 (for wild type strains) and a

chromosomal region adjacent to the cassettes. The expected amplicon size for each

strain is as shown. Flanking region is indicated by blue boxes and K. pneumoniae

chromosomal region adjacent to the mutant cassettes is denoted by dashed line.

269 |

The expected band for the Kpneu_1 wild type strain was 2142 bp and

2224 bp for ompK35 and ompK36, respectively. While, the expected band

for its isogenic mutant strains, ΔompK35 and ΔompK36 were 1839 bp and

1883 bp, respectively. A detectable reduction in the size of wild type strain

and mutant strain was observed on a 1.0 % agarose gel by electrophoresis.

Using the same protocol of electroporation, ΔompK35 and ΔompK36 isogenic

mutant strains of C3091 were also successfully obtained (Figure 6-17). The

expected band for the C3091 wild type strain was 2142 bp and 2242 bp for

ompK35 and ompK36, respectively. While the expected band for its isogenic

mutant strains, ΔompK35 and ΔompK36 were 1839 bp and 1894 bp,

respectively. Strain Kpneu_110, however could not be genetically modified.

Kpneu_110 was notable because it had a hypermucoviscous morphology

(discussed in the discussion below).

270 |

Figure 6-17 Confirmation of mutant strains by PCR.

The amplification of the chromosomal regions adjacent to mutant cassettes of

A) ompK35 and B) ompK36 of Kpneu_1, and C) ompK35 and D) ompK36 of C3091.

Confirmation of the gene replacement can be observed by the differences in the

amplicon size amplified by the external primers between wild type and mutant

strains. Expected size for ompK35 and ompK36 of wild type strain can be seen at

2142 bp or 2224 bp for Kpneu_1, and 2142 bp and 2242 bp for C3091, respectively.

The expected size for Kpneu_1∆ompK35 and C3091∆ompK35 is 1839 bp. While,

the expected size for Kpneu_1∆ompK36 and C3091∆ompK36 is 1883 bp and 1894,

respectively. M: HyperLadder I (Bioline, UK). Lane 1A and 1B: Genomic DNA from

Kpneu_1; Lane 1C and 1D: Genomic DNA from C3091; Lane 2A: Genomic DNA

from Kpneu_1∆ompK35; Lane 2B: Genomic DNA from Kpneu_1∆ompK36; Lane 2C

and 3C: Genomic DNA from C3091∆ompK35; Lane 2D and 3D: Genomic DNA from

C3091∆ompK36; Lane 3A, 3B, 4C and 4D: negative control (no template control).

271 |

6.3.6 Development of complemented strains

6.3.6.1 Construction of ompK35 complementation

construct harbouring tet(M)

Complementation constructs were developed by cloning ompK35 or ompK36

with their putative promoters into pUC19-tet(M). The tet(M) resistance gene

was cloned in pUC19 plasmid vector by Dr Supathep Tansirichaiya (Eastman

Dental Institute, UCL). For the purpose of explanation, only the results of the

ompK35 complementation construct of Kpneu_1 and C3091 will be described

in detail since no complementation construct was successfully obtained for

ompK36. The ompK35 fragments were amplified using primer sets of

OmpK35_SphI_GC_F_2 and OmpK35_SphI_GC_R_2 from Kpneu_1 and

C3091 genomic DNA as template. A 1506 bp PCR product was purified, and

the purified PCR product and pUC19-tet(M) were digested with SphI. The

digested pUC19-tet(M) was then dephosphorylated and ligated with the

purified digested ompK35 fragment. Figure 6-18 below shows all the

expected fragments. The amplicons of ompK35 is 1506 bp. The digested

bands of pUC19-tet(M) and target genes were expected at 5276 bp and 1506

bp, respectively. While the ligation of pUC19-tet(M) and the targeted genes

was confirmed by existence of multiple faint bands at ~ 8000 bp (ligation

product of 5276 bp of pUC19-tet(M) back bone and 1506 bp of ompK35).

272 |

Figure 6-18 Construction of ompK35 complementation constructs.

The figure shows A) Simulation of the SphI digestion pattern of pUC19-tet(M) and

ompK35 amplified fragment on a 1 % agarose gel. The pUC19-tet(M) backbone is

observed at 5276 bp while digested ompK35 fragment is at 1506 bp . Simulation

was run using Snap Gene (GSL Biotech LLC, US), B) The agarose gel shows

plasmid DNA, amplified PCR product, digested product and ligation product of

complementation construct. Lane 1: pUC19-tet(M) plasmid DNA. Lane 2: ompK35

with promoter amplicon derived from Kpneu_1. Lane 3: ompK35 with promoter

amplicon derived from C3091. Lane 4: purified digested plasmid. Lane 5: purified

SphI digested fragment of Kpneu_1. Lane 6: purified SphI digested fragment of

C3091. Lane 7: Ligated product of pUC19-tet(M) and ompK35 of Kpneu_1. Lane 8:

Unligated control of pUC19-tet(M) and ompK35 from Kpneu_1. Lane 9: Ligated

product of pUC19-tet(M) and ompK35 of C3091. Lane 10: Unligated control of

pUC19-tet(M) and ompK35 from C3091. M: HyperLadder I, (Bioline UK). All of the

expected bands were denoted by red arrow. Expected bands can be seen for

plasmid DNA, and amplified product upon amplification. The 5276 bp band for

pUC19-tet(M) backbone and 1506 bp of ompK35 fragments were also observed from

the gel. The ligation was suspected to occur successfully since multiple bands were

observed upon ligation in comparison to the unligated control (without T4 DNA

ligase) which showed only two bands; 5276 bp plasmid backbone and 1506 bp

ompK35 fragments. Our obtained gel also showed the same bands as predicted by

Snap Gene (GSL Biotech LLC, US). The plasmid backbone and insert are indicated

by red arrows.

A) B)

273 |

After transformation of the ligated product into E. coli competent cells, the

colonies were selected and grown for plasmid extraction in LB-tet 12.5 µg/ml

broth. The plasmids were digested and the plasmids with the expected insert

(Figure 6-19) were sent for sequencing.

Figure 6-19 Confirmation of the complementation constructs by digestion.

The figure shows the electrophoresis gel of the restriction digest to confirm the

plasmid constructs. M: HyperLadder I (Bioline, UK). Lane 1-8: Digested product for

putative Kpneu_1ompK35ompK35 constructs. Lane 9-16: Digested product for

putative C3091ompK35ompK35 constructs. Lane 1, 3, 5, 7, 9, 11, 13 and 15:

Undigested plasmid. Lane 2, 4, 6, 8, 10, 12, 14, and 16: SphI digested plasmid. No

expected bands were seen for putative Kpneu_1ompK35ompK35 construct upon

digestion. However, three C3091ompK35ompK35 constructs were seen to

harbour a correct construct by showing ~5500 bp of pUC19-tet(M) backbone and

~1500 of ompK35 with promoter fragment. Constructs 10, 12, and 16 were then sent

for sequencing. The plasmid backbone and insert are indicated by red arrows.

Unfortunately, upon sequence analysis, the constructs did not harbour any of

the expected insert, ompK35 with putative promoter.

274 |

6.3.7 Antimicrobial susceptibility testing

It is very important to determine the factors contributing to the emergence of

antimicrobial resistant K. pneumoniae as it is one of the significant growing

concerns in healthcare settings (O'Neill, 2016). Other than acquisition of

antimicrobial resistance genes, impermeability of the outer membrane layer

(Nikaido, 2003), loss or modification of porin(s) (Pagès, et al., 2008) are also

known to contribute to these factors.

To investigate the role of OmpK35 and OmpK36 in antibiotic and antiseptic

resistance, the antimicrobial susceptibility testing was conducted on the wild

type strains and their isogenic mutant strains. The antibiotics tested in this

study were from different families and had different modes of action. They

were -lactam (cefotaxime), fluoroquinolone (ciprofloxacin), tetracycline

(tetracycline) and the aminoglycoside (gentamicin). While the quaternary

ammonium compound (QAC) antiseptics were cetylpyridinium chloride (CPC)

and cetyltrimethylammonium bromide (CTAB). Due to growing number of

hospital-acquired K. pneumoniae infection cases, it has become important to

investigate the resistance mechanism of QACs in K. pneumoniae as QACs

are among the main components for detergents and antiseptics used in

hospitals.

275 |

Table 6-4 summarises the median MIC (µg/ml) and the range of cefotaxime,

ciprofloxacin, tetracycline, gentamicin, CPC, and CTAB tested for Kpneu_1

and C3091, and their isogenic mutants: Kpneu_1ompK35,

Kpneu_1ompK36, C3091ompK35, and C3091ompK36. The clinical

strain, Kpneu_1 was, in general resistant to more antibiotics compared to the

laboratory strain, C3091 (based on the EUCAST breakpoint). Deletion of

ompK35 and ompK36 was observed to cause increased resistance to several

antibiotics especially to cefotaxime and tetracycline. The increased

resistance to cefotaxime was observed in C3091ompK36 which was

2.5 times higher compared to its wild type and C3091ompK35 counterpart.

However, the increased resistance in strain Kpneu_1 and its isogenic mutants

could not be identified since the median MIC of cefotaxime for all the strains

was > 32 µg/ml and this value is out of the E- test range for cefotaxime.

Moreover, increased resistance to tetracycline was also observed in

Kpneu_1ompK35 and C3091ompK35 compared to their wild type and

ompK36 mutant counterparts. The median MIC of tetracycline for

Kpneu_1ompK35 was 0.75 µg/ml, which was 1.5 times higher than the

median MICs of tetracycline for Kpneu_1 and Kpneu_1ompK36, which were

at 0.5 µg/ml. While the median MIC of tetracycline for C3091ompK35

(1.5 µg/ml) was 1.5 times higher than the median MIC of C3091 (0.75 µg/ml)

and 3 times higher than the median MIC of C3091ompK36 which was only

0.5 µg/ml.

276 |

Table 6-4 Median MIC of antibiotics and antiseptics for Kpneu_1, C3091, and their isogenic mutants.

Strain Kpneu_1 Kpneu_1∆ompK35 Kpneu_1∆ompK36 C3091 C3091∆ompK35 C3091∆ompK36

Median MIC (µg/ml)

Range

(µg/ml)

Median MIC (µg/ml)

Range

(µg/ml)

Median MIC (µg/ml)

Range

(µg/ml)

Median MIC (µg/ml)

Range

(µg/ml)

Median MIC (µg/ml)

Range

(µg/ml)

Median MIC (µg/ml)

Range

(µg/ml)

E-test MIC

Cefotaxime >32 >32 >32 >32 >32 >32

0.047 (0.032-0.047)

0.047 0.047 0.125 (0.094-0.125)

Ciprofloxacin 0.016 0.016 0.016

(0.016-0.032)

0.016 (0.016-0.032)

0.012 (0.012-0.016)

0.016

0.016

0.012

(0.012-0.016)

Tetracycline 0.5

(0.38-0.5)

0.75 (0.75-1.5)

0.5 0.5 0.75 (0.38-0.75)

1.5 1.5 0.5 0.5

Gentamicin 6.0 6.0 6.0 6.0 4.0

(4.0-6.0)

0.38 0.38 0.25 0.25 0.25 0.25

Broth microdilution

CPC 64 64 64 64 64 64

64 64 64 64 64 64

CTAB 128 128 128 (64-128)

128 (64-128)

128 128 128 128 128 128

*median is from three independent experiments

277 |

6.3.8 Growth kinetics of wild type and mutant strains upon

cultivation in TSB and AUM

Growth curves performed in TSB and AUM for Kpneu_1, C3091, and their

isogenic mutant strains showed that all the strains grew faster in TSB

compared to AUM. The difference can be seen from the growth curve which

showed that all the strains reached mid-log phase quicker in TSB in

comparison to AUM (Figure 6-20 and Figure 6-21), and also by the growth rate

constant (µ) and doubling time (td) which are presented in Table 6-5. The

growth rate constant and doubling time was determined using the calculation

mentioned in Chapter 4 section 4.2.3.1. The growth curves of Kpneu_1 and

C3091 wild type strains were almost similar with their isogenic mutants upon

cultivation in TSB and AUM. All strains reached mid-log phase after 2 h

cultivation in TSB (Figure 6-20) while they approached mid-log phase after

4 h cultivation in AUM (Figure 6-21).

278 |

279 |

Figure 6-20 Growth curve of Kpneu_1, C3091, and isogenic mutants in TSB.

The figure shows growth curve of A) Kpneu_1 and its isogenic mutant strains and B) C3091 and its isogenic mutant strains in TBS plotted in two

different scales: i) linear scale and ii) logarithmic scale. Data presented are the mean of two independent experiments and error bars indicates

the standard deviations. All of the strains showed almost similar growth rate upon cultivation in TSB especially when reaching the exponential

phase. Mid-log phase (indicated by the red arrow) of all the strains was observed after 2 h. The OD1/t0 and OD2/t for growth rate constant and

doubling time determination are as labelled. The graph lines are indicated as follow: blue: Kpneu_1, turquoise: Kpneu_1∆ompK35, light orange:

Kpneu_1∆ompK36, red: C3091, dark green: C3091∆ompK35 and purple: C3091∆ompK36. The OD1 and OD2 are the absorbance value and t0

and t are the time, at late and early log phase used for calculation based on the equation in Chapter 4 section 4.2.3.1

280 |

281 |

Figure 6-21 Growth curve of Kpneu_1, C3091, and isogenic mutants in AUM.

The figure shows growth curve of A) Kpneu_1 and its isogenic mutant strains and B) C3091 and its isogenic mutant strains in AUM plotted in

two different scales: i) linear scale and ii) logarithmic scale. Data presented are the mean of two independent experiments and error bars

indicates the standard deviations. All of the strains showed almost similar growth rate upon cultivation in AUM especially when reaching the

exponential phase. Mid-log phase (indicated by the red arrow) of all the strains was observed after 4 h. The OD1/t0 and OD2/t for growth rate

constant and doubling time determination are as labelled. The graph lines are indicated as follow: blue: Kpneu_1, turquoise: Kpneu_1∆ompK35,

light orange: Kpneu_1∆ompK36, red: C3091, dark green: C3091∆ompK35 and purple: C3091∆ompK36. The OD1 and OD2 are the absorbance

value and t0 and t, are the time at late and early log phase used for calculation based on the equation in Chapter 4 section 4.2.3.1

282 |

The growth rate constant (µ) and doubling time (td) of Kpneu_1 and C3091,

and their isogenic mutant strains were determined according to the calculation

showed in Chapter 4, section 4.2.3.1. Doubling time for strain Kpneu_1 and

C3091 and their isogenic ∆ompK35 strains, Kpneu_1∆ompK35 and

C3091∆ompK35 was 35 min in TSB. While their isogenic mutants,

Kpneu_1∆ompK36 and C3091∆ompK36 had doubling time of 29 min. The

doubling time for all the wild type and mutant strains upon cultivation in AUM

(Figure 6-21) was longer compared to TSB. The doubling times of strain

Kpneu_1 and C3091 upon cultivation in AUM were 103 min and 100 min,

respectively. Their mutant strains showed slightly shorter doubling times

although no obvious differences could be seen from the growth curve. The

doubling times of Kpneu_1∆ompK35 and Kpneu_1∆ompK36 upon cultivation

in AUM were 100 min and 97 min, respectively. While the doubling times of

C3091∆ompK35 and C3091∆ompK36 upon cultivation in AUM were 96 min

and 99 min, respectively (Table 6-5).

283 |

Table 6-5 Growth rate constant (µ) and doubling time (td) of Kpneu_1 and C3091, and their isogenic mutant strains upon cultivation

in TSB and AUM.

Strain

Media

Equation 1

(h-1)

Equation 2

(min)

µ= (log10 OD2 – log10 OD1) 2.303/ (t- t0) td = ln 2 / µ

Kpneu_1 TSB 1.20 35

AUM 0.40 103

Kpneu_1∆ompK35 TSB 1.17 35

AUM 0.42 100

Kpneu_1∆ompK36 TSB 1.40 29

AUM 0.43 97

C3091 TSB 1.20 35

AUM 0.41 100

C3091∆ompK35 TSB 1.18 35

AUM 0.43 96

C3091∆ompK36 TSB 1.31 32

AUM 0.42 99

284 |

6.3.9 Role of OmpK35 and OmpK36 in biofilm formation by

K. pneumoniae in 96-well polystyrene tissue culture treated

plates

In order to determine a role of OmpK35 and OmpK36 in biofilm formation by

K. pneumoniae, both of the strains, Kpneu_1 and C3091 and their isogenic

mutants were used in this experiment. No significant difference was seen

between Kpneu_1 and its isogenic mutant strains upon cultivation in either

medium. In addition, no significant difference in biofilm formation by strain

C3091 and C3091∆ompK35 could be seen upon cultivation in TSB. However,

strain C3091 was found to form slightly lower biofilm

(p < 0.05) in AUM compared to strain C3091∆ompK36 (Figure 6-22).

285 |

Figure 6-22 Biofilm formation of Kpneu_1, C3091, and their isogenic mutants

in TSB and AUM.

The graphs show biofilm formation by A) Kpneu_1 and its isogenic mutant and

B) C3091 and its isogenic mutant upon cultivation in i) TSB and ii) AUM. Bars show

average reading obtained using CV assays from three independent experiments. The

data was analysed by a Kruskal-wallis with Dunn’s multiple comparison. The error

bars indicate standard deviation of n=15. No significant difference can be seen

between the Kpneu_1 wild type strain and its isogenic mutant strains upon cultivation

in either of the medium. No significant difference was seen between C3091 and

C3091∆ompK35 upon cultivation in TSB, while a significant higher of biofilm formation

was seen between C3091∆ompK36 compared to C3091 upon cultivation in AUM.

Statistical significance is indicated by * represents p < 0.05.

286 |

6.3.10 Role of OmpK35 and OmpK36 in biofilm formation

by K. pneumoniae on catheter pieces

The role of OmpK36 and OmpK35 in biofilm formation were next determined

in a more clinically relevant assay. The capacity of strain Kpneu_1 and its

isogenic mutant to form biofilms on catheter pieces was evaluated by

cultivating these strains in TSB. While the capacity of the other, strain C3091

and its isogenic mutants, were evaluated by growing these strains in AUM.

The media was chosen based on our previous data in Chapter 4 (Figure 4-10),

which showed that strain Kpneu_1 formed more biofilm in TSB while strain

C3091 formed more biofilm upon cultivation in AUM.

The capacity of Kpneu_1, C3091 and their isogenic mutants to adhere on

silicone catheter surfaces and eventually form a biofilm was evaluated by CV

staining, and enumeration of the cells harvested from the surface of the

catheter was also conducted. No significant difference was seen between the

wild type and mutant strains upon CV staining and enumeration of the cells.

No significant difference of biofilm formation of Kpneu_1 wild type,

Kpneu_1∆ompK35, and Kpneu_1∆ompK36 was observed upon cultivation of

these strains on catheter surfaces in TSB under shaking conditions. The

C3091 wild type, C3091∆ompK35, and C3091∆ompK36 were also shown to

have similar capacity to form biofilms on catheter surfaces upon cultivation in

AUM under the same condition as Kpneu_1 and its mutants.

287 |

Figure 6-23 Biofilm formation and cell viability of Kpneu_1 and its isogenic

mutants on silicone catheter surfaces.

The graph shows biofilm formation by Kpneu_1 and its isogenic mutants on silicone

catheter surfaces upon cultivation in TSB at 37 °C for 24 h. A) The biofilm was

quantified by CV staining. The error bars indicate standard deviation from three

independent experiments with data points for each strain as represented. B) Box

and whiskers plot shows the CFU/cm2 on catheter surfaces of Kpneu_1 and its

isogenic mutants post 24 h cultivation in TSB. Bars show A) average absorbance

reading obtained using CV staining assay and B) median value of CFU/cm2 from three

independent experiments (average value is signposted by +). The data was analysed

by a Kruskal-Wallis with Dunn’s multiple comparison test. The blank was TSB without

any inoculated bacteria. The error bars indicate standard deviation of nine replicates.

No significant difference could be seen between Kpneu_1 and its isogenic mutant

strains upon cultivation in TSB in terms of biofilm formation on silicone catheter pieces

and cell viability harvested from the silicone catheter pieces.

288 |

Figure 6-24 Biofilm formation and cell viability of C3091 and its isogenic

mutants on silicone catheter surfaces.

The graph shows biofilm formation by C3091 and its isogenic mutants on silicone

catheter surfaces upon cultivation in AUM at 37 °C for 24 h. A) The biofilm was

quantified by CV staining. The error bars indicate standard deviation from three

independent experiments with data points for each strain as represented. B) Box and

whiskers plot shows the CFU/cm2 on catheter surfaces of C3091 and its isogenic

mutants post 24 h cultivation in AUM. Bars show A) average absorbance reading

obtained using CV staining assay and B) median value of CFU/cm2 from three

independent experiments (average value is signposted by +). The data was analysed

by a Kruskal-Wallis with Dunn’s multiple comparison test. The blank was AUM without

any inoculated bacteria. The error bars indicate standard deviation of nine replicates.

No significant difference could be seen between C3091 and its isogenic mutant

strains upon cultivation in AUM in terms of biofilm formation on silicone catheter

pieces and cell viability harvested from the silicone catheter pieces.

289 |

6.4 Discussion

In the initial part of this study, the University of California, San Francisco,

UCSF Chimera 1.12rc programme (Pettersen, et al., 2004); an open access

molecular modelling software for visualisation, and analysis of protein

structure and functional modelling was used to illustrate and superimpose the

3D structures of OmpC, OmpK36, and predicted structure of OmpK35.

Structure-based alignment is known to be more reliable than sequence-based

alignment since protein structure is thought to be more conserved than

sequence (Chothia, et al. , 1986).

This software was chosen since it provides access to many analysis tools that

were important in our study such as sequence alignment, structure alignment

or superimposition. Protein structures can be visualised automatically with

their amino acid sequences. In addition, this programme can superimpose two

different structures without the existence of pre-existing sequence alignment,

and it can match the sequences and secondary structures of two proteins,

even if they have very low sequence identity. Our protein structural analysis

data demonstrated that OmpC of E. coli and OmpK36 of K. pneumoniae, and

also OmpK36 and OmpK35 have a high degree of structural similarity and a

high percentage of amino acid sequence identity. We then hypothesised that

OmpC, OmpK36 and OmpK35 could possibly play similar functions since they

were shown to have a high degree of structural similarity. Our hypothesis was

also based on the concept that orthologs, possibly play the same or similar

functions in different species (Lee, et al. , 2007). This concept, however, is not

applicable for all orthologs since the specific protein conformation or protein

290 |

fold adopted by any particular protein does not necessarily lead to the same

function (Lee, et al., 2007).

We then developed our ompK35 and ∆ompK36 mutant strains of Kpneu_1

and C3091 by site-directed mutagenesis by homologous recombination using

λ-red recombinase system. Our result showed that a single deletion of

ompK35 and ompK36 occurred by replacement of these genes in the genome

of wild type strains with aad9 from the mutant cassettes. Unfortunately, no

complemented strains were obtained and this will be discussed later.

At the beginning of this study, ompK35 and ompK36 mutant constructs were

designed with erm1(B) as a selective marker. Unfortunately, we had

difficulties with the screening for putative mutant colonies. Uncountable

K. pneumoniae colonies were found to grow on LB plates even with a very

high concentration of erythromycin (~ 400 µg/ml). The background growth of

non-recombinant negative control colonies was also observed on this

selective medium. Hence, the mutant constructs were then re-developed

using a different selective marker to eliminate the appearance of false positive

colonies. The growth of false positive colonies could be due to an intrinsic

resistance of K. pneumoniae to erythromycin.

We concluded that erm1(B) was not a suitable marker to be used for

K. pneumoniae mutagenesis, even though all the strains used in this study

did not harbour erm1(B). This is probably because the outer membrane (OM)

layer of K. pneumoniae is composed of asymmetric bilayer of LPS and

phospholipids. It is proven that intrinsic resistance against erythromycin could

291 |

be due to intact hydrophilic LPS at the OM layer of several Gram-negative

bacteria, which results in slow passive diffusion of erythromycin and

eventually provides a permeability barrier to this lipophilic drug (Nikaido,

2003). The erythromycin resistance also might be due to the existence of an

active efflux pump, which pumps the antibiotics out of the cells through the

membrane that leads to less concentrated antibiotics in the cytoplasm

(Webber, et al. , 2003). Small hydrophilic porin channels that are located in

the OM layer also inhibit the entrance of large hydrophobic molecules such

as erythromycin (Delcour, 2003; Pagès, et al., 2008). This is due to the

existence of an eyelet structure (see Figure 6-1) in porins that provides size

exclusion limits to the entrance of those big molecules (Delcour, 2009).

Since most of the strains used in this study were resistant to almost all

available antibiotics including those antibiotics that are commonly used as

selective markers (see Table 2-3) during cloning, we encountered another

problem to choose a suitable selective marker for the mutagenesis. The initial

mutant constructs were modified by replacing the erm1(B) with aad9. aad9

which encodes for spectinomycin adenyltransferase was first identified in

E. faecalis LDR55 plasmid, pDL412 (LeBlanc, et al. , 1991). This gene was

chosen but with a limitation that only three strains were subjected to be

manipulated, Kpneu_1, Kpneu_110, and C3091, since they had a lower MIC

of spectinomycin (12.5 µg/ml) than the MIC could be used for selection

(100 µg/ml). These strains were also confirmed to not carry aad9 by PCR.

The second reason of the failure to obtain mutant strains was thought to be

due to a host native RecBCD exonuclease pathway in K. pneumoniae

292 |

(C. Chen, et al. , 2016). This pathway is comprised of heterotrimer RecBCD

enzyme, consisting of three distinctly different polypeptides, RecB, RecC and

RecD. This RecBCD complex binds to the end of double-stranded DNA.

Operating from a 3’-5’ translocation polarity at the end of the double stranded

DNA, RecBCD unwinds the duplex by its fast and weak helicases, RecD and

RecB, respectively. Upon recognition of a cross-over hotspot instigator (Chi)

(5’ GCTGGTGG 3’) site on one strand by RecC, the enzyme complex cleaves

the other complementary strand, and exposes the new 3’ single strands end

with the Chi site. The recombinase protein (RecA) is then nucleated onto the

ssDNA to form a nucleoprotein filament. The RecA nucleoprotein filament

scavenges for a homologous strand from the host genome and catalyses the

DNA strand invasion by pairing the invading ssDNA with the host

complementary strand, and displacing the non-complementary strand by

formation of a Holliday junction (Amundsen, et al. , 2007; Chen, et al., 2016;

Dillingham, et al. , 2008). After this process, the RecBCD complex is then

dissembled.

Apart from recombinase, this potent exonuclease is also responsible for rapid

degradation of invading linear DNA (Wilkinson, et al. , 2016). This RecBCD

system directly targets foreign DNA elements that invade into host cells, and

without Chi site, degradation would possibly occur continuously until the whole

duplex DNA strand is digested. This might be happening in our

K. pneumoniae strains resulting in no mutant strains being obtained at first.

An alternative approach was then taken to overcome the RecBCD pathway in

K. pneumoniae strains, which was to increase the recombination efficiency of

linear foreign DNA into the bacterial chromosome.

293 |

Taken together, these approaches allowed us to obtain the ompK35 and

∆ompK36 mutant strains derived from Kpneu_1 and C3091 by reconstructing

new mutant constructs and adapting the recombineering method by λ-red

recombinase using a helper plasmid, pKD46. Exo, Bet and Gam, which are

encoded by exo, bet and gam derived from the red operon of λ-bacteriophage,

also play a significant role in hindering the degradation of our mutant cassettes

by RecBCD pathway upon electroporation and indirectly mediates the

homologous recombination to occur.

In this study, we could not manage to develop a mutant strain for Kpneu_110.

Upon visualization, colonies of this strain had a very distinct

hypermucoviscous phenotype (see Figure 1-2 in Chapter 1) compared to the

other two strains. During preparation of electrocompetent cells of

Kpneu_110, centrifugation and washing step were slightly inefficient since the

cells were difficult to be pelleted down. Our observation similar with a

previous study by Pan, et al. (2011) which mentioned the constraints of using

hypermucoviscous K. pneumoniae strains. This strain, however, was able to

take up the pKD46-apra helper plasmid since it was shown to harbour the

pKD46-apra by amplification of the apramycin resistance gene, aac(3)-IV and

pKD46-apra plasmid backbone (data not shown), but, no homologous

recombination between the target gene, ompK35 or ompK36 with selectable

marker aad9 was observed upon electroporation. The decreased efficiency

of homologous recombination in this strain might be because of the

physicochemical hindrance conferred by thick capsule structure and this

remains to be explored.

294 |

No complemented strains were obtained in this study. Our result might

suggest that usage of a high copy number plasmid such as pUC19 was

unsuitable for development of ompK35 or ompK36 complemented strains

since this may result in high levels of OmpK35 or OmpK36 that might

destabilise the outer membrane. Also, we hypothesised that the host cells

could not survive since abundant production of porins can be detrimental to

the growth of cells (K. L. Jones, et al. , 2000) and lead to porous cells.

Therefore, to develop complemented strains of ompK35 or ompK36, these

genes should be expressed in a low or single copy number plasmid. A low

copy number plasmid could be used instead of high copy number plasmid

(Sugawara, et al., 2016).

In addition, we also could not manage to obtain the complementation

constructs. As shown in Figure 6-18 and Figure 6-19, we managed to get the

putative mutant constructs with the expected size, however, they did not

harbour the complementation cassettes after confirmation by sequencing.

This could be because of a high degree similarity of ompK35 or ompK36 of

K. pneumoniae with ompF and ompC of E. coli which leads to homologous

recombination that could possibly occur. The fact that most of E. coli cloning

hosts are recA deficient to minimise any unexpected homologous

recombination, this homologous recombination is very unlikely to occur but it

is also important to consider this factor since there is still a chance it will

happen. Using a different cloning host as an example, Bacillus subtilis and a

low copy number plasmid, pHCMC05 or a copy control plasmid as in

pCC1BAC would possibly increase the efficiency, but further study is needed

to confirm this hypothesis.

295 |

The effect of OmpK35 and OmpK36 porin loss on antibiotic and antiseptic

resistance was determined using E-test and broth microdilution assays,

respectively. Loss of OmpK36 was shown to have an effect on median MIC

of cefotaxime for strain C3091. This can be seen from our data, which

showed that C3091ompK36, the mutant strain derived from laboratory

strain, had 2.5 times higher median MIC of cefotaxime than for C3091 and

C3091ompK35. The reason is possible due to the lack of channels that can

be used to transport the cefotaxime into the bacterial cell, however, this

hypothesis needs to be confirmed together with the complementation strains.

This data is in contrast to the previous findings by Tsai, et al. (2011) which

showed that loss of one or both of OmpK35 or OmpK36 porin did not affect

the susceptibility of K. pneumoniae strain NVT2001 to cefotaxime. The

median MIC of Kpneu_1 and its isogenic mutant strains, however, could not

be confirmed since it was higher than the detection limit of cefotaxime E-test

strip.

Another prominent difference that could be seen was the increased median

MIC of tetracycline for C3091ompK35 in comparison to its parental strain

and C3091ompK36. The data shows that Kpneu_1ompK35 also had a

higher median MIC of tetracycline than Kpneu_1 and Kpneu_1ompK36.

This suggested that OmpK35 might play a role for tetracycline transportation

across the membrane for both of these strains. Our data is in accordance

with the previous finding by Thanassi, et al. (1995) who also proved that

uptake of tetracycline was reduced in an E. coli ompF mutant, and increased

resistance in mutants with decreased expression of ompF. These finding

296 |

together with our finding clearly suggesting that tetracycline might use

OmpK35 or its homolog, OmpF as a channel to cross the outer membrane

layer of E. coli and K. pneumoniae. Since we did not have the complemented

strains, further study is warranted to confirm this hypothesis.

No large differences could be seen between Kpneu_1 and C3091 and their

isogenic mutants for the median MIC of ciprofloxacin and gentamicin. The

median MIC of ciprofloxacin was the same for Kpneu_1 and its isogenic

mutants, while C3091ompK35 had just only one dilution step of higher of

median MIC of ciprofloxacin (from 0.012 µg/ml to 0.016 µg/ml) compared to

its parental strain, C3091 and C3091ompK36. This is in agreement with the

finding by Hernández-Allés, et al. (2000) who found that deficiency in either

OmpK35 or OmpK36 porin expression did not result in increased MICs of

fluoroquinolones including the hydrophilic compounds, ciprofloxacin, and

gentamicin. A previous study also showed decreased MIC of ciprofloxacin

caused by the expression of OmpK35, was just one dilution step lower

(0.25 µg/ml) than the strain that expressed only OmpK36, and strain that

expressed both of the porins which the MIC was at 0.5 µg/ml (Doménech-

Sánchez, et al., 2003).

We, however, observed that C3091∆ompK35 and C3091∆ompK36 had

increased susceptibility to gentamicin compared to C3091 strain. The median

MIC of both of the mutant strains was found to be 1.5 times lower than their

wild type. In addition, Kpneu_1∆ompK36 was also found to have lower

median MIC to gentamicin compared to its wild type and Kpneu_1∆ompK35.

This observation is not in line with the previous study by Doménech-Sánchez,

297 |

et al. (2003) which showed that lack of OmpK35 or OmpK36 did not result in

any difference in MIC of gentamicin (4 µg/ml) in comparison to the strain that

expressed both of the porins.

Hernández-Allés, et al. (2000) also concluded that loss of expression of a

single porin either OmpK35 or OmpK36 or both porins did not alter the MICs

of aminoglycosides, fluoroquinolones, tetracycline, and chloramphenicol but

the MICs of β-lactams including cefotaxime were observed to increase for

ompK35/ompK36 porin-deficient mutants. Nonetheless, this is not the case

in our study since we found that absence of OmpK35 porin had a difference

in median MIC of tetracycline and loss of OmpK36 porin of C3091 showed

about 2.5 times higher median MIC of cefotaxime compared to its parental

strain and C3091ompK36. The role of porins in antibiotic susceptibility is

well-established. However, the variance of antibiotic susceptibility between

strains could also be due to the existence of other mechanisms of resistance

as an example, having efflux pumps or harbouring any resistance genes

against several classes of antibiotics in the same bacterial cells.

Due to the fact that OmpF and OmpC are charge preference, which prefer

neutral or cationic molecules over anionic molecules, we then hypothesised

that loss of OmpK35 and OmpK36, the homologs of OmpF and OmpC, might

affect MIC of CTAB and CPC against K. pneumoniae (Nikaido, 1994). QACs

are monocationic surfactants, which reduce the surface tension (interfacial

tension) of liquid, or between liquid and solid and form micelles. These

micelles can be easily dispersed in liquid. The following steps of the QACs

mechanisms against bacteria were proposed by McDonnell (2007):

298 |

i) absorption of QACs via the cell wall ii) interaction of QACs with the bacterial

cytoplasmic membrane (phospholipid and LPS) followed by disorientation of

the bacterial membrane iii) released of intracellular lower-weight compounds

iv) degradation of the proteins and DNA v) induction of cell autolysis by

autolytic enzymes (Denyer, et al. , 1998; Gerba, 2015; Maillard, 2002;

McDonnell, et al. , 1999). Data from our investigation, however, showed that

lack of porins, OmpK35 and OmpK36 did not affect the median MIC of CTAB

and CPC. This suggests that OmpK35 and OmpK36 may not be involved in

the entrance of these antiseptics into K. pneumoniae.

Since there was an effect on growth of the ΔompK35 and ΔompK36 (Figure

6-20 and Figure 6-21) mutants compared to wild types in TSB and AUM, it is

suggested that porins could possibly play a role in growth of

K. pneumoniae under high-osmolarity environment. Our observation,

however, is not in agreement with a previous study by Tsai, et al. (2011). This

study showed that wild type strain of NVT2001S had similar growth rate with

its ompK35 and ompK36 mutant strains which were grown in LB and their

generation time was 27 min for each strain. Our data showed similar growth

rates were observed between Kpneu_1 and C3091 and their ∆ompK35 mutant

strains upon growth in TSB with a doubling time of 35 min. Kpneu_1∆ompK36

and C3091∆ompK36 were observed to grow faster in TSB with doubling times

of 29 min and 32 min respectively. All the mutant strains also were observed

to have shorter doubling time upon cultivation in AUM in comparison to their

wild type strains. Thus, our data suggests that loss of OmpK35 does not alter

growth of K. pneumoniae, but loss of OmpK36 may have a very small effect

on the generation time upon growth in TSB. Whilst, loss of OmpK35 and

299 |

OmpK36 could possibly have an effect in generation time upon cultivation in

the nutrient-poor medium, AUM.

No significant difference was observed for biofilm formation by Kpneu_1

compared to its isogenic mutant strains in either of the media tested. In

addition, the same observation was obtained for biofilm formation by C3091

compared to C3091∆ompK35 upon cultivation in 96-well microtitre plates in

TSB. However, a small but significant increase could be seen for biofilm

formation by C3091∆ompK36 compared to C3091. Given the proposed role

of OmpC in biofilm formation by E. coli, we were quite surprised to observe

that loss of OmpK36 slightly increased the biofilm formation by

C3091∆ompK36. Our data suggests that OmpK36 possibly has a small but

significant effect on biofilm formation by K. pneumoniae upon cultivation in

AUM in 96-well microtitre plate.

We then performed another biofilm formation assays to investigate the role of

OmpK35 and OmpK36 in biofilm formation by Kpneu_1, C3091 and its

isogenic mutants in more clinically relevant assay. No significant difference

could be seen for cells number and also biofilm mass produced by Kpneu_1,

C3091 and their isogenic mutants upon cultivation in TSB and AUM,

respectively. This assay suggested that OmpK35 and OmpK36 do not

contribute to the biofilm formation by the tested K. pneumoniae strains upon

growth in TSB and AUM on catheter pieces.

300 |

6.5 Conclusion

In this chapter, we have shown that the amino acid sequences and protein

structure of OmpK36 are highly similar with OmpC of E. coli. In addition,

comparative structural alignment of OmpK36 and a putative structure of

OmpK35 shows high degree of similarity, despite sharing 60 % identity at the

amino acid sequence level.

Our data also shows that erm1(B) is not a suitable selective marker for

K. pneumoniae and the genetic manipulation in K. pneumoniae requires a

helper plasmid, pKD46 since it is likely that the RecBCD machinery would

affect the occurrence of allelic exchange. This is because this helper plasmid

harbours exo, beta, and gam which encode for the protein that facilitate the

occurrence of homologous recombination of the foreign DNA with

K. pneumoniae genome.

In addition, our findings have suggested that OmpK35 might possibly play a

role for transportation of tetracycline in strains Kpneu_1 and C3091, while

OmpK36 was shown to be involved in increased MIC of cefotaxime for

strain C3091ompK36. We have also demonstrated that both of these porins

do not affect the susceptibility of Kpneu_1 and C3091 to QACs such as CTAB

and CPC.

The biofilm formation assays in 96-well microtitre plates have shown that

OmpK35 and OmpK36 do not play a role in biofilm formation by Kpneu_1

upon cultivation in TSB and AUM in 96-well microtitre plates with the

exception of a slight increase in biofilm formation can be seen for

C3091∆ompK36 compared to C3091. Another clinically relevant biofilm

301 |

formation assays on silicone catheter pieces has also shown that OmpK35

and OmpK36 are not involved in the biofilm formation by Kpneu_1 and C3091

on catheter pieces upon growth in TSB and AUM, respectively.

302 |

Chapter 7.0

General discussion and

future directions

303 |

7.0 General discussion and future directions

7.1 General discussion

The rapid emergence of antibiotic resistant bacteria has become a serious

threat to global public health. It is estimated that by 2050, there will be almost

10 million individuals who will die due to antibiotic resistant bacterial infections

(O'Neill, 2016). K. pneumoniae is known as one of the most prevalent species

that causes antibiotic resistant nosocomial infections, and as reported by

O'Neill (2016), K. pneumoniae has been observed to increase in resistance

rapidly. Apart from the acquisition of antibiotic resistance genes, the

emergence of antibiotic resistant K. pneumoniae strains, that could resist

most of the antibiotics available in clinical settings, is probably due to the

capacity of this pathogen to form biofilms on medical devices, for example,

urinary catheters. This phenomenon is known as biofilm-mediated antibiotic

resistance (Vuotto, et al., 2014).

Despite the large amount of research on mechanisms of antibiotic resistance

by K. pneumoniae, very little information exists about the detailed mechanisms

of biofilm formation by this pathogen. Taking this challenge as our research

theme, this study aimed to investigate the mechanisms of biofilm formation by

K. pneumoniae as this information could aid the development of therapeutic

agents against K. pneumoniae infection, and subsequently could provide

further understanding on the contributing factors of biofilm-mediated antibiotic

resistance by K. pneumoniae. Five multidrug resistant clinical isolates and two

laboratory strains, which acted as biofilm producer controls, were used in this

study. The studies presented in this thesis mainly focused on: i) the effect of

304 |

exposure of sub-inhibitory concentrations of antibiotics on biofilm formation by

K. pneumoniae ii) the composition of biofilm EPS of K. pneumoniae upon

cultivation in two different media, TSB and AUM iii) the role of type 1 fimbriae

(T1F) in biofilm formation by K. pneumoniae, and iv) the role of OmpK35 and

OmpK36 in biofilm formation by K. pneumoniae.

The use of antibiotics (both appropriate and inappropriate) will expose bacteria

to low concentrations of antibiotics. This study was conducted as several

previous studies have shown that exposure of bacterial culture to sub-

inhibitory concentrations of antibiotics increased the biofilm formation by

clinically relevant pathogen, for example, P. aeruginosa. A study by Bagge, et

al. (2004) showed that exposure of P. aeruginosa biofilms to sub-inhibitory

concentrations of antibiotics, for example, 0.2 µg/ml of imipenem, increased

the biofilm formation by P. aeruginosa by upregulating the genes which encode

for alginate biosynthesis.

As previously discussed, the capacity to form biofilms by K. pneumoniae is

established as one of the causative factors contributing to the emergence of

antibiotics resistant K. pneumoniae strains. Thus, the aim of the study

reported in Chapter 3 was to investigate the effect of exposure to sub-MICs of

antibiotics, gentamicin and ciprofloxacin, on biofilm formation by

K. pneumoniae. We found that the biofilm formation by all the

K. pneumoniae strains used in this study was not affected by the sub-MICs of

commonly found antibiotics, gentamicin, and ciprofloxacin in clinical settings.

305 |

To our knowledge, we have provided for the first time data on the effect of

antibiotic exposure on biofilm formation by K. pneumoniae. These results have

ruled out the risk factors for the antibiotic-induced biofilm formation by

K. pneumoniae particularly on the multidrug resistant strains in clinical settings.

The findings are clinically relevant since K. pneumoniae is also amongst the

prevalent species that causes nosocomial infections, and the antibiotics used

in this study are amongst the most common antibiotics prescribed for infections

with this bacterium, particularly gentamicin and ciprofloxacin (Enzler, et al.,

2011; Hickerson, et al., 2006; Parkinson, et al., 2016).

In the Chapter 4, we investigated the composition of K. pneumoniae biofilm

EPS since it is not very well understood. Our findings show a positive effect

of proteinase K on biofilm formation by Kpneu_110. Biofilm of Kpneu_110

was found to increase upon treatment with 100 µg/ml of proteinase K in TSB.

Since Kpneu_110 has very distinct morphological characteristics compared

to the rest of the strains (see Figure 1-2), it would be worthwhile to investigate

this observation in the future as this kind of information may be very important

in identifying the contributing factors that could enhance the biofilm formation

by hypermucoviscous K. pneumoniae strains instead of disrupting their

biofilms. Further work is needed to determine if there is any link between the

increased biofilms formation of hypermucoviscous of K. pneumoniae strains

upon treatment with proteinase K.

306 |

We also report for the first time on the analysis of the orientation of fim-switch;

fimS, in multiple antibiotic resistant strain of K. pneumoniae upon cultivation in

two different media in Chapter 5. fimS is an invertible DNA element that

controls the expression of major subunit of type 1 fimbriae (T1F), fimA and

subsequently leads to the expression of T1F. We also showed four of the

strains used in this study had biofilm cells with fimS being in the ‘ON and OFF’

orientation (Kpneu_1, Kpneu_65, C3091, and CG43S3) and three strains

(Kpneu_63, Kpneu_64, and Kpneu_110) had biofilm cells with fimS being in

the ‘OFF’ orientation upon cultivation in TSB. All the strains, however, had

biofilm cells with fimS in the ‘OFF’ orientation upon cultivation in AUM. This

suggested that T1F could be involved in biofilm formation of four of the

examined K. pneumoniae strains used in this study upon cultivation in TSB,

however, the role of T1F in biofilm formation upon cultivation in AUM is still

unclear.

Our investigation on the gene sequences in T1F operon revealed multiple

SNPs that were specific for both of the ‘ON and OFF’ and ‘OFF’ groups.

Interestingly, one of the SNPs resulted in truncated FimK and it is conserved

in all strains which had the fimS being in the ‘OFF’ orientation. We

hypothesised that the truncated FimK could influence the orientation of fimS

since it was only present in strains where the fimS was always in the ‘OFF’

orientation, but further study is warranted to confirm this hypothesis. Upon

analysis of DNA sequences of genes in the T1F operon of all available

K. pneumoniae strains on GenBank (up until August 2016), it revealed out of

56 strains, 22 strains were shown to have this particular SNP that resulted in

a truncated FimK. Most of the strains were identified as clinical isolates from

307 |

different samples such as blood, urine, perirectal swabs and sputum.

However, not all of the genome sequences of clinical K. pneumoniae strains

used in this analysis were identified to have this particular SNP.

We then investigated other proteins that might be involved in biofilm formation

by K. pneumoniae, which were OmpK35 and OmpK36. In the final results

chapter (Chapter 6), defined isogenic mutants of strain Kpneu_1 and C3091,

were constructed by deleting ompK35, and ompK36 and replacing these

genes with spectinomycin resistance gene, aad9. We also tested the capacity

of ompK35 and ompK36 mutant strains to form biofilm in 96-well microtitre

plates and on catheter surfaces together with their wild type strains. Our study

suggested that OmpK35 contributes to increase resistance to tetracycline,

while OmpK36 plays a role in increased resistance to cefotaxime for strains

C3091∆ompK35 and C3091∆ompK36, respectively. This, however, could not

be confirmed since we do not have complemented strains. Due to the fact that

both of these antibiotics, tetracycline and cefotaxime, at sub-inhibitory

concentrations could induce biofilm formation by other Gram-negative bacteria

such as S. enterica ser. Infantis (Tezel, et al., 2016) and S. enterica ser.

Typhimurium (Majtan, et al. , 2008), we hypothesised that, by loss of OmpK35

or OmpK36, K. pneumoniae could possibly form more biofilm upon exposure

to the sub-inhibitory concentration of tetracycline and cefotaxime and this

particular phenomenon is quite difficult to manage. However, further study is

warranted to confirm this hypothesis.

308 |

OmpK35 and OmpK36 were shown to play no role in biofilm formation by

K. pneumoniae in our experimental model, apart from the slight increase of

biofilm formation in C3091∆ompK36 in comparison to control upon cultivation

in 96-well microtitre plates. Due to the fact that K. pneumoniae could adhere

and invade human bladder in pathogenesis of UTIs (Oelschlaeger, et al., 1997;

Sahly, et al. , 2008), it is worth investigating the role of OmpK35 and OmpK36

in mechanisms of adhesion and invasion of K. pneumoniae to human epithelial

cells such as T24 human bladder carcinoma cells.

To conclude, this study extends the current knowledge about mechanisms of

biofilm formation by K. pneumoniae. Exposure to sub-MICs of gentamicin, and

ciprofloxacin do not increase biofilm formation by

K. pneumoniae. We have shown that proteins might play a role in biofilm

formation by K. pneumoniae. Further investigation on the role of one of the

surface proteins, T1F, suggested that role of T1F in biofilm formation by

K. pneumoniae is still speculative, since it has a different role upon cultivation

in different environments. Our study also suggested that OmpK35 and

OmpK36 contributed to increased resistance to tetracycline and cefotaxime

but this needs to be confirmed with the complemented strains. In addition,

OmpK35 and OmpK36 do not play a role in biofilm formation by K. pneumoniae

upon cultivation on clinically relevant surfaces, on catheter surfaces.

It is hoped that this research, which expands the current knowledge and

understanding of mechanisms of biofilm formation by K. pneumoniae could

contribute to future development of effective treatments for biofilm-associated

infection caused by K. pneumoniae.

309 |

7.2 Future directions

Exposure to sub-inhibitory concentrations of gentamicin and ciprofloxacin was

shown to have no effect on biofilm formation by

K. pneumoniae. Since sub-inhibitory concentrations of antibiotics are

suggested to stimulate or repress a large number of gene expressions in target

bacteria at transcription level, (Goh, et al. , 2002), it is worth investigating the

effect of other antibiotics on biofilm formation by K. pneumoniae. This

investigation could possibly provide us more information on risk factors that

could contribute to the antibiotic-induced biofilm formation by K. pneumoniae.

Since the resistance of K. pneumoniae to antibiotics is established to arise

rapidly, and one of the known causes is the capacity to form biofilm by this

pathogen, these data would be very beneficial in highlighting the underlying

effect of the antibiotic exposure on biofilm formation by

K. pneumoniae. If we could identify the antibiotics that could induce biofilm

formation by K. pneumoniae, it is useful to test the clinical isolates from the

patients for biofilm inducibility against the antibiotics before a prescription of

antibiotics in order to control the antibiotic-induced biofilm formation by this

pathogen.

We also found that Kpneu_110 had a very low capacity to form biofilms in

AUM. Knowing this hypermucosviscous strain was isolated from UTI, this is

something worth further investigation. As hypermucoviscous K. pneumoniae

strains are normally associated with hypervirulence, it could be very beneficial

if we could obtain more information about the gene regulation in biofilm

formation by this hypermucosviscous strain (Shon, et al., 2013). Gene

310 |

expression study by comparing between planktonic cells and biofilms of

Kpneu_110 upon cultivation in AUM could be a very good stepping stone for

us to understand the differential gene expression of this particular strain in a

different growth state cultivated in a clinically relevant medium. In addition, it

is also worth comparing the differential gene expression of biofilm cells of

Kpneu_110 and biofilm cells of Kpneu_65 for example, since Kpneu_65 was

observed to form more biofilm in AUM compared to the rest of the strains. By

having this comparison, we could obtain an important information on what

genes are involved in biofilm formation by K. pneumoniae in nutrient-poor

medium. The study on differential gene expressions of biofilm cells upon

growth in clinically relevant medium is utmost important in order for us to

identify the drug targets that could be useful for future development for

treatment of the biofilm-associated infections caused by

K. pneumoniae particularly by hypermucoviscous strains.

Moreover, since biofilm formation of Kpneu_110 was found to be induced by

proteinase K, further investigation into this observation could also be

advantageous. The positive effect on biofilm formation by proteinase K could

provide us some interesting data regarding a putative self-secretion dispersion

enzyme produces by strain Kpneu_110 that could be useful in controlling the

biofilm formation by other bacterial. This putative enzyme might be further

developed to be used as a potential candidate agent to treat biofilm-associated

infections. One of the classic example of this kind of enzyme was previously

identified. Dispersin B, or DspB that produced by A. actinomycetemcomitans

is established to degrade the glycosidic bonds of β-1, 6-N-acetylglucosamine

or PNAG (Ramasubbu, et al., 2005).

311 |

In addition, the SNPs found upon analysis of the gene sequences in T1F

operon could also be further investigated. Particularly the effect of truncation

of FimK on the orientation of fimS in Kpneu_63, Kpneu_64, and Kpneu_110,

since FimK is suggested to play a role in the orientation of fimS. If we could

confirm the role of FimK in the orientation of fimS which controls the expression

of T1F in K. pneumoniae, it could provide us a better understanding on the

regulation of T1F and subsequently determine the role of T1F in biofilm cells

of K. pneumoniae. The exploration of this particular virulence factor could lead

us to a better way to counteract the biofilm-associated K. pneumoniae

infections by using T1F as a drug target.

312 |

Chapter 8.0

References

313 |

8.0 References

Abbott, S. L. (2007). Klebsiella, Enterobacter, Citrobacter, and Serratia. In P.

R. Murray (Ed.), Manual of Clinical Microbiology (9th ed.). Washington

DC: ASM Press.

Abdallah, H. M., Wintermans, B. B., Reuland, E. A., et al. (2015). Extended-

spectrum β-lactamase- and carbapenemase-producing

Enterobacteriaceae isolated from Egyptian patients with suspected

blood stream infection. PLOS ONE, 10(5), e0128120.

Ahmad, I., Rouf, S. F., Sun, L., et al. (2016). BcsZ inhibits biofilm phenotypes

and promotes virulence by blocking cellulose production in Salmonella

enterica serovar Typhimurium. Microbial Cell Factories, 15(1), 177.

Ahmed, N. A., Petersen, F. C., & Scheie, A. A. (2009). AI-2/LuxS is involved

in increased biofilm formation by Streptococcus intermedius in the

presence of antibiotics. Antimicrobial Agents and Chemotherapy,

53(10), 4258-4263.

Aka, S. T., & Haji, S. H. (2015). Sub-MIC of antibiotics induced biofilm

formation of Pseudomonas aeruginosa in the presence of

chlorhexidine. Brazilian Journal of Microbiology, 46(1), 149-154.

Albertí, S., Marques, G., Hernández-Allés, S., et al. (1996). Interaction

between complement subcomponent C1q and the Klebsiella

pneumoniae porin OmpK36. Infection and Immunity, 64(11), 4719-

4725.

Albertí, S., Rodríquez-Quiñones, F., Schirmer, T., et al. (1995). A porin from

Klebsiella pneumoniae: sequence homology, three-dimensional model,

and complement binding. Infection and Immunity, 63(3), 903-910.

314 |

Albiger, B., Glasner, C., Struelens, M. J., et al. (2015). Carbapenemase-

producing Enterobacteriaceae in Europe: assessment by national

experts from 38 countries, May 2015. Eurosurveillance, 20(45), 30062.

Alcántar-Curiel, M. D., Blackburn, D., Saldaña, Z., et al. (2013). Multi-

functional analysis of Klebsiella pneumoniae fimbrial types in

adherence and biofilm formation. Virulence, 4(2), 129-138.

Allen, B. L., Gerlach, G. F., & Clegg, S. (1991). Nucleotide sequence and

functions of mrk determinants necessary for expression of type 3

fimbriae in Klebsiella pneumoniae. Journal of Bacteriology, 173(2), 916-

920.

Allen, H. K., Donato, J., Wang, H. H., et al. (2010). Call of the wild: antibiotic

resistance genes in natural environments. Nature Reviews

Microbiology, 8(4), 251-259.

Almaghrabi, R., Clancy, C. J., Doi, Y., et al. (2014). Carbapenem-resistant

Klebsiella pneumoniae strains exhibit diversity in aminoglycoside-

modifying enzymes, which exert differing effects on plazomicin and

other agents. Antimicrobial Agents and Chemotherapy, 58(8), 4443-

4451.

Alonso, A., Sanchez, P., & Martinez, J. L. (2001). Environmental selection of

antibiotic resistance genes. Environmental Microbiology, 3(1), 1-9.

Altschul, S. F., Gish, W., Miller, W., Myers, E. W., & Lipman, D. J. (1990). Basic

local alignment search tool. Journal of Molecular Cell Biology, 215(3),

403-410.

Amako, K., Meno, Y., & Takade, A. (1988). Fine structures of the capsules of

Klebsiella pneumoniae and Escherichia coli K1. Journal of Bacteriology,

170(10), 4960-4962.

315 |

Amalaradjou, M. A. R., & Venkitanarayanan, K. (2011). Natural approaches

for controlling urinary tract infections. In P. Tenke (Ed.), Urinary Tract

Infections: InTech.

Amalaradjou, M. A. R., & Venkitanarayanan, K. (2013). Role of bacterial

biofilms in catheter-associated urinary tract infections (CAUTI) and

strategies for their control. In T. Nelius (Ed.), Recent Advances in the

Field of Urinary Tract Infections. Rijeka: InTech.

Ammor, M. S., Flórez, A. B., Van Hoek, A. H., et al. (2007). Molecular

characterization of intrinsic and acquired antibiotic resistance in lactic

acid bacteria and bifidobacteria. Journal of Molecular Microbiology &

Biotechnology, 14(1-3), 6-15.

Amundsen, S. K., Taylor, A. F., Reddy, M., & Smith, G. R. (2007). Intersubunit

signaling in RecBCD enzyme, a complex protein machine regulated by

Chi hot spots. Genes & Development, 21(24), 3296-3307.

Andersson, D. I., & Hughes, D. (2014). Microbiological effects of sublethal

levels of antibiotics. Nature Reviews Microbiology, 12(7), 465-478.

Arnold, R. S., Thom, K. A., Sharma, S., et al. (2011). Emergence of Klebsiella

pneumoniae carbapenemase (KPC)-producing bacteria. Southern

Medical Journal, 104(1), 40-45.

Athamna, A., Ofek, I., Keisari, Y., et al. (1991). Lectinophagocytosis of

encapsulated Klebsiella pneumoniae mediated by surface lectins of

guinea pig alveolar macrophages and human monocyte-derived

macrophages. Infection and Immunity, 59(5), 1673-1682.

Azeredo, J., Azevedo, N. F., Briandet, R., et al. (2017). Critical review on

biofilm methods. Critical Review in Microbiology, 43(3), 313-351.

316 |

Aziz, R. K., Bartels, D., Best, A. A., et al. (2008). The RAST Server: Rapid

Annotations using Subsystems Technology. BMC Genomics, 9(1), 75.

Bagge, N., Schuster, M., Hentzer, M., et al. (2004). Pseudomonas aeruginosa

biofilms exposed to imipenem exhibit changes in global gene

expression and β-lactamase and alginate production. Antimicrobial

Agents and Chemotherapy, 48(4), 1175-1187.

Bagley, S. T. (1985). Habitat association of Klebsiella species. Infection

Control & Hospital Epidemiology, 6(2), 52-58.

Bagley, S. T., Seidler, R. J., & Brenner, D. J. (1981). Klebsiella planticola sp.

nov.: A new species of Enterobacteriaceae found primarily in nonclinical

environments. Current Microbiology, 6(2), 105-109.

Balasubramaniam, D., Arockiasamy, A., Kumar, P. D., Sharma, A., &

Krishnaswamy, S. (2012). Asymmetric pore occupancy in crystal

structure of OmpF porin from Salmonella typhi. Journal of Structural

Biology, 178(3), 233-244.

Bales, P. M., Renke, E. M., May, S. L., Shen, Y., & Nelson, D. C. (2013).

Purification and characterization of biofilm-associated EPS

exopolysaccharides from ESKAPE organisms and other pathogens.

PLOS ONE, 8(6), e67950.

Balestrino, D., Ghigo, J. M., Charbonnel, N., Haagensen, J. A., & Forestier, C.

(2008). The characterization of functions involved in the establishment

and maturation of Klebsiella pneumoniae in vitro biofilm reveals dual

roles for surface exopolysaccharides. Environmental Microbiology,

10(3), 685-701.

Bandeira, M., Borges, V., Gomes, J. P., Duarte, A., & Jordao, L. (2017).

Insights on Klebsiella pneumoniae biofilms assembled on different

317 |

surfaces using phenotypic and genotypic approaches. Microorganisms,

5(2), 16.

Baquero, F., Martínez, J., & Cantón, R. (2008). Antibiotics and antibiotic

resistance in water environments. Current Opinion in Biotechnology,

19(3), 260-265.

Baquero, F., Negri, M. C., Morosini, M. I., & Blázquez, J. (1998). Antibiotic-

selective environments. Clinical Infectious Diseases, 27 (Supplement

1), S5-11.

Barnhart, M. M., & Chapman, M. R. (2006). Curli biogenesis and function.

Annual Review of Microbiology, 60, 131-147.

Barthelemy, M., Peduzzi, J., Ben Yaghlane, H., & Labia, R. (1988). Single

amino acid substitution between SHV-1 beta-lactamase and

cefotaxime-hydrolyzing SHV-2 enzyme. FEBS Letters, 231(1), 217-

220.

Bascomb, S., Lapage, S. P., Willcox, W. R., & Curtis, M. A. (1971). Numerical

classification of the tribe Klebsielleae. Microbiology, 66(3), 279-295.

Bauernfeind, A., Schweighart, S., & Grimm, H. (1990). A new plasmidic

cefotaximase in a clinical isolate of Escherichia coli. Infection, 18(5),

294-298.

Bauernfeind, A., Stemplinger, I., Jungwirth, R., & Giamarellou, H. (1996).

Characterization of the plasmidic beta-lactamase CMY-2, which is

responsible for cephamycin resistance. Antimicrobial Agents and

Chemotherapy, 40(1), 221-224.

318 |

Benenson, S., Navon-Venezia, S., Carmeli, Y., et al. (2009). Carbapenem-

resistant Klebsiella pneumoniae endocarditis in a young adult:

Successful treatment with gentamicin and colistin. International Journal

of Infectious Diseases, 13(5), e295-298.

Bennett, P. M. (2008). Plasmid encoded antibiotic resistance: acquisition and

transfer of antibiotic resistance genes in bacteria. British Journal of

Pharmacology, 153(Supplement 1), S347-357.

Berendonk, T. U., Manaia, C. M., Merlin, C., et al. (2015). Tackling antibiotic

resistance: the environmental framework. Nature Reviews

Microbiology, 13(5), 310-317.

Berman, H. M., Westbrook, J., Feng, Z., et al. (2000). The Protein Data Bank.

Nucleic Acids Research, 28(1), 235-242.

Berne, C., Ducret, A., Hardy, G. G., & Brun, Y. V. (2015). Adhesins involved in

attachment to abiotic surfaces by Gram-negative bacteria. Microbiology

Spectrum, 3(4).

Bhavnani, S. M., Ambrose, P. G., Craig, W. A., Dudley, M. N., & Jones, R. N.

(2006). Outcomes evaluation of patients with ESBL- and non–ESBL-

producing Escherichia coli and Klebsiella species as defined by CLSI

reference methods: Report from the SENTRY Antimicrobial

Surveillance Program. Diagnostic Microbiology and Infectious Disease,

54(3), 231-236.

Biasini, M., Bienert, S., Waterhouse, A., et al. (2014). SWISS-MODEL:

modelling protein tertiary and quaternary structure using evolutionary

information. Nucleic Acids Research, 42(W1), W252-W258.

Bisognano, C., Vaudaux, P. E., Lew, D. P., Ng, E. Y., & Hooper, D. C. (1997).

Increased expression of fibronectin-binding proteins by

319 |

fluoroquinolone-resistant Staphylococcus aureus exposed to

subinhibitory levels of ciprofloxacin. Antimicrobial Agents and

Chemotherapy, 41(5), 906-913.

Blair, J. M., Webber, M. A., Baylay, A. J., Ogbolu, D. O., & Piddock, L. J.

(2015). Molecular mechanisms of antibiotic resistance. Nature Reviews

Microbiology, 13(1), 42-51.

Blickwede, M., Valentin-Weigand, P., & Schwarz, S. (2004). Subinhibitory

concentrations of florfenicol enhance the adherence of florfenicol-

susceptible and florfenicol-resistant Staphylococcus aureus. Journal of

Antimicrobial Chemotherapy, 54(1), 286-288.

Boddicker, J. D., Anderson, R. A., Jagnow, J., & Clegg, S. (2006). Signature-

tagged mutagenesis of Klebsiella pneumoniae to identify genes that

influence biofilm formation on extracellular matrix material. Infection

and Immunity, 74(8), 4590-4597.

Boehm, A., Steiner, S., Zaehringer, F., et al. (2009). Second messenger

signalling governs Escherichia coli biofilm induction upon ribosomal

stress. Molecular Microbiology, 72(6), 1500-1516.

Botelho-Nevers, E., Gouriet, F., Lepidi, H., et al. (2007). Chronic nasal

infection caused by Klebsiella rhinoscleromatis or Klebsiella ozaenae:

two forgotten infectious diseases. International Journal of Infectious

Diseases, 11(5), 423-429.

Bratu, S., Landman, D., Haag, R., & et al. (2005). Rapid spread of

carbapenem-resistant Klebsiella pneumoniae in New York City: a new

threat to our antibiotic armamentarium. Archives of Internal Medicine,

165(12), 1430-1435.

320 |

Bratu, S., Tolaney, P., Karumudi, U., et al. (2005). Carbapenemase-producing

Klebsiella pneumoniae in Brooklyn, NY: Molecular epidemiology and in

vitro activity of polymyxin B and other agents. Journal of Antimicrobial

Chemotherapy, 56(1), 128-132.

Bren, A., Park, J. O., Towbin, B. D., et al. (2016). Glucose becomes one of the

worst carbon sources for E.coli on poor nitrogen sources due to

suboptimal levels of cAMP. Scientific Reports, 6, 24834.

Brisse, S., Fevre, C., Passet, V., et al. (2009). Virulent clones of Klebsiella

pneumoniae: Identification and evolutionary scenario based on

genomic and phenotypic characterization. PLOS ONE, 4(3), e4982.

Brisse, S., Grimont, F., & Grimont, P. A. D. (2006). The Genus Klebsiella. In

M. Dworkin, S. Falkow, E. Rosenberg, K.-H. Schleifer & E.

Stackebrandt (Eds.), The Prokaryotes: Volume 6: Proteobacteria:

Gamma Subclass (pp. 159-196). New York, NY: Springer New York.

Brisse, S., Passet, V., Haugaard, A. B., et al. (2013). wzi gene sequencing, a

rapid method for determination of capsular type for Klebsiella strains.

Journal of Clinical Microbiology, 51(12), 4073-4078.

Brooks, T., & Keevil, C. W. (1997). A simple artificial urine for the growth of

urinary pathogens. Letters in Applied Micobiology, 24(3), 203-206.

Brown, C., & Seidler, R. J. (1973). Potential pathogens in the environment:

Klebsiella pneumoniae, a taxonomic and ecological enigma. Applied

Microbiology, 25(6), 900-904.

Buchanan, S. K., Smith, B. S., Venkatramani, L., et al. (1999). Crystal structure

of the outer membrane active transporter FepA from Escherichia coli.

Nature Structural & Molecular Biology, 6(1), 56-63.

321 |

Buckley, J., Coffin, S. E., Lautenbach, E., et al. (2007). Outcome of

Escherichia coli and/or Klebsiella bloodstream infection in children with

central venous catheters. Infection Control & Hospital Epidemiology,

28(11), 1308-1310.

Busch, A., & Waksman, G. (2012). Chaperone–usher pathways: diversity and

pilus assembly mechanism. Philosophical Transactions of the Royal

Society B: Biological Sciences, 367(1592), 1112-1122.

Bush, K., & Jacoby, G. A. (2010). Updated functional classification of β-

lactamases. Antimicrobial Agents and Chemotherapy, 54(3), 969-976.

Bush, K., Jacoby, G. A., & Medeiros, A. A. (1995). A functional classification

scheme for beta-lactamases and its correlation with molecular

structure. Antimicrobial Agents and Chemotherapy, 39(6), 1211-1233.

Cabeen, M. T., & Jacobs-Wagner, C. (2005). Bacterial cell shape. Nature

Reviews Microbiology, 3(8), 601-610.

Cabello, F. C. (2006). Heavy use of prophylactic antibiotics in aquaculture: a

growing problem for human and animal health and for the environment.

Environmental Microbiology, 8(7), 1137-1144.

Carter, J. S., Bowden, F. J., Bastian, I., et al. (1999). Phylogenetic evidence

for reclassification of Calymmatobacterium granulomatis as Klebsiella

granulomatis comb. nov. International Journal of Systematic

Bacteriology, 49 (Pt 4), 1695-1700.

Carugo, O., & Pongor, S. (2001). A normalized root-mean-square distance for

comparing protein three-dimensional structures. Protein Science, 10(7),

1470-1473.

322 |

Ceri, H., Olson, M. E., Stremick, C., et al. (1999). The Calgary biofilm device:

New technology for rapid determination of antibiotic susceptibilities of

bacterial biofilms. Journal of Clinical Microbiology, 37(6), 1771-1776.

Chaignon, P., Sadovskaya, I., Ragunah, C., et al. (2007). Susceptibility of

Staphylococcal biofilms to enzymatic treatments depends on their

chemical composition. Applied Microbiology and Biotechnology, 75(1),

125-132.

Chen, C., Wei, D., Liu, P., et al. (2016). Inhibition of RecBCD in Klebsiella

pneumoniae by Gam and its effect on the efficiency of gene

replacement. Journal of Basic Microbiology, 56(2), 120-126.

Chen, D., & van de Ven, T. G. (2016). Morphological changes of sterically

stabilized nanocrystalline cellulose after periodate oxidation. Cellulose,

23(2), 1051-1059.

Chen, F. J., Chan, C. H., Huang, Y. J., et al. (2011). Structural and mechanical

properties of Klebsiella pneumoniae type 3 fimbriae. Journal of

Bacteriology, 193(7), 1718-1725.

Chen, J. H., Siu, L. K., Fung, C. P., et al. (2010). Contribution of outer

membrane protein K36 to antimicrobial resistance and virulence in

Klebsiella pneumoniae. Antimicrobial Agents and Chemotherapy,

65(5), 986-990.

Chen, K. M., Chiang, M. K., Wang, M., et al. (2014). The role of pgaC in

Klebsiella pneumoniae virulence and biofilm formation. Microbial

Pathogenesis, 77, 89-99.

Chen, L., Todd, R., Kiehlbauch, J., Walters, M., & Kallen, A. (2017). Notes from

the field: pan-resistant New Delhi metallo-beta-lactamase-producing

323 |

Klebsiella pneumoniae — Washoe County, Nevada, 2016. Morbidity

and Mortality Weekly Report 2017, 66(33).

Chothia, C., & Lesk, A. M. (1986). The relation between the divergence of

sequence and structure in proteins. EMBO Journal, 5(4), 823-826.

Clegg, S., & Murphy, C. N. (2016). Epidemiology and virulence of Klebsiella

pneumoniae. Microbiology Spectrum, 4(1).

Clements, A., Gaboriaud, F., Duval, J. F., et al. (2008). The major surface-

associated saccharides of Klebsiella pneumoniae contribute to host cell

association. PLOS ONE, 3(11), e3817.

CLSI. (2013). Clinical and Laboratory Standards Institute (CLSI)-Performance

standards for antimicrobial susceptibility testing. Clinical and Laboratory

Standards Institute, 33(1).

Collins, C. M., & D'Orazio, S. F. D. (1996). Virulence determinants of

uropathogenic Klebsiella pneumoniae In H. L. Mobley & J. W. Warren

(Eds.), Urinary Tract Infections; Molecular Pathogenesis and Clinical

Management (pp. 299-312). Washington D. C.: American Society for

Microbiology Press.

Combet, C., Blanchet, C., Geourjon, C., & Deléage, G. (2000). NPS@:

Network protein sequence analysis. Trends in Biochemical Science,

25(3), 147-150.

Connell, S. R., Tracz, D. M., Nierhaus, K. H., & Taylor, D. E. (2003). Ribosomal

protection proteins and their mechanism of tetracycline resistance.

Antimicrobial Agents and Chemotherapy, 47(12), 3675-3681.

324 |

Craig, W. A. (1998). Pharmacokinetic/pharmacodynamic parameters:

rationale for antibacterial dosing of mice and men. Clinical Infectious

Diseases, 26(1), 1-10.

Crawford, R. W., Reeve, K. E., & Gunn, J. S. (2010). Flagellated but not

hyperfimbriated Salmonella enterica serovar Typhimurium attaches to

and forms biofilms on cholesterol-coated surfaces. Journal of

Bacteriology, 192(12), 2981-2990.

Da Re, S., & Ghigo, J. M. (2006). A CsgD-independent pathway for cellulose

production and biofilm formation in Escherichia coli. Journal of

Bacteriology, 188(8), 3073-3087.

Danelon, C., & Winterhalter, M. (2004). Reconstitution of general diffusion

pores from bacterial outer membranes. In R. Benz (Ed.), Bacterial and

Eukaryotic Porins; Structure, Function, Mechanism (pp. 99-117).

Germany: Wiley-VCH Verlag GmbH & Co.

Danziger, L. H., & Neuhauser, M. (2011). Beta-Lactam Antibiotics. In S. C.

Piscitelli, K. A. Rodvold & M. P. Pai (Eds.), Drug Interactions in

Infectious Diseases (pp. 203-242). Totowa, NJ: Humana Press.

Das, T., Sehar, S., & Manefield, M. (2013). The roles of extracellular DNA in

the structural integrity of extracellular polymeric substance and bacterial

biofilm development. Environmental Microbiology Reports, 5(6), 778-

786.

Datsenko, K. A., & Wanner, B. L. (2000). One-step inactivation of

chromosomal genes in Escherichia coli K-12 using PCR products.

Proceedings of the National Academy of Sciences, 97(12), 6640-6645.

Datta, N., & Kontomichalou, P. (1965). Penicillinase synthesis controlled by

infectious R factors in Enterobacteriaceae. Nature, 208(5007), 239-241.

325 |

Davies, D. (2003). Understanding biofilm resistance to antibacterial agents.

Nature Reviews Drug Discovery, 2(2), 114-122.

de Sanctis, J., Teixeira, L., van Duin, D., et al. (2014). Complex prosthetic joint

infections due to carbapenemase-producing Klebsiella pneumoniae: a

unique challenge in the era of untreatable infections. International

Journal of Infectious Diseases, 25, 73-78.

Delcour, A. H. (2003). Solute uptake through general porins. Frontiers in

Bioscience: A journal and Virtual Library, 8, 1055-1071.

Delcour, A. H. (2009). Outer membrane permeability and antibiotic resistance.

Biochima et Biophysica Acta (BBA) - Proteins and Proteomics, 1794(5),

808-816.

Denyer, S. P., & Stewart, G. S. A. B. (1998). Mechanisms of action of

disinfectants. International Biodeterioration & Biodegradation, 41(3),

261-268.

Deris, Z. Z., Heidi, H. Y., Davis, K., et al. (2012). The combination of colistin

and doripenem is synergistic against Klebsiella pneumoniae at multiple

inocula and suppresses colistin resistance in an in vitro

pharmacokinetic/pharmacodynamic model. Antimicrobial Agents and

Chemotherapy, 56(10), 5103-5112.

Di Martino, P., Bertin, Y., Girardeau, J. P., et al. (1995). Molecular

characterization and adhesive properties of CF29K, an adhesin of

Klebsiella pneumoniae strains involved in nosocomial infections.

Infection and Immunity, 63(11), 4336-4344.

Di Martino, P., Livrelli, V., Sirot, D., Joly, B., & Darfeuille-Michaud, A. (1996).

A new fimbrial antigen harbored by CAZ-5/SHV-4-producing Klebsiella

326 |

pneumoniae strains involved in nosocomial infections. Infection and

Immunity, 64(6), 2266-2273.

Dillingham, M. S., & Kowalczykowski, S. C. (2008). RecBCD enzyme and the

repair of double-stranded DNA breaks. Microbiology & Molecular

Biology Reviews, 72(4), 642-671.

Doménech-Sánchez, A., Hernández-Allés, S., Martínez-Martínez, L., Benedí,

V. J., & Albertí, S. (1999). Identification and characterization of a new

porin gene of Klebsiella pneumoniae: its role in beta-lactam antibiotic

resistance. Journal of Bacteriology, 181(9), 2726-2732.

Doménech-Sánchez, A., Martínez-Martínez, L., Hernández-Allés, S., et al.

(2003). Role of Klebsiella pneumoniae OmpK35 porin in antimicrobial

resistance. Antimicrobial Agents and Chemotherapy, 47(10), 3332-

3335.

Domenico, P., Tomas, J. M., Merino, S., Rubires, X., & Cunha, B. A. (1999).

Surface antigen exposure by bismuth dimercaprol suppression of

Klebsiella pneumoniae capsular polysaccharide. Infection and

Immunity, 67(2), 664-669.

Doorduijn, D. J., Rooijakkers, S. H., van Schaik, W., & Bardoel, B. W. (2016).

Complement resistance mechanisms of Klebsiella pneumoniae.

Immunobiology, 221(10), 1102-1109.

Dorman, C. J., & Corcoran, C. P. (2009). Bacterial DNA topology and

infectious disease. Nucleic Acids Research, 37(3), 672-678.

Dou, W., Thompson-Jaeger, S., Laulederkind, S. J., et al. (2005). Defective

expression of Tamm-Horsfall protein/uromodulin in COX-2-deficient

mice increases their susceptibility to urinary tract infections. American

Journal of Physiology: Renal Physiology, 289(1), F49-60.

327 |

Ducel, G., Fabry, J., & Nicolle, L. (2002). Prevention of hospital acquired

infections: a practical guide. Geneva: World Health Organization.

Dueholm, M. S., Sondergaard, M. T., Nilsson, M., et al. (2013). Expression of

Fap amyloids in Pseudomonas aeruginosa, P. fluorescens, and P.

putida results in aggregation and increased biofilm formation.

Microbiology Open, 2(3), 365-382.

Duguid, J. P. (1959). Fimbriae and adhesive properties in Klebsiella strains

Journal of General Microbiology, 21(1), 271-286.

Dulek, D. E., Newcomb, D. C., Goleniewska, K., et al. (2014). Allergic airway

inflammation decreases lung bacterial burden following acute Klebsiella

pneumoniae infection in a neutrophil- and CCL8-dependent manner.

Infection and Immunity, 82(9), 3723-3739.

Dutzler, R., Rummel, G., Albertí, S., et al. (1999). Crystal structure and

functional characterization of OmpK36, the osmoporin of Klebsiella

pneumoniae. Structure, 7(4), 425-434.

ECDC. (2017). Annual Report of the European Antimicrobial Resistance

Surveillance Network (EARS-Net)- Surveillance of Antimicrobial

Resistance in Europe 2016. Retrieved from Stockholm:

https://ecdc.europa.eu/en/publications-data/antimicrobial-resistance-

surveillance-europe-2016

Enzler, M. J., Berbari, E., & Osmon, D. R. (2011). Antimicrobial prophylaxis in

adults. Mayo Clinic Proceedings, 86(7), 686-701.

EUCAST. (2014). Breakpoint tables for interpretation of MICs and zone

diameters. EUCAST: Clinical breakpoints. Version 4. Retrieved from

http://www.eucast.org

328 |

Fabrega, A., Madurga, S., Giralt, E., & Vila, J. (2009). Mechanism of action of

and resistance to quinolones. Microbial Biotechnology, 2(1), 40-61.

Fane, M. E. L., Dollo, I., Sodqi, M., et al. (2005). Septicaemia caused by

community acquired Klebsiella pneumonia complicated with

endocarditis and meningoencephalitis: A case report. Journal of

Cardiovascular Diseases & Diagnosis, 4(3).

Fang, C. T., Shih, Y. J., Cheong, C. M., & Yi, W. C. (2016). Rapid and accurate

determination of lipopolysaccharide O-antigen types in Klebsiella

pneumoniae with a novel PCR-based O-genotyping method. Journal of

Clinical Microbiology, 54(3), 666-675.

Fang, X., Ahmad, I., Blanka, A., et al. (2014). GIL, a new c-di-GMP-binding

protein domain involved in regulation of cellulose synthesis in

Enterobacteria. Molecular Microbiology, 93(3), 439-452.

Feldman, M. B., Terry, D. S., Altman, R. B., & Blanchard, S. C. (2010).

Aminoglycoside activity observed on single pre-translocation ribosome

complexes. Nature Chemical Biology, 6(1), 54-62.

Feneley, R. C., Hopley, I. B., & Wells, P. N. T. (2015). Urinary catheters:

history, current status, adverse events and research agenda. Journal of

Medical Engineering & Technology, 39(8), 459-470.

Feng, X., Oropeza, R., Walthers, D., & Kenney, L. J. (2003). OmpR

phosphorylation and its role in signaling and pathogenesis. ASM News

American Society for Microbiology, 69(8), 390-395.

Feng, Y., Johnston, R. N., Liu, G. R., & Liu, S. L. (2013). Genomic comparison

between Salmonella Gallinarum and Pullorum: Differential pseudogene

formation under common host restriction. PLOS ONE, 8(3), e59427.

329 |

Fernández, L., & Hancock, R. E. (2012). Adaptive and mutational resistance:

role of porins and efflux pumps in drug resistance. Clinical Microbiology

Reviews, 25(4), 661-681.

Ferrieres, L., Hancock, V., & Klemm, P. (2007). Biofilm exclusion of

uropathogenic bacteria by selected asymptomatic bacteriuria

Escherichia coli strains. Microbiology, 153(6), 1711-1719.

Fiser, A. (2010). Template-based protein structure modeling. Methods in

Molecular Biology, 673, 73-94.

Flemming, H. C. (2011). The perfect slime. Colloids and Surfaces B:

Biointerfaces, 86(2), 251-259.

Flemming, H. C., & Wingender, J. (2010). The biofilm matrix. Nature Reviews

Microbiology, 8(9), 623-633.

Flores-Mireles, A. L., Walker, J. N., Caparon, M., & Hultgren, S. J. (2015).

Urinary tract infections: epidemiology, mechanisms of infection and

treatment options. Nature Reviews Microbiology, 13(5), 269-284.

Fong, J. N., & Yildiz, F. H. (2015). Biofilm matrix proteins. Microbiology

Spectrum, 3(2).

Foster, P. L. (2007). Stress-induced mutagenesis in bacteria. Critical Reviews

in Biochemistry and Molecular Biology, 42(5), 373-397.

Fournet-Fayard, S., Joly, B., & Forestier, C. (1995). Transformation of wild type

Klebsiella pneumoniae with plasmid DNA by electroporation. Journal of

Microbiological Methods, 24(1), 49-54.

330 |

Foxman, B. (2010). The epidemiology of urinary tract infection. Nature

Reviews Urology, 7(12), 653-660.

French, G. L., Shannon, K. P., & Simmons, N. (1996). Hospital outbreak of

Klebsiella pneumoniae resistant to broad-spectrum cephalosporins and

beta-lactam-beta-lactamase inhibitor combinations by hyperproduction

of SHV-5 beta-lactamase. Journal of Clinical Microbiology, 34(2), 358-

363.

Fung, C.-P., Lin, Y.-T., Lin, J.-C., et al. (2012). Klebsiella pneumoniae in

gastrointestinal tract and pyogenic liver abscess. Emerging Infectious

Diseases, 18(8), 1322-1325.

Galdiero, E., Kampanaraki, A., Mignogna, E., & Galdiero, M. (2008). Porins in

the inflammatory and immunological response following Salmonella

infections. In Y. Kumar (Ed.), Salmonella- A Diversified Superbug.

Rijeka, Croatia: InTech Open Access Publisher.

Galimand, M., Courvalin, P., & Lambert, T. (2003). Plasmid-mediated high-

level resistance to aminoglycosides in Enterobacteriaceae due to 16S

rRNA methylation. Antimicrobial Agents and Chemotherapy, 47(8),

2565-2571.

García-Sureda, L., Doménech-Sánchez, A., Barbier, M., et al. (2011).

OmpK26, a novel porin associated with carbapenem resistance in

Klebsiella pneumoniae. Antimicrobial Agents and Chemotherapy,

55(10), 4742-4747.

Geisinger, E., & Isberg, R. R. (2015). Antibiotic modulation of capsular

exopolysaccharide and virulence in Acinetobacter baumannii. PLOS

Pathogens, 11(2), e1004691.

331 |

Gerba, C. P. (2015). Quaternary ammonium biocides: efficacy in application.

Applied and Environmental Microbiology, 81(2), 464-469.

Gerlach, G. F., Allen, B. L., & Clegg, S. (1988). Molecular characterization of

the type 3 (MR/K) fimbriae of Klebsiella pneumoniae. Journal of

Bacteriology, 170(8), 3547-3553.

Gilan, I., & Sivan, A. (2013). Effect of proteases on biofilm formation of the

plastic-degrading actinomycete Rhodococcus ruber C208. FEMS

Microbiology Letters, 342(1), 18-23.

Goh, E.-B., Yim, G., Tsui, W., et al. (2002). Transcriptional modulation of

bacterial gene expression by subinhibitory concentrations of antibiotics.

Proceedings of the National Academy of Sciences, 99(26), 17025-

17030.

Gonzalez-Escobedo, G., Marshall, J. M., & Gunn, J. S. (2011). Chronic and

acute infection of the gall bladder by Salmonella typhi: understanding

the carrier state. Nature Reviews Microbiology, 9(1), 9-14.

Goodman, S. D., Obergfell, K. P., Jurcisek, J. A., et al. (2011). Biofilms can be

dispersed by focusing the immune system on a common family of

bacterial nucleoid-associated proteins. Mucosal Immunology, 4(6), 625-

637.

Haeggman, S., Lofdahl, S., & Burman, L. G. (1997). An allelic variant of the

chromosomal gene for class A beta-lactamase K2, specific for

Klebsiella pneumoniae, is the ancestor of SHV-1. Antimicrobial Agents

and Chemotherapy, 41(12), 2705-2709.

Hall, B. G., Acar, H., Nandipati, A., & Barlow, M. (2014). Growth rates made

easy. Molecular Biology and Evolution, 31(1), 232-238.

332 |

Hall, T. A. (1999). BioEdit: a user-friendly biological sequence alignment editor

and analysis program for Windows 95/98/NT. Paper presented at the

Nucleic Acids Symposium Series.

Hardesty, C., Ferran, C., & DiRienzo, J. M. (1991). Plasmid-mediated sucrose

metabolism in Escherichia coli: characterization of scrY, the structural

gene for a phosphoenolpyruvate-dependent sucrose

phosphotransferase system outer membrane porin. Journal of

Bacteriology, 173(2), 449-456.

Harmsen, M., Lappann, M., Knøchel, S., & Molin, S. (2010). Role of

extracellular DNA during biofilm formation by Listeria monocytogenes.

Applied and Environmental Microbiology, 76(7), 2271-2279.

Hashemi, A., Fallah, F., Erfanimanesh, S., et al. (2014). Detection of β-

lactamases and outer membrane porins among Klebsiella pneumoniae

strains isolated in Iran. Scientifica, 2014, 6.

Hasman, H., Chakraborty, T., & Klemm, P. (1999). Antigen-43-mediated

autoaggregation of Escherichia coli is blocked by fimbriation. Journal of

Bacteriology, 181(16), 4834-4841.

Hennequin, C., & Forestier, C. (2009). oxyR, a LysR-type regulator involved in

Klebsiella pneumoniae mucosal and abiotic colonization. Infection and

Immunity, 77(12), 5449-5457.

Hernández-Allés, S., Albertí, S., Álvarez, D., et al. (1999). Porin expression in

clinical isolates of Klebsiella pneumoniae. Microbiology, 145(3), 673-

679.

Hernández-Allés, S., Conejo, M., Pascual, A., et al. (2000). Relationship

between outer membrane alterations and susceptibility to antimicrobial

333 |

agents in isogenic strains of Klebsiella pneumoniae. Journal of

Antimicrobial Chemotherapy, 46(2), 273-277.

Hickerson, A. D., & Carson, C. C. (2006). The treatment of urinary tract

infections and use of ciprofloxacin extended release. Expert Opinion on

Investigational Drugs, 15(5), 519-532.

Hickman, J. W., Tifrea, D. F., & Harwood, C. S. (2005). A chemosensory

system that regulates biofilm formation through modulation of cyclic

diguanylate levels. Proceedings of the National Academy of Sciences,

102(40), 14422-14427.

Hidron, A. I., Edwards, J. R., Patel, J., et al. (2008). NHSN annual update:

antimicrobial-resistant pathogens associated with healthcare-

associated infections: annual summary of data reported to the National

Healthcare Safety Network at the Centers for Disease Control and

Prevention, 2006-2007. Infection Control and Hospital Epidemiology,

29(11), 996-1011.

Hirsch, E. B., & Tam, V. H. (2010). Detection and treatment options for

Klebsiella pneumoniae carbapenemases (KPCs): an emerging cause

of multidrug-resistant infection. Journal of Antimicrobial Chemotherapy,

65(6), 1119-1125.

Hobley, L., Harkins, C., MacPhee, C. E., & Stanley-Wall, N. R. (2015). Giving

structure to the biofilm matrix: an overview of individual strategies and

emerging common themes. FEMS Microbiology Reviews, 39(5), 649-

669.

Hocquet, D., Vogne, C., El Garch, F., et al. (2003). MexXY-OprM efflux pump

is necessary for a adaptive resistance of Pseudomonas aeruginosa to

aminoglycosides. Antimicrobial Agents and Chemotherapy, 47(4),

1371-1375.

334 |

Hoffman, L. R., D'Argenio, D. A., MacCoss, M. J., et al. (2005). Aminoglycoside

antibiotics induce bacterial biofilm formation. Nature, 436(7054), 1171-

1175.

Holden, V. I., Breen, P., Houle, S., Dozois, C. M., & Bachman, M. A. (2016).

Klebsiella pneumoniae siderophores induce inflammation, bacterial

dissemination, and HIF-1α stabilization during pneumonia. mBIO, 7(5),

e01397-01316.

Hopkins, S., Shaw, K., & Simpson, L. (2012). English National Point

Prevalence Survey on Healthcare Associated Infections and

Antimicrobial Use, 2011: Preliminary data. Retrieved from

http://www.hpa.org.uk/webc/HPAwebFile/HPAwebC/1317134305239.

Hsueh, P.-R., Badal, R. E., Hawser, S. P., et al. Epidemiology and

antimicrobial susceptibility profiles of aerobic and facultative Gram-

negative bacilli isolated from patients with intra-abdominal infections in

the Asia-Pacific region: 2008 results from SMART (Study for Monitoring

Antimicrobial Resistance Trends). International Journal of Antimicrobial

Agents, 36(5), 408-414.

Huang, T.-W., Lam, I., Chang, H.-Y., et al. (2014). Capsule deletion via a λ-

Red knockout system perturbs biofilm formation and fimbriae

expression in Klebsiella pneumoniae MGH 78578. BMC Research

Notes, 7, 13-13.

Huertas, M. G., Zarate, L., Acosta, I. C., et al. (2014). Klebsiella pneumoniae

yfiRNB operon affects biofilm formation, polysaccharide production and

drug susceptibility. Microbiology, 160(12), 2595-2606.

Humeniuk, C., Arlet, G., Gautier, V., et al. (2002). Beta-lactamases of Kluyvera

ascorbata, probable progenitors of some plasmid-encoded CTX-M

types. Antimicrobial Agents and Chemotherapy, 46(9), 3045-3049.

335 |

Humphries, R. M., Kelesidis, T., Bard, J. D., et al. (2010). Successful treatment

of pan-resistant Klebsiella pneumoniae pneumonia and bacteraemia

with a combination of high-dose tigecycline and colistin. Journal of

Medical Microbiology, 59(11), 1383-1386.

Ipe, D. S., Horton, E., & Ulett, G. C. (2016). The basics of bacteriuria:

Strategies of microbes for persistence in urine. Frontiers in Cellular and

Infection Microbiology, 6, 14.

Izard, D., Ferragut, C., Gavini, F., et al. (1981). Klebsiella terrigena, a new

species from soil and water International Journal of Systematic

Bacteriology, 31(2), 116-127.

Jacoby, G. A., Chow, N., & Waites, K. B. (2003). Prevalence of plasmid-

mediated quinolone resistance. Antimicrobial Agents and

Chemotherapy, 47(2), 559-562.

Jacoby, G. A., Mills, D. M., & Chow, N. (2004). Role of β-lactamases and

porins in resistance to ertapenem and other β-lactams in Klebsiella

pneumoniae. Antimicrobial Agents and Chemotherapy, 48(8), 3203-

3206.

Jagnow, J., & Clegg, S. (2003). Klebsiella pneumoniae MrkD-mediated biofilm

formation on extracellular matrix- and collagen-coated surfaces.

Microbiology, 149(9), 2397-2405.

Jahn, C. E., Selimi, D. A., Barak, J. D., & Charkowski, A. O. (2011). The

Dickeya dadantii biofilm matrix consists of cellulose nanofibres, and is

an emergent property dependent upon the type III secretion system and

the cellulose synthesis operon. Microbiology, 157(10), 2733-2744.

336 |

Jiang, X., Espedido, B. A., Partridge, S. R., et al. (2009). Paradoxical effect of

Klebsiella pneumoniae OmpK36 porin deficiency. Pathology, 41(4),

388-392.

Johnson, J. G., Murphy, C. N., Sippy, J., Johnson, T. J., & Clegg, S. (2011).

Type 3 fimbriae and biofilm formation are regulated by the

transcriptional regulators MrkHI in Klebsiella pneumoniae. Journal of

Bacteriology, 193(14), 3453-3460.

Johnson, J. G., Spurbeck, R. R., Sandhu, S. K., & Matson, J. S. (2014).

Genome sequence of Klebsiella pneumoniae urinary tract isolate

Top52. Genome Announcements, 2(4), e00668-00614.

Jones, K. L., Kim, S. W., & Keasling, J. D. (2000). Low-copy plasmids can

perform as well as or better than high-copy plasmids for metabolic

engineering of bacteria. Metabolic Engineering, 2(4), 328-338.

Jones, M. E., Sahm, D. F., Martin, N., et al. (2000). Prevalence of gyrA, gyrB,

parC, and parE mutations in clinical isolates of Streptococcus

pneumoniae with decreased susceptibilities to different

fluoroquinolones and originating from Worldwide Surveillance Studies

during the 1997-1998 respiratory season. Antimicrobial Agents and

Chemotherapy, 44(2), 462-466.

Jorgensen, I., & Seed, P. C. (2012). How to make it in the urinary tract: A

tutorial by Escherichia coli. PLOS Pathogens, 8(10), e1002907.

Kabsch, W., & Sander, C. (1983). Dictionary of protein secondary structure:

pattern recognition of hydrogen-bonded and geometrical features.

Biopolymers, 22(12), 2577-2637.

Kaczmarek, F. M., Dib-Hajj, F., Shang, W., & Gootz, T. D. (2006). High-level

carbapenem resistance in a Klebsiella pneumoniae clinical isolate is

337 |

due to the combination of blaACT-1 β-lactamase production, porin

OmpK35/36 insertional inactivation, and down-regulation of the

phosphate transport porin PhoE. Antimicrobial Agents and

Chemotherapy, 50(10), 3396-3406.

Kaplan, J. B. (2011). Antibiotic-induced biofilm formation. The International

Journal of Artificial Organs, 34(9), 737-751.

Kaplan, J. B., Ragunath, C., Ramasubbu, N., & Fine, D. H. (2003). Detachment

of Actinobacillus actinomycetemcomitans biofilm cells by an

endogenous β-hexosaminidase activity. Journal of Bacteriology,

185(16), 4693-4698.

Karim, A., Poirel, L., Nagarajan, S., & Nordmann, P. (2001). Plasmid-mediated

extended-spectrum beta-lactamase (CTX-M-3 like) from India and gene

association with insertion sequence ISEcp1. FEMS Microbiology

Letters, 201(2), 237-241.

Kato, J., Nishimura, Y., Imamura, R., et al. (1990). New topoisomerase

essential for chromosome segregation in E. coli. Cell, 63(2), 393-404.

Khandige, S., & Møller-Jensen, J. (2016). Fimbrial phase variation: stochastic

or cooperative? Current Genetics, 62(2), 237-241.

Kishi, K., Yasuda, T., & Takeshita, H. (2001). DNase I: structure, function, and

use in medicine and forensic science. Legal Medicine, 3(2), 69-83.

Klebba, P. E. (2002). Mechanism of maltodextrin transport through LamB.

Research in Microbiology, 153(7), 417-424.

Knothe, H., Shah, P., Krcmery, V., Antal, M., & Mitsuhashi, S. (1983).

Transferable resistance to cefotaxime, cefoxitin, cefamandole and

338 |

cefuroxime in clinical isolates of Klebsiella pneumoniae and Serratia

marcescens. Infection, 11(6), 315-317.

Ko, W. C., Paterson, D. L., Sagnimeni, A. J., et al. (2002). Community-acquired

Klebsiella pneumoniae bacteremia: global differences in clinical

patterns. Emerging Infectious Diseases, 8(2), 160-166.

Kong, Q., Beanan, J. M., Olson, R., et al. (2012). Biofilm formed by a

hypervirulent (hypermucoviscous) variant of Klebsiella pneumoniae

does not enhance serum resistance or survival in an in vivo abscess

model. Virulence, 3(3), 309-318.

Korea, C. G., Ghigo, J. M., & Beloin, C. (2011). The sweet connection: Solving

the riddle of multiple sugar-binding fimbrial adhesins in Escherichia coli.

Bioessays, 33(4), 300-311.

Kotra, L. P., Haddad, J., & Mobashery, S. (2000). Aminoglycosides:

perspectives on mechanisms of action and resistance and strategies to

counter resistance. Antimicrobial Agents and Chemotherapy, 44(12),

3249-3256.

Krewulak, K. D., & Vogel, H. J. (2008). Structural biology of bacterial iron

uptake. Biochimica et Biophysica Acta (BBA) - Biomembranes, 1778(9),

1781-1804.

Kropec, A., Maira-Litran, T., Jefferson, K. K., et al. (2005). Poly-N-

acetylglucosamine production in Staphylococcus aureus is essential for

virulence in murine models of systemic infection. Infection and

Immunity, 73(10), 6868-6876.

Kumar, A., Mallik, D., Pal, S., et al. (2015). Escherichia coli O8-antigen

enhances biofilm formation under agitated conditions. FEMS

Microbiology Letters, 362(15).

339 |

LaBombardi, V. J., Rojtman, A., & Tran, K. (2006). Use of cefepime for the

treatment of infections caused by extended spectrum β-lactamase–

producing Klebsiella pneumoniae and Escherichia coli. Diagnostic

Microbiology and Infectious Disease, 56(3), 313-315.

Landman, D., Bratu, S., & Quale, J. (2009). Contribution of OmpK36 to

carbapenem susceptibility in KPC-producing Klebsiella pneumoniae.

Journal of Medical Microbiology, 58(10), 1303-1308.

Latasa, C., Solano, C., Penadés, J. R., & Lasa, I. (2006). Biofilm-associated

proteins. Comptes Rendus Biologies, 329(11), 849-857.

Lautrop, H. (1956). Gelatin liquefying Klebsiella strains (Bacterium oxytocum

(Flügge)). Acta Pathologica Microbiologica Scandinavica, 39(5), 375-

384.

Le Quéré, B., & Ghigo, J. M. (2009). BcsQ is an essential component of the

Escherichia coli cellulose biosynthesis apparatus that localizes at the

bacterial cell pole. Molecular Microbiology, 72(3), 724-740.

Lebeaux, D., Chauhan, A., Rendueles, O., & Beloin, C. (2013). From in vitro

to in vivo models of bacterial biofilm-related infections. Pathogens, 2(2),

288-356.

Lebeaux, D., Ghigo, J. M., & Beloin, C. (2014). Biofilm-related infections:

bridging the gap between clinical management and fundamental

aspects of recalcitrance toward antibiotics. Microbiology and Molecular

Biology Reviews, 78(3), 510-543.

LeBlanc, D. J., Lee, L. N., & Inamine, J. M. (1991). Cloning and nucleotide

base sequence analysis of a spectinomycin adenyltransferase AAD(9)

determinant from Enterococcus faecalis. Antimicrobial Agents and

Chemotherapy, 35(9), 1804-1810.

340 |

Leclercq, R. (2002). Mechanisms of resistance to macrolides and

lincosamides: nature of the resistance elements and their clinical

implications. Clinical Infectious Diseases, 34(4), 482-492.

Lee, D., Redfern, O., & Orengo, C. (2007). Predicting protein function from

sequence and structure. Nature Reviews Molecular Cell Biology, 8(12),

995-1005.

Levison, M. E., & Levison, J. H. (2009). Pharmacokinetics and

pharmacodynamics of antibacterial agents. Infectious Disease Clinics

of North America, 23(4), 791-815.

Lewis, K. (2001). Riddle of biofilm resistance. Antimicrobial Agents and

Chemotherapy, 45(4), 999-1007.

Liao, H., Zhang, X. Z., Rollin, J. A., & Zhang, Y. H. (2011). A minimal set of

bacterial cellulases for consolidated bioprocessing of lignocellulose.

Biotechnology Journal, 6(11), 1409-1418.

Limoli, D. H., Jones, C. J., & Wozniak, D. J. (2015). Bacterial extracellular

polysaccharides in biofilm formation and function. Microbiology

Spectrum, 3(3).

Lin, C.-W., Huang, Y.-W., Hu, R.-M., Chiang, K.-H., & Yang, T.-C. (2009). The

role of AmpR in regulation of L1 and L2 β-lactamases in

Stenotrophomonas maltophilia. Research in Microbiology, 160(2), 152-

158.

Lin, W.-P., Wang, J.-T., Chang, S.-C., et al. (2016). The antimicrobial

susceptibility of Klebsiella pneumoniae from community settings in

Taiwan, a Trend Analysis. Scientific Reports, 6, 36280.

341 |

Linares, J. F., Gustafsson, I., Baquero, F., & Martinez, J. L. (2006). Antibiotics

as intermicrobial signaling agents instead of weapons. Proceedings of

the National Academy of Sciences, 103(51), 19484-19489.

Liu, Y., & Tay, J.-H. (2002). The essential role of hydrodynamic shear force in

the formation of biofilm and granular sludge. Water Research, 36(7),

1653-1665.

Llobet, E., March, C., Giménez, P., & Bengoechea, J. A. (2009). Klebsiella

pneumoniae OmpA confers resistance to antimicrobial peptides.

Antimicrobial Agents and Chemotherapy, 53(1), 298-302.

Loveday, H. P., Wilson, J. A., Pratt, R. J., et al. (2014). epic3: National

evidence-based guidelines for preventing healthcare-acssociated

infections in NHS hospitals in England. Journal of Hospital Infection,

86(Supplement 1), S1-S70.

Lugtenberg, B., & Van Alphen, L. (1983). Molecular architecture and

functioning of the outer membrane of Escherichia coli and other Gram-

negative bacteria. Biochimica et Biophysica Acta (BBA)- Reviews on

Biomembranes, 737(1), 51-115.

Lynch, A. S., & Robertson, G. T. (2008). Bacterial and fungal biofilm infections.

Annual Review of Medicine, 59, 415-428.

Mack, D., Fischer, W., Krokotsch, A., et al. (1996). The intercellular adhesin

involved in biofilm accumulation of Staphylococcus epidermidis is a

linear beta-1,6-linked glucosaminoglycan: purification and structural

analysis. Journal of Bacteriology, 178(1), 175-183.

Magnet, S., & Blanchard, J. S. (2005). Molecular insights into aminoglycoside

action and resistance. Chemical Reviews, 105(2), 477-498.

342 |

Mah, T. F. C., & O'Toole, G. A. (2001). Mechanisms of biofilm resistance to

antimicrobial agents. Trends in Microbiology, 9(1), 34-39.

Maillard, J. Y. (2002). Bacterial target sites for biocide action. Journal of

Applied Microbiology, 92(Supplement 1), 16S-27S.

Majtan, J., Majtanova, L., Xu, M., & Majtan, V. (2008). In vitro effect of

subinhibitory concentrations of antibiotics on biofilm formation by

clinical strains of Salmonella enterica serovar Typhimurium isolated in

Slovakia. Journal of Applied Microbiology, 104(5), 1294-1301.

Mallea, M., Chevalier, J., Bornet, C., et al. (1998). Porin alteration and active

efflux: two in vivo drug resistance strategies used by Enterobacter

aerogenes. Microbiology, 144 (11), 3003-3009.

Martinez-Martinez, L., Pascual, A., & Jacoby, G. A. (1998). Quinolone

resistance from a transferable plasmid. Lancet, 351(9105), 797-799.

Martínez‐Martínez, L. (2008). Extended-spectrum β-lactamases and the

permeability barrier. Clinical Microbiology & Infection, 14(Supplement

1), 82-89.

McCoy, W. F., Bryers, J. D., Robbins, J., & Costerton, J. W. (1981).

Observations of fouling biofilm formation. Canadian Journal of

Microbiology, 27(9), 910-917.

McCusker, M. P., Turner, E. C., & Dorman, C. J. (2008). DNA sequence

heterogeneity in Fim tyrosine-integrase recombinase-binding elements

and functional motif asymmetries determine the directionality of the fim

genetic switch in Escherichia coli K-12. Molecular Microbiology, 67(1),

171-187.

343 |

McDonnell, G., & Russell, A. D. (1999). Antiseptics and disinfectants: Activity,

action, and resistance. Clinical Microbiology Reviews, 12(1), 147-179.

McDonnell, G. E. (2007). Antisepsis, disinfection, and sterilization: Types,

action and resistance. Washington D. C.: American Society of

Microbiology.

McKenna, M. (2013). The last resort. Nature, 499(7459), 394-396.

Medeiros, A. A. (1984). Beta-lactamases. British Medical Bulletin, 40(1), 18-

27.

Meile, K., Zhurinsh, A., & Spince, B. (2014). Aspects of periodate oxidation of

carbohydrates for the analysis of pyrolysis liquids. Journal of

Carbohydrate Chemistry, 33(3), 105-116.

Messenger, A. J., & Barclay, R. (1983). Bacteria, iron and pathogenicity.

Biochemical Education, 11(2), 54-63.

Miller, S. I. (2016). Antibiotic resistance and regulation of the Gram-negative

bacterial outer membrane barrier by host innate immune molecules.

mBIO, 7(5), e01541-01516.

Mingeot-Leclercq, M. P., Glupczynski, Y., & Tulkens, P. M. (1999).

Aminoglycosides: activity and resistance. Antimicrobial Agents and

Chemotherapy, 43(4), 727-737.

Montanaro, L., Poggi, A., Visai, L., et al. (2011). Extracellular DNA in biofilms.

International Journal of Artificial Organs, 34(9), 824-831.

Moreira, J. M., Gomes, L. C., Araújo, J. D., et al. (2013). The effect of glucose

concentration and shaking conditions on Escherichia coli biofilm

344 |

formation in microtiter plates. Chemical Engineering Science, 94, 192-

199.

Morgan, J. L. W., McNamara, J. T., & Zimmer, J. (2014). Mechanism of

activation of bacterial cellulose synthase by cyclic-di-GMP. Nature

Structural & Molecular Biology, 21(5), 489-496.

Muheim, C., Götzke, H., Eriksson, A. U., et al. (2017). Increasing the

permeability of Escherichia coli using MAC13243. Scientific Reports,

7(1), 17629.

Mullany, P., Wilks, M., Lamb, I., et al. (1990). Genetic analysis of a tetracycline

resistance element from Clostridium difficile and its conjugal transfer to

and from Bacillus subtilis. Journal of General Microbiology, 136(7),

1343-1349.

Mulvey, M. A., Schilling, J. D., Martinez, J. J., & Hultgren, S. J. (2000). Bad

bugs and beleaguered bladders: interplay between uropathogenic

Escherichia coli and innate host defenses. Proceedings of the National

Academy of Sciences, 97(16), 8829-8835.

Munita, J. M., & Arias, C. A. (2016). Mechanisms of antibiotic resistance.

Microbiology Spectrum, 4(2), 1-37.

Murphy, C. N., & Clegg, S. (2012). Klebsiella pneumoniae and type 3 fimbriae:

nosocomial infection, regulation and biofilm formation. Future

Microbiology, 7(8), 991-1002.

Murphy, C. N., Mortensen, M. S., Krogfelt, K. A., & Clegg, S. (2013). Role of

Klebsiella pneumoniae type 1 and type 3 fimbriae in colonizing silicone

tubes implanted into the bladders of mice as a model of catheter-

associated urinary tract infections. Infection and Immunity, 81(8), 3009-

3017.

345 |

Naparstek, L., Carmeli, Y., Navon-Venezia, S., & Banin, E. (2014). Biofilm

formation and susceptibility to gentamicin and colistin of extremely

drug-resistant KPC-producing Klebsiella pneumoniae. Journal of

Antimicrobial Chemotherapy, 69(4), 1027-1034.

Naville, M., Ghuillot-Gaudeffroy, A., Marchais, A., & Gautheret, D. (2011).

ARNold: A web tool for the prediction of Rho-independent transcription

terminators. RNA Biology, 8(1), 11-13.

Neal, D. E. (2008). Complicated urinary tract infections. Urologic Clinics of

North America, 35(1), 13-22.

Nicholas, K. B., & Nicholas, H. B. J. (1997). GeneDoc: a tool for editing and

annotating multiple sequence alignments.

http://www.psc.edu/biomed/genedoc.

Nicolle, L. E. (2014). Catheter associated urinary tract infections. Antimicrobial

Resistance and Infection Control, 3(1), 23.

Nikaido, H. (1988). Bacterial resistance to antibiotics as a function of outer

membrane permeability. Journal of Antimicrobial Chemotherapy,

22(Supplement A), 17-22.

Nikaido, H. (1994). Porins and specific diffusion channels in bacterial outer

membranes. The Journal of Biological Chemistry, 269(6), 3906-3908.

Nikaido, H. (1996a). Multidrug efflux pumps of Gram-negative bacteria.

Journal of Bacteriology, 178(20), 5853-5859.

Nikaido, H. (1996b). Outer membrane. In F. C. Neidhardt (Ed.), Escherichia

coli and Salmonella: Cellular and Molecular Biology (pp. 29-47).

Washington, DC: ASM Press.

346 |

Nikaido, H. (2003). Molecular basis of bacterial outer membrane permeability

revisited. Microbiology & Molecular Biology Reviews, 67(4), 593-656.

Nordmann, P., Cuzon, G., & Naas, T. (2009). The real threat of Klebsiella

pneumoniae carbapenemase-producing bacteria. The Lancet

Infectious Diseases, 9(4), 228-236.

Nordmann, P., Naas, T., & Poirel, L. (2011). Global spread of carbapenemase-

producing Enterobacteriaceae. Emerging Infectious Diseases, 17(10),

1791-1798.

Nuccio, S. P., & Bäumler, A. J. (2007). Evolution of the chaperone/usher

assembly pathway: fimbrial classification goes Greek. Microbiology &

Molecular Biology Reviews, 71(4), 551-575.

Nucleo, E., Steffanoni, L., Fugazza, G., et al. (2009). Growth in glucose-based

medium and exposure to subinhibitory concentrations of imipenem

induce biofilm formation in a multidrug-resistant clinical isolate of

Acinetobacter baumannii. BMC Microbiology, 9(1), 270.

O'Neill, J. (2016). Tackling drug-resistant infections globally: Final report and

recommendations.

O'Toole, G. A. (2011). Microtiter dish biofilm formation assay. Journal of

Visualized Experiments : JoVE(47), 2437.

Oelschlaeger, T. A., & Tall, B. D. (1997). Invasion of cultured human epithelial

cells by Klebsiella pneumoniae isolated from the urinary tract. Infection

and Immunity, 65(7), 2950-2958.

Ojo, K. K., Ulep, C., Van Kirk, N., et al. (2003). The mef(A) gene predominates

among seven macrolide resistance genes identified in Gram-negative

347 |

strains representing 13 genera, isolated from healthy Portuguese

children. Antimicrobial Agents and Chemotheraphy, 48(69), 3451-3456.

Okshevsky, M., & Meyer, R. L. (2014). Evaluation of fluorescent stains for

visualizing extracellular DNA in biofilms. Journal of Microbiological

Methods, 105, 102-104.

Old, D. C., Tavendale, A., & Senior, B. W. (1985). A comparative study of the

type-3 fimbriae of Klebsiella species. Journal of Medical Microbiology,

20(2), 203-214.

Olivares, E., Badel-Berchoux, S., Provot, C., et al. (2016). The BioFilm Ring

test: a rapid method for routine analysis of Pseudomonas aeruginosa

biofilm formation kinetics. Journal of Clinical Microbiology, 54(3), 657-

661.

Onori, R., Gaiarsa, S., Comandatore, F., et al. (2015). Tracking nosocomial

Klebsiella pneumoniae infections and outbreaks by whole-genome

analysis: Small-scale Italian scenario within a single hospital. Journal of

Clinical Microbiology, 53(9), 2861-2868.

Ørskov, I. (1984). Genus v. Klebsiella. In N. R. Krieg & J. R. Holt (Eds.),

Bergey's Manual of Systematic Bacteriology (Vol. 1, pp. 461-465):

Springer.

Paczosa, M. K., & Mecsas, J. (2016). Klebsiella pneumoniae: Going on the

offense with a strong defense. Microbiology & Molecular Biology

Reviews, 80(3), 629-661.

Pagès, J. M., James, C. E., & Winterhalter, M. (2008). The porin and the

permeating antibiotic: a selective diffusion barrier in Gram-negative

bacteria. Nature Reviews Microbiology, 6(12), 893-903.

348 |

Pak, J., Pu, Y., Zhang, Z. T., Hasty, D. L., & Wu, X. R. (2001). Tamm-Horsfall

protein binds to type 1 fimbriated Escherichia coli and prevents E. coli

from binding to uroplakin Ia and Ib receptors. Journal of Biological

Chemistry, 276(13), 9924-9930.

Pan, Y. J., Lin, T. L., Chen, Y. H., et al. (2013). Capsular types of Klebsiella

pneumoniae revisited by wzc sequencing. PLOS ONE, 8(12), e80670.

Pan, Y. J., Lin, T. L., Hsu, C. R., & Wang, J. T. (2011). Use of a Dictyostelium

model for isolation of genetic loci associated with phagocytosis and

virulence in Klebsiella pneumoniae. Infection and Immunity, 79(3), 997-

1006.

Pangon, B., Bizet, C., Bure, A., et al. (1989). In vivo selection of a cephamycin-

resistant, porin-deficient mutant of Klebsiella pneumoniae producing a

TEM-3 beta-lactamase. The Journal of Infectious Diseases, 159(5),

1005-1006.

Paris, T., Skali-Lami, S., & Block, J.-C. (2007). Effect of wall shear rate on

biofilm deposition and grazing in drinking water flow chambers.

Biotechnology and Bioengineering, 97(6), 1550-1561.

Parkinson, R., Holden, S., & Clarkson, A. (2016). Guideline for antibiotic

surgical prophylaxis in adult urology. Nottingham University Hospital.

Paterson, D. L. (2006). Resistance in Gram-negative bacteria:

Enterobacteriaceae. The American Journal of Medicine, 119(6), S20-

S28.

Paterson, D. L., Ko, W. C., Von Gottberg, A., et al. (2004). Antibiotic therapy

for Klebsiella pneumoniae bacteremia: Implications of production of

extended-spectrum β-lactamases. Clinical Infectious Diseases, 39(1),

31-37.

349 |

Paterson, D. L., Siu, L. K., & Chang, F. Y. (2014). Klebsiella species (K.

pneumoniae, K. oxytoca, K. ozaenae and K. rhinoscleromatis).

Retrieved from http://www.antimicrobe.org/b107.asp

Peters, A. C., & Wimpenny, J. W. T. (1987). A constant-depth laboratory model

film fermentor. Biotechnology and Bioengineering, 32(3), 263-270.

Pettersen, E. F., Goddard, T. D., Huang, C. C., et al. (2004). UCSF Chimera -

A visualization system for exploratory research and analysis. Journal of

Computational Chemistry, 25(13), 1605-1612.

Podschun, R., & Ullmann, U. (1998). Klebsiella spp. as nosocomial pathogens:

epidemiology, taxonomy, typing methods, and pathogenicity factors.

Clinical Microbiology Reviews, 11(4), 589-603.

Poole, K. (2012). Bacterial stress responses as determinants of antimicrobial

resistance. Journal of Antimicrobial Chemotherapy, 67(9), 2069-2089.

Prakash, M. (2008). Substitutional evoution. In Encyclopaedia of Gene

Evolution - Molecular Genetics (pp. 165-193). New Delhi: Discovery

Publishing House New Delhi.

Pratt, L. A., & Kolter, R. (1998). Genetic analysis of Escherichia coli biofilm

formation: roles of flagella, motility, chemotaxis and type I pili. Molecular

Microbiology, 30(2), 285-293.

Prigent-Combaret, C., Brombacher, E., Vidal, O., et al. (2001). Complex

regulatory network controls initial adhesion and biofilm formation in

Escherichia coli via regulation of the csgD gene. Journal of

Bacteriology, 183(24), 7213-7223.

350 |

Prigent-Combaret, C., Vidal, O., Dorel, C., & Lejeune, P. (1999). Abiotic

surface sensing and biofilm-dependent regulation of gene expression

in Escherichia coli. Journal of Bacteriology, 181(19), 5993-6002.

Prokesch, B. C., TeKippe, M., Kim, J., et al. (2016). Primary osteomyelitis

caused by hypervirulent Klebsiella pneumoniae. The Lancet Infectious

Diseases, 16(9), e190-e195.

Pruss, B. M., Besemann, C., Denton, A., & Wolfe, A. J. (2006). A complex

transcription network controls the early stages of biofilm development

by Escherichia coli. Journal of Bacteriology, 188(11), 3731-3739.

Pultz, N. J., Stiefel, U., & Donskey, C. J. (2005). Effects of daptomycin,

linezolid, and vancomycin on establishment of intestinal colonization

with vancomycin-resistant enterococci and extended-spectrum-beta-

lactamase-producing Klebsiella pneumoniae in mice. Antimicrobial

Agents and Chemotherapy, 49(8), 3513-3516.

Qi, Y., Pei, J., & Grishin, N. V. (2002). C-terminal domain of gyrase A is

predicted to have a β-propeller structure. Proteins: Structure, Function

and Bioinformatics, 47(3), 258-264.

Rachid, S., Ohlsen, K., Witte, W., Hacker, J., & Ziebuhr, W. (2000). Effect of

subinhibitory antibiotic concentrations on polysaccharide intercellular

adhesin expression in biofilm-forming Staphylococcus epidermidis.

Antimicrobial Agents and Chemotherapy, 44(12), 3357-3363.

Raffi, H. S., Bates, J. M., Flournoy, D. J., & Kumar, S. (2012). Tamm-Horsfall

protein facilitates catheter associated urinary tract infection. BMC

Research Notes, 5(1), 532.

Ramasubbu, N., Thomas, L. M., Ragunath, C., & Kaplan, J. B. (2005).

Structural analysis of dispersin B, a biofilm-releasing glycoside

351 |

hydrolase from the periodontopathogen Actinobacillus

actinomycetemcomitans. Journal of Molecular Biology, 349(3), 475-

486.

Ramirez, M. S., & Tolmasky, M. E. (2010). Aminoglycoside modifying

enzymes. Drug Resistance Updates, 13(6), 151-171.

Rawat, D., & Nair, D. (2010). Extended-spectrum β-lactamases in Gram

negative bacteria. Journal of Global Infectious Diseases, 2(3), 263-274.

Raymond, T., Wiesen, J., Rehm, S., & Auron, M. (2014). Carbapenem-

resistant Klebsiella pneumoniae prosthetic valve endocarditis: A feared

combination of technology and emerging pathogens. Infectious

Diseases in Clinical Practice, 22(2), 113-115.

Redgrave, L. S., Sutton, S. B., Webber, M. A., & Piddock, L. J. V. (2014).

Fluoroquinolone resistance: mechanisms, impact on bacteria, and role

in evolutionary success. Trends in Microbiology, 22(8), 438-445.

Reisner, A., Maierl, M., Jorger, M., et al. (2014). Type 1 fimbriae contribute to

catheter-associated urinary tract infections caused by Escherichia coli.

Journal of Bacteriology, 196(5), 931-939.

Rice, P., Longden, I., & Bleasby, A. (2000). EMBOSS: The European

Molecular Biology Open Software Suite. Trends in Genetics, 16(6), 276-

277.

Richter, L., Stepper, C., Mak, A., et al. (2007). Development of a microfluidic

biochip for online monitoring of fungal biofilm dynamics. Lab on a Chip,

7(12), 1723-1731.

352 |

Robicsek, A., Strahilevitz, J., Jacoby, G. A., et al. (2005). Fluoroquinolone-

modifying enzyme: a new adaptation of a common aminoglycoside

acetyltransferase. Nature Medicine, 12, 83.

Rodrigues, D. F., & Elimelech, M. (2009). Role of type 1 fimbriae and mannose

in the development of Escherichia coli K12 biofilm: from initial cell

adhesion to biofilm formation. Biofouling, 25(5), 401-411.

Rolhion, N., Carvalho, F. A., & Darfeuille-Michaud, A. (2007). OmpC and the

σE regulatory pathway are involved in adhesion and invasion of the

Crohn's disease-associated Escherichia coli strain LF82. Molecular

Microbiology, 63(6), 1684-1700.

Romero, D., Traxler, M. F., Lopez, D., & Kolter, R. (2011). Antibiotics as signal

molecules. Chemical Reviews, 111(9), 5492-5505.

Römling, U. (2012). Cyclic di-GMP, an established secondary messenger still

speeding up. Environmental Microbiology, 14(8), 1817-1829.

Römling, U., & Galperin, M. Y. (2015). Bacterial cellulose biosynthesis:

diversity of operons, subunits, products, and functions. Trends in

Microbiology, 23(9), 545-557.

Ronald, A. (2002). The etiology of urinary tract infection: traditional and

emerging pathogens. The American Journal of Medicine, 113(1), 14-

19.

Rose, A. S., Bradley, A. R., Valasatava, Y., et al. (2016). Web-based molecular

graphics for large complexes. Paper presented at the Proceedings of

the 21st International Conference on Web3D Technology, Anaheim,

California.

353 |

Rose, A. S., & Hildebrand, P. W. (2015). NGL Viewer: a web application for

molecular visualization. Nucleic Acids Research, 43(W1), 576-579.

Rosen, D. A., Pinkner, J. S., Jones, J. M., et al. (2008). Utilization of an

intracellular bacterial community pathway in Klebsiella pneumoniae

urinary tract infection and the effects of FimK on type 1 pilus expression.

Infection and Immunity, 76(7), 3337-3345.

Rosen, D. A., Pinkner, J. S., Walker, J. N., et al. (2008). Molecular variations

in Klebsiella pneumoniae and Escherichia coli FimH affect function and

pathogenesis in the urinary tract. Infection and Immunity, 76(7), 3346-

3356.

Rosenberg, E. Y., Ma, D., & Nikaido, H. (2000). AcrD of Escherichia coli is an

aminoglycoside efflux pump. Journal of Bacteriology, 182(6), 1754-

1756.

Rosenblueth, M., Martinez, L., Silva, J., & Martinez-Romero, E. (2004).

Klebsiella variicola, a novel species with clinical and plant-associated

isolates. Systematic and Applied Microbiology, 27(1), 27-35.

Roux, D., Cywes-Bentley, C., Zhang, Y. F., et al. (2015). Identification of poly-

N-acetylglucosamine as a major polysaccharide component of the

Bacillus subtilis biofilm matrix. Journal of Biological Chemistry, 290(31),

19261-19272.

Roy, S., Datta, S., Viswanathan, R., Singh, A. K., & Basu, S. (2013).

Tigecycline susceptibility in Klebsiella pneumoniae and Escherichia coli

causing neonatal septicaemia (2007-10) and role of an efflux pump in

tigecycline non-susceptibility. Journal of Antimicrobial Chemotherapy,

68(5), 1036-1042.

354 |

Ruiz, E., Sáenz, Y., Zarazaga, M., et al. (2012). qnr, aac(6′)-Ib-cr and qepA

genes in Escherichia coli and Klebsiella spp.: genetic environments and

plasmid and chromosomal location. Journal of Antimicrobial

Chemotherapy, 67(4), 886-897.

Russo, T. A., Olson, R., MacDonald, U., Beanan, J., & Davidson, B. A. (2015).

Aerobactin, but not yersiniabactin, salmochelin, or enterobactin,

enables the growth/survival of hypervirulent (hypermucoviscous)

Klebsiella pneumoniae ex vivo and in vivo. Infection and Immunity,

83(8), 3325-3333.

Saenger, W. (2013). Proteinase K In N. D. Rawlings & G. Salvesen (Eds.),

Handbook of Proteolytic Enzymes (Vol. 1, pp. 3240-3242). Europe:

Elsevier Ltd.

Sahly, H., Keisari, Y., & Ofek, I. (2009). Manno(rhamno)biose-containing

capsular polysaccharides of Klebsiella pneumoniae enhance opsono-

stimulation of human polymorphonuclear leukocytes. Journal of Innate

Immunity, 1(2), 136-144.

Sahly, H., Navon-Venezia, S., Roesler, L., et al. (2008). Extended-spectrum

β-lactamase production is associated with an increase in cell invasion

and expression of fimbrial adhesins in Klebsiella pneumoniae.

Antimicrobial Agents and Chemotherapy, 52(9), 3029-3034.

Sahly, H., Podschun, R., Oelschlaeger, T. A., et al. (2000). Capsule impedes

adhesion to and invasion of epithelial cells by Klebsiella pneumoniae.

Infection and Immunity, 68(12), 6744-6749.

Sakazaki, R., Tamura, K., Kosako, Y., & Yoshizaki, E. (1989). Klebsiella

ornithinolytica sp. nov., formerly known as ornithine-positive Klebsiella

oxytoca. Current Microbiology, 18(4), 201-206.

355 |

Saldaña, Z., A., M., Carrillo-Casas, E. M., et al. (2014). Production of the

Escherichia coli common pilus by uropathogenic E. coli is associated

with adherence to HeLa and HTB-4 cells and invasion of mouse bladder

urothelium. PLOS ONE, 9(7), e101200.

Sánchez, R., & Sali, A. (1998). Large-scale protein structure modeling of the

Saccharomyces cerevisiae genome. Proceedings of the National

Academy of Sciences, 95(23), 13597-13602.

Sandora, T. J., & Goldmann, D. A. (2012). Preventing lethal hospital outbreaks

of antibiotic-resistant bacteria. New England Journal of Medicine,

367(23), 2168-2170.

Saur, T., Morin, E., Habouzit, F., Bernet, N., & Escudié, R. (2017). Impact of

wall shear stress on initial bacterial adhesion in rotating annular reactor.

PLOS ONE, 12(2), e0172113.

Sbrana, F., Malacarne, P., Bassetti, M., et al. (2016). Risk factors for ventilator

associated pneumonia due to carbapenemase-producing Klebsiella

pneumoniae in mechanically ventilated patients with tracheal and rectal

colonization. Minerva Anestesiologica, 82(6), 635-640.

Schembri, M. A., Blom, J., Krogfelt, K. A., & Klemm, P. (2005). Capsule and

fimbria interaction in Klebsiella pneumoniae. Infection and Immunity,

73(8), 4626-4633.

Schembri, M. A., Dalsgaard, D., & Klemm, P. (2004). Capsule shields the

function of short bacterial adhesins. Journal of Bacteriology, 186(5),

1249-1257.

Schmidt, A. J., Ryjenkov, D. A., & Gomelsky, M. (2005). The ubiquitous protein

domain EAL is a cyclic diguanylate-specific phosphodiesterase:

356 |

enzymatically active and inactive EAL domains. Journal of Bacteriology,

187(14), 4774-4781.

Schmieder, R., & Edwards, R. (2012). Insights into antibiotic resistance

through metagenomic approaches. Future Microbiology, 7(1), 73-89.

Schroll, C., Barken, K. B., Krogfelt, K. A., & Struve, C. (2010). Role of type 1

and type 3 fimbriae in Klebsiella pneumoniae biofilm formation. BMC

Microbiology, 10(179), 2-10.

Schulz, G. E. (2004). The structures of general porins. In R. Benz (Ed.),

Bacterial and Eukaryotic Porins; Structure, Function, Mechanism (pp.

25-40). Germany: 2004 Wiley-Vch Verlag GmbH & Co.

Schwartz, K., Stephenson, R., Hernandez, M., Jambang, N., & Boles, B. R.

(2010). The use of drip flow and rotating disk reactors for

Staphylococcus aureus biofilm analysis. Journal of Visualized

Experiments : JoVE(46).

Seltmann, G., & Holst, O. (2002). Proteins of the outer membrane. In The

bacterial cell wall (pp. 89): Springer Science & Business Media.

Seneviratne, C. J., Yip, J. W. Y., Chang, J. W. W., Zhang, C. F., &

Samaranayake, L. P. (2013). Effect of culture media and nutrients on

biofilm growth kinetics of laboratory and clinical strains of Enterococcus

faecalis. Archives of Oral Biology, 58(10), 1327-1334.

Serafini-Cessi, F., Malagolini, N., & Cavallone, D. (2003). Tamm-Horsfall

glycoprotein: biology and clinical relevance. American Journal of Kidney

Diseases, 42(4), 658-676.

357 |

Shakib, P., Ghafourian, S., Zolfaghary, M. R., et al. (2012). Prevalence of

OmpK35 and OmpK36 porin expression in beta-lactamase and non-

betalactamase- producing Klebsiella pneumoniae. Biologics : Targets

& Therapy, 6, 1-4.

Sharples, G. J., & Lloyd, R. G. (1990). A novel repeated DNA sequence

located in the intergenic regions of bacterial chromosomes. Nucleic

Acids Research, 18(22), 6503-6508.

Shoji, S., Walker, S. E., & Fredrick, K. (2009). Ribosomal translocation: One

step closer to the molecular mechanism. ACS Chemical Biology, 4(2),

93-107.

Shon, A. S., Bajwa, R. P., & Russo, T. A. (2013). Hypervirulent

(hypermucoviscous) Klebsiella pneumoniae: A new and dangerous

breed. Virulence, 4(2), 107-118.

Shu, J. C., Chia, J. H., Kuo, A. J., Su, L. H., & Wu, T. L. (2009). A 7-year

surveillance for ESBL-producing Escherichia coli and Klebsiella

pneumoniae at a university hospital in Taiwan: the increase of CTX-M-

15 in the ICU. Epidemiology and Infection, 138(2), 253-263.

Sievers, F., Wilm, A., Dineen, D., et al. (2011). Fast, scalable generation of

high‐quality protein multiple sequence alignments using Clustal

Omega. Molecular Systems Biology, 7(1), 539.

Simpson, I. N., Harper, P. B., & O'Callaghan, C. H. (1980). Principal beta-

lactamases responsible for resistance to beta-lactam antibiotics in

urinary tract infections. Antimicrobial Agents and Chemotherapy, 17(6),

929-936.

358 |

Singla, S., Harjai, K., & Chhibber, S. (2013). Susceptibility of different phases

of biofilm of Klebsiella pneumoniae to three different antibiotics. Journal

of Antibiotics, 66(2), 61-66.

Siu, L. K., Yeh, K. M., Lin, J. C., Fung, C. P., & Chang, F. Y. (2012). Klebsiella

pneumoniae liver abscess: a new invasive syndrome. The Lancet

Infectious Diseases, 12(11), 881-887.

Smith, D. R., Price, J. E., Burby, P. E., et al. (2017). The production of curli

amyloid fibers is deeply integrated into the biology of Escherichia coli.

Biomolecules, 7(4).

Smyth, E. T. M., McIlvenny, G., Enstone, J. E., et al. (2008). Four country

healthcare associated infection prevalence survey 2006: overview of

the results. Journal of Hospital Infection, 69(3), 230-248.

Sokurenko, E. V., Vogel, V., & Thomas, W. E. (2008). Catch bond mechanism

of force-enhanced adhesion: counter-intuitive, elusive but …

widespread? Cell Host & Microbe, 4(4), 314-323.

Solano, C., Garcia, B., Valle, J., et al. (2002). Genetic analysis of Salmonella

enteritidis biofilm formation: critical role of cellulose. Molecular

Microbiology, 43(3), 793-808.

Solovyev, V., & Salamov, A. (2011). Automatic annotation of microbial

genomes and metagenomic sequences. In R. W. Li (Ed.),

Metagenomics and its Applications in Agriculture, Biomedicine and

Environmental Studies (pp. 61-78): Nova Science Publisher, Inc.

Song, T., Duperthuy, M., & Wai, S. N. (2016). Sub-optimal treatment of

bacterial biofilms. Antibiotics, 5(2), 23.

359 |

Sougakoff, W., Goussard, S., & 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

Microbiology Letters, 56(3), 343-348.

Staerk, K., Khandige, S., Kolmos, H. J., Moller-Jensen, J., & Andersen, T. E.

(2016). Uropathogenic Escherichia coli express type 1 fimbriae only in

surface adherent populations under physiological growth conditions.

The Journal of Infectious Diseases, 213(3), 386-394.

Stahlhut, S. G., Struve, C., Krogfelt, K. A., & Reisner, A. (2012). Biofilm

formation of Klebsiella pneumoniae on urethral catheters requires either

type 1 or type 3 fimbriae. FEMS Immunology and Medical Microbiology,

65(2), 350-359.

Stahlhut, S. G., Tchesnokova, V., Struve, C., et al. (2009). Comparative

structure-function analysis of mannose-specific FimH adhesins from

Klebsiella pneumoniae and Escherichia coli. Journal of Bacteriology,

191(21), 6592-6601.

Steimle, A., Autenrieth, I. B., & Frick, J. S. (2016). Structure and function: Lipid

A modifications in commensals and pathogens. International Journal of

Medical Microbiology, 306(5), 290-301.

Stewart, P. S. (2002). Mechanisms of antibiotic resistance in bacterial biofilms.

International Journal of Medical Microbiology, 292(2), 107-113.

Stewart, P. S., & William Costerton, J. (2001). Antibiotic resistance of bacteria

in biofilms. The Lancet, 358(9276), 135-138.

Stickler, D., Young, R., Jones, G., Sabbuba, N., & Morris, N. (2003). Why are

Foley catheters so vulnerable to encrustation and blockage by

crystalline bacterial biofilm? Urological Research, 31(5), 306-311.

360 |

Stoodley, P., Dodds, I., Boyle, J. D., & Lappin-Scott, H. M. (1998). Influence of

hydrodynamics and nutrients on biofilm structure. Journal of Applied

Microbiology, 85(Supplement 1), 19S-28S.

Strahilevitz, J., Jacoby, G. A., Hooper, D. C., & Robicsek, A. (2009). Plasmid-

mediated quinolone resistance: a multifaceted threat. Clinical

Microbiology Reviews, 22(4), 664-689.

Struve, C., Bojer, M., & Krogfelt, K. A. (2008). Characterization of Klebsiella

pneumoniae type 1 fimbriae by detection of phase variation during

colonization and infection and impact on virulence. Infection and

Immunity, 76(9), 4055-4065.

Struve, C., & Krogfelt, K. A. (2004). Pathogenic potential of environmental

Klebsiella pneumoniae isolates. Environmental Microbiology, 6(6), 584-

590.

Sugawara, E., Kojima, S., & Nikaido, H. (2016). Klebsiella pneumoniae major

porins OmpK35 and OmpK36 allow more efficient diffusion of β-lactams

than their Escherichia coli homologs OmpF and OmpC. Journal of

Bacteriology, 198(23), 3200-3208.

Sun, L., Bertelshofer, F., Greiner, G., & Böckmann, R. A. (2016).

Characteristics of sucrose transport through the sucrose-specific porin

ScrY studied by molecular dynamics simulations. Frontiers in

Bioengineering and Biotechnology, 4(9).

Sutcliffe, J., Blumenthal, R., Walter, A., & Fouldsi, J. (1983). Escherichia coli

outer membrane protein K is a porin. Journal of Bacteriology, 156(2),

867-872.

Taglialegna, A., Lasa, I., & Valle, J. (2016). Amyloid structures as biofilm matrix

scaffolds. Journal of Bacteriology, 198(19), 2579-2588.

361 |

Tai, P. C., & Davis, B. D. (1979). Triphasic concentration effects of gentamicin

on activity and misreading in protein synthesis. Biochemistry, 18(1),

193-198.

Tambyah, P. A., & Maki, D. G. (2000). Catheter-associated urinary tract

infection is rarely symptomatic: a prospective study of 1,497

catheterized patients. Archives of Internal Medicine, 160(5), 678-682.

Tangy, F., Moukkadem, M., Vindimian, E., Capmau, M. L., & Goffic, F. (1985).

Mechanism of action of gentamicin components. The FEBS Journal,

147(2), 381-386.

Tenover, F. C. (2006). Mechanisms of antimicrobial resistance in bacteria. The

American Journal of Medicine, 119(6), S3-10.

Tetz, G. V., Artemenko, N. K., & Tetz, V. V. (2009). Effect of DNase and

antibiotics on biofilm characteristics. Antimicrobial Agents and

Chemotherapy, 53(3), 1204-1209.

Tezel, B. U., Akçelik, N., Yüksel, F. N., Karatuğ, N. T., & Akçelik, M. (2016).

Effects of sub-MIC antibiotic concentrations on biofilm production of

Salmonella Infantis. Biotechnology & Biotechnological Equipment,

30(6), 1184-1191.

Thanassi, D. G., Suh, G. S., & Nikaido, H. (1995). Role of outer membrane

barrier in efflux-mediated tetracycline resistance of Escherichia coli.

Journal of Bacteriology, 177(4), 998-1007.

Thomas, B., & Tolley, D. (2008). Concurrent urinary tract infection and stone

disease: pathogenesis, diagnosis and management. Nature Clinical

Practice Urology, 5(12), 668-675.

362 |

Thomas, W. (2008). Catch bonds in adhesion. Annual Review of Biomedical

Engineering, 10(1), 39-57.

Thornton, M. M., Chung-Esaki, H. M., Irvin, C. B., et al. (2012). Multicellularity

and antibiotic resistance in Klebsiella pneumoniae grown under

bloodstream-mimicking fluid dynamic conditions. Journal of Infectious

Diseases, 206(4), 588-595.

Tran, J. H., & Jacoby, G. A. (2002). Mechanism of plasmid-mediated quinolone

resistance. Proceedings of the National Academy of Sciences, 99(8),

5638-5642.

Tran, J. H., Jacoby, G. A., & Hooper, D. C. (2005). Interaction of the plasmid-

encoded quinolone resistance protein Qnr with Escherichia coli DNA

gyrase. Antimicrobial Agents and Chemotherapy, 49(1), 118-125.

Tsai, Y. K., Fung, C. P., Lin, J. C., et al. (2011). Klebsiella pneumoniae outer

membrane porins OmpK35 and OmpK36 play roles in both

antimicrobial resistance and virulence. Antimicrobial Agents and

Chemotherapy, 55(4), 1485-1493.

Tsao, N., Kuo, C. F., Chiu, C. C., et al. (2015). Protection against Klebsiella

pneumoniae using lithium chloride in an intragastric infection model.

Antimicrob Agents Chemother, 59(3), 1525-1533.

Tsuji, A., Kaneko, Y., Takahashi, K., Ogawa, M., & Goto, S. (1982). The effects

of temperature and pH on the growth of eight enteric and nine glucose

non-fermenting species of gram-negative rods. Microbiology and

Immunology, 26(1), 15-24.

Turner, P. J. (2005). Extended-spectrum β-lactamases. Clinical Infectious

Diseases, 41(Supplement 4), S273 - S275.

363 |

Tzouvelekis, L. S., Tzelepi, E., Tassios, P. T., & Legakis, N. J. (2000). CTX-

M-type beta-lactamases: an emerging group of extended-spectrum

enzymes. International Journal of Antimicrobial Agents, 14(2), 137-

142.

van Duin, D., & Bonomo, R. A. (2016). Ceftazidime/avibactam and

ceftolozane/tazobactam: Second-generation β-lactam/β-lactamase

inhibitor combinations. Clinical Infectious Diseases, 63(2), 234-241.

van Hoek, A. H., Mevius, D., Guerra, B., et al. (2011). Acquired antibiotic

resistance genes: an overview. Frontiers in Microbiology, 2, 1-27.

Vetsch, M., Puorger, C., Spirig, T., et al. (2004). Pilus chaperones represent a

new type of protein-folding catalyst. Nature, 431(7006), 329-333.

Vinogradov, E., Frirdich, E., MacLean, L. L., et al. (2002). Structures of

lipopolysaccharides from Klebsiella pneumoniae. Elucidation of the

structure of the linkage region between core and polysaccharide O

chain and identification of the residues at the non-reducing termini of

the O chains. The Journal of Biological Chemistry, 277(28), 25070-

25081.

Virgilio, E., Castaldo, P., Catta, F., Tarantino, G., & Cavallini, M. (2016).

Abdominal surgical site infection due to Klebsiella pneumoniae

carbapenemase‐producing K. pneumoniae. International Wound

Journal, 13(5), 1075-1076.

Vuong, C., Kocianova, S., Voyich, J. M., et al. (2004). A crucial role for

exopolysaccharide modification in bacterial biofilm formation, immune

evasion, and virulence. Journal of Biological Chemistry, 279(52),

54881-54886.

364 |

Vuotto, C., Longo, F., Balice, M. P., Donelli, G., & Varaldo, P. E. (2014).

Antibiotic resistance related to biofilm formation in Klebsiella

pneumoniae. Pathogens, 3(3), 743-758.

Vuotto, C., Longo, F., Pascolini, C., et al. (2017). Biofilm formation and

antibiotic resistance in Klebsiella pneumoniae urinary strains. Journal

of Applied Microbiology, 123(4), 1003-1018.

Walthers, D., Go, A., & Kenney, L. J. (2004). Regulation of porin gene

expression by the two-component regulatory system EnvZ/OmpR. In R.

Benz (Ed.), Bacterial and Eukaryotic Porins: Structure, Function,

Mechanism (pp. 1-24). Germany: 2004 Wiley-Vch Verlag GmbH & Co.

Wang, H., Yan, Y., Rong, D., et al. (2016). Increased biofilm formation ability

in Klebsiella pneumoniae after short-term exposure to a simulated

microgravity environment. MicrobiologyOpen, 5(5), 793-801.

Wang, L. C., Morgan, L. K., Godakumbura, P., Kenney, L. J., & Anand, G. S.

(2012). The inner membrane histidine kinase EnvZ senses osmolality

via helix‐coil transitions in the cytoplasm. The EMBO Journal, 31(11),

2648-2659.

Wang, X., Preston, J. F., & Romeo, T. (2004). The pgaABCD locus of

Escherichia coli promotes the synthesis of a polysaccharide adhesin

required for biofilm formation. Journal of Bacteriology, 186(9), 2724-

2734.

Wang, X. D., Cai, J. C., Zhou, H. W., Zhang, R., & Chen, G. X. (2009).

Reduced susceptibility to carbapenems in Klebsiella pneumoniae

clinical isolates associated with plasmid-mediated β-lactamase

production and OmpK36 porin deficiency. Journal of Medical

Microbiology, 58(9), 1196-2202.

365 |

Wang, Z. C., Huang, C. J., Huang, Y. J., Wu, C. C., & Peng, H. L. (2013). FimK

regulation on the expression of type 1 fimbriae in Klebsiella pneumoniae

CG43S3. Microbiology, 159(7), 1402-1415.

Webber, M. A., & Piddock, L. J. V. (2003). The importance of efflux pumps in

bacterial antibiotic resistance. Journal of Antimicrobial Chemotherapy,

51(1), 9-11.

Wei, D., Sun, J., Shi, J., Liu, P., & Hao, J. (2013). New strategy to improve

efficiency for gene replacement in Klebsiella pneumoniae. Journal of

Industrial Microbiology & Biotechnology, 40(5), 523-527.

Wei, D., Wang, M., Shi, J., & Hao, J. (2012). Red recombinase assisted gene

replacement in Klebsiella pneumoniae. Journal of Industrial

Microbiology & Biotechnology, 39(8), 1219-1226.

Weinstein, R. A., & Hooper, D. C. (2005). Efflux pumps and nosocomial

antibiotic resistance: A primer for hospital epidemiologists. Clinical

Infectious Diseases, 40(12), 1811-1817.

Whitchurch, C. B., Tolker-Nielsen, T., Ragas, P. C., & Mattick, J. S. (2002).

Extracellular DNA required for bacterial biofilm formation. Science,

295(5559), 1487.

Wiebe, H., Gurlebeck, D., Gross, J., et al. (2015). YjjQ represses transcription

of flhDC and additional loci in Escherichia coli. Journal of Bacteriology,

197(16), 2713-2720.

Wiegand, I., Hilpert, K., & Hancock, R. E. W. (2008). Agar and broth dilution

methods to determine the minimal inhibitory concentration (MIC) of

antimicrobial substances. Nature Protocols, 3(2), 163-175.

366 |

Wilkinson, M., Troman, L., Ismah, W. A. W. N., et al. (2016). Structural basis

for the inhibition of RecBCD by Gam and its synergistic antibacterial

effect with quinolones. eLife, 5, e22963.

Wilson, D. N. (2014). Ribosome-targeting antibiotics and mechanisms of

bacterial resistance. Nature Reviews Microbiology, 12(1), 35-48.

Wilson, M. L., & Gaido, L. (2004). Laboratory diagnosis of urinary tract

infections in adult patients. Clinical Infectious Diseases, 38(8), 1150-

1158.

Wimpenny, J., Manz, W., & Szewzyk, U. (2000). Heterogeneity in biofilms.

FEMS Microbiology Reviews, 24(5), 661-671.

Wolfaardt, G. M., Lawrence, J. R., Robarts, R. D., Caldwell, S. J., & Caldwell,

D. E. (1994). Multicellular organization in a degradative biofilm

community. Applied and Environmental Microbiology, 60(2), 434-446.

Wu, C. C., Huang, Y. J., Fung, C. P., & Peng, H. L. (2010). Regulation of the

Klebsiella pneumoniae Kpc fimbriae by the site-specific recombinase

KpcI. Microbiology, 156(7), 1983-1992.

Wu, M. C., Lin, T. L., Hsieh, P. F., Yang, H. C., & Wang, J. T. (2011). Isolation

of genes involved in biofilm formation of a Klebsiella pneumoniae strain

causing pyogenic liver abscess. PLOS ONE, 6(8), e23500.

Wu, X. R., Sun, T. T., & Medina, J. J. (1996). In vitro binding of type 1-

fimbriated Escherichia coli to uroplakins Ia and Ib: relation to urinary

tract infections. Proceedings of the National Academy of Sciences,

93(18), 9630-9635.

367 |

Yanisch-Perron, C., Vieira, J., & Messing, J. (1985). Improved M13 phage

cloning vectors and host strains: nucleotide sequences of the M13mp18

and pUC19 vectors. Gene, 33(1), 103-119.

Yigit, H., Queenan, A. M., Anderson, G. J., et al. (2001). Novel carbapenem-

hydrolyzing β-lactamase, KPC-1, from a carbapenem-resistant strain of

Klebsiella pneumoniae. Antimicrobial Agents and Chemotherapy,

45(4), 1151-1161.

Yinnon, A. M., Butnaru, A., Raveh, D., Jerassy, Z., & Rudensky, B. (1996).

Klebsiella bacteraemia: community versus nosocomial infection. QJM :

An International Journal of Medicine, 89(12), 933-941.

Yong, D., Toleman, M. A., Giske, C. G., et al. (2009). Characterization of a

new metallo-beta-lactamase gene, bla(NDM-1), and a novel

erythromycin esterase gene carried on a unique genetic structure in

Klebsiella pneumoniae sequence type 14 from India. Antimicrobial

Agents and Chemotherapy, 53(12), 5046-5054.

Zhang, Y., Li, P., Yin, Y., Li, F., & Zhang, Q. (2017). In vitro activity of

tigecycline in combination with rifampin, doripenem or ceftazidime

against carbapenem-resistant Klebsiella pneumoniae bloodstream

isolates. The Journal of Antibiotics, 70(2), 193-195.

Zhao, W.-H., & Hu, Z.-Q. (2013). Epidemiology and genetics of CTX-M

extended-spectrum β-lactamases in Gram-negative bacteria. Critical

Reviews in Microbiology, 39(1), 79-101.

Zheng, J.-x., Lin, Z.-w., Chen, C., et al. (2018). Biofilm formation in Klebsiella

pneumoniae bacteremia strains was found to be associated with CC23

and the presence of wcaG. Frontiers in Cellular and Infection

Microbiology, 8, 21.

368 |

Zhou, A., Kang, T. M., Yuan, J., et al. (2015). Synergistic interactions of

vancomycin with different antibiotics against Escherichia coli:

trimethoprim and nitrofurantoin display strong synergies with

vancomycin against wild-type E. coli. Antimicrobial Agents and

Chemotherapy, 59(1), 276-281.

Zogaj, X., Nimtz, M., Rohde, M., Bokranz, W., & Romling, U. (2001). The

multicellular morphotypes of Salmonella typhimurium and Escherichia

coli produce cellulose as the second component of the extracellular

matrix. Molecular Microbiology, 39(6), 1452-1463.

Zweig, M., Schork, S., Koerdt, A., et al. (2014). Secreted single-stranded DNA

is involved in the initial phase of biofilm formation by Neisseria

gonorrhoeae. Environmental Microbiology, 16(4), 1040-1052.

369 |

Appendices

370 |

Appendix 1: Periodate oxidation of PNAG monomer. Figure shows a periodate cleavage of a 2,3-diol on PNAG monomer to give a dicarbonyl

compound. Figure is drawn with ChemDraw Professional 16.0.1 (with the assistance of Hanim Salami Mohd Saupi, Imperial College London.)

371 |

Appendix 2: Possible reaction mechanism between crystal violet compound and oxidised PNAG monomer that could result to the high CV

absorbance reading upon CV staining of the treated pre-formed biofilm with NaIO4. Figure is drawn with ChemDraw Professional 16.0.1 (with

the assistance of Dr Shikh Mohd Shahrul Nizan Shikh Zahari, USIM, Malaysia).

372 |

Appendix 3: Amino acid variations identified in FimB, FimA, FimD and FimF.

Several amino acid variations were identified upon multiple sequence alignment of

amino acid sequences of the strains used in this study. The amino acid variations

were found in FimB at position Q15P for Kpneu_1 and FimA at position T32A for

Kpneu_65. Several amino acid variations were also identified in FimD at position

T131A for Kpneu_65 and C3091, H160Q and V392A for CG43S3, and also P830A

for Kpneu_1, Kpneu_65, and C3091. In addition, two amino acid variations were

identified in FimF at position A10T for Kpneu_65 and C3091, and T169I for Kpneu_1.

373 |

Appendix 4: Intergenic region sequence of fimB and fimE.

Figure shows the sequence obtained from the multiple sequence alignments of the intergenic region of fimB and fimE of all of the strains used in

this study. Several SNPs were identified in the intergenic region of the ‘ON’ and OFF’ group (Kpneu_1, Kpneu_65, C3091, and CG43S3), and

the ‘OFF’ group (Kpneu_64, Kpneu_64, and Kpneu_110). The SNPs are boxed in blue and coloured in red. The start codon of fimE (GTG) is

underlined. The location of the SNPs is based on the fimE start codon and is presented in the table below.

374 |

Appendix 5: Multiple sequence alignment of an intergenic region of fimB and fimE.

The SNPs specific for ‘ON and OFF’ (Kpneu_1, Kpneu_65, C3091, and CG43S3) and ‘OFF’ (Kpneu_63, Kpneu_64, and Kpneu_110) groups are

boxed. The start codon of fimE (GTG) and the predicted promoter by BPROM are underlined and labelled. The ‘ON and OFF’ and ‘OFF’ groups

are as indicated.

375 |

Appendix 6: SNPs identified in the intergenic region of fimB and fimE.

Intergenic region between fimB and fimE

The location of SNPs

upstream from the start

codon of fimE

Kpneu_1

Kpneu_63

Kpneu_64

Kpneu_65

Kpneu_110

C3091

CG43S3

-61 T C C T C T T

-63 A T T A T A A

-204 A G G A G A A

-247 G A A G A G G

-257 A G G A G A A

-265 T G G T G T T

376 |

Appendix 7: Intergenic region sequence of fimE and fimA.

Figure shows the sequence obtained from the multiple sequence alignments of fimS of all of the strains used in this study. Several SNPs were

identified in the intergenic region of the ‘ON’ and OFF’ group (Kpneu_1, Kpneu_65, C3091, and CG43S3), and the ‘OFF’ group (Kpneu_64,

Kpneu_64, and Kpneu_110). The SNPs in the fimS are boxed in blue and coloured in purple. The start codon (ATG) of fimA is underlined. The

location of the SNPs is based on the fimA start codon and is presented in the table below.

377 |

Appendix 8: Multiple sequence alignments of fimS.

The SNPs specific for ‘ON and OFF’ (Kpneu_1, Kpneu_65, C3091, and CG43S3) and ‘OFF’ (Kpneu_63, Kpneu_64, and Kpneu_110) groups are

boxed. The start codon of fimA (ATG), the inverted repeat (IR), the HinfI site and the predicted promoter by BPROM are underlined and labelled.

The ‘ON and OFF’ and ‘OFF’ groups are as indicated.

378 |

Appendix 9: SNPs identified in the intergenic region of fimE and fimA.

*There was a 1 bp deletion identified for strain Kpneu_63 and Kpneu_64 at -140 nucleotide site resulted in 1 bp different of SNPs location in

comparison to the rest of the strains.

Intergenic region of fimE and fimA

The location of SNPs

upstream from the start

codon of fimA

Kpneu_1

Kpneu_63

Kpneu_64

Kpneu_65

Kpneu_110

C3091

CG43S3

-321

A

A

C

A

A

-320*

C

C

-204

A

A

C

A

A

-203*

C

C

379 |

Appendix 10: Multiple sequence alignment of FimF.

The multiple amino acid sequence alignment of FimF was conducted using GeneDoc. The amino acid substitutions resulted from the SNPs are

boxed in red. The ‘ON and OFF’ and ‘OFF’ groups are as indicated.

380 |

Appendix 11: Multiple sequence alignment of FimI.

The multiple amino acid sequence alignment of FimI was conducted using GeneDoc. The amino acid substitutions resulted from the SNPs are

boxed in red. The ‘ON and OFF’ and ‘OFF’ groups are as indicated.

381 |

Appendix 12: An ompK35 replacement with add9 in Kpneu_1.

Schematic diagram of homologous recombination is shown between the Kpneu_1 wild type and its isogenic mutant, Kpneu_1∆ompK35.

Verification of ompK35 mutants can be seen by the allelic exchange occurred between targeted gene, ompK35 and aad9 gene. The upper

diagram indicated the downstream region of wild type and its mutant while the bottom diagram indicated the upstream region of wild type and its

mutant. The HindIII sites and RBS are lined in blue and stop and start codons of ompK35 and aad9 are lined in red.

382 |

Appendix 13: An ompK36 replacement with aad9 in Kpneu_1.

Schematic diagram of homologous recombination is shown between the Kpneu_1 wild type and its isogenic mutant, Kpneu_1∆ompK36.

Verification of ompK36 mutants can be seen by the allelic exchange occurred between targeted gene, ompK36 and aad9 gene. The upper

diagram indicated the downstream region of wild type and its mutant while the bottom diagram indicated the upstream region of wild type and its

mutant. The HindIII sites and RBS are lined in blue and stop and start codons of ompK35 and aad9 are lined in red.

383 |

Appendix 14: 3D structures of OmpC of E. coli.

The figures show the 3D structures of OmpC of E. coli (PDB ID 2J1N). A) Top view of trimeric structure, B) Top view of monomeric structure

C) Side view of trimeric structure, and D) Side view of monomeric structure. The top and side view of the trimeric and monomeric structures were

diagrammed using Chimera 1.12rc. The β-barrel structures are coloured in light blue, the α-helices are denoted in blue, the L3 restriction loop is

coloured in red, and the ligands (dodecane, magnesium and chloride) are coloured in yellow.

384 |

Appendix 15: 3D structures of OmpK36 of K. pneumoniae.

The figures show the 3D structures of OmpK36 of K. pneumoniae (PDB ID 1OSM). A) Top view of trimeric structure, B) Top view of monomeric

structure, C) Side view of trimeric structure, and D) Side view of monomeric structure. The top and side view of the trimeric and monomeric

structures were diagrammed using Chimera 1.12rc. The β-barrel structures are coloured in light blue, the α-helices are denoted in blue, the L3

restriction loop is coloured in red, and the ligands (dodecane) are coloured in yellow.

385 |

Appendix 16: Predicted 3D structures of OmpK35 of K. pneumoniae.

The figure shows the predicted 3D structure of OmpK35. A) Top view of trimeric structure, B) Top view of monomeric structure, C) Side view of

trimeric structure, and D) Side view of monomeric structure. The 3D structure was predicted using 1OSM as a template by SWISS-MODEL. The

top and side view of the trimeric and monomeric structures were diagrammed using Chimera 1.12rc. The predicted β-barrel structures are coloured

in pink, the predicted α-helices are denoted in blue, and the predicted L3 restriction loop is coloured in red.

386 |

Appendix 17: Structural alignment of OmpC and OmpK36.

The figure shows the structural alignment obtained from the superimposition of the

3D models of OmpC of E. coli as a reference/template and OmpK36 of K. pneumoniae

as a target. The non-superimposable regions are highlighted in green. Secondary

structures elements are denoted in blue (α-helices). The L3 restriction loops are

coloured in red. The structural alignment was carried out using UCSF Chimera

1.12rc. The amino acid identity is 80.12 % and the RMSD is 0.841 Å. The differences

between OmpK36 and OmpC structure conformations are found in L2, L4, L5, L7 and

B8 (the details of the amino acid sequence can be seen in Figure 6-13).

387 |

Appendix 18: Structural alignment of OmpK36 and predicted OmpK35.

The figure shows the structural alignment obtained from the superimposition of the

3D models of OmpK36 of K. pneumoniae as a reference/template and putative 3D

structure of OmpK35 as a target. The non-superimposable regions are highlighted in

green. Secondary structures elements are denoted in blue (α-helices). The L3

restriction loops are coloured in red. The structural alignment was carried out using

UCSF Chimera 1.12rc. The amino acid identity is 59.35 % and the RMSD is 0.433 Å.

The differences between OmpK36 and putative OmpK35 structure conformations are

found in L1, L2, L3, L4, L5 and L6 (the details of the amino acid sequence can be

seen in Figure 6-13).