Sunteți pe pagina 1din 87

THE EFFECT OF ANTIMICROBIAL

PEPTIDES ON BACTERIAL BIOFILMS


Focusing on prevention of biofilm formation of urinary tract infection isolates
by the antimicrobial peptides 1037 and LL-37

Andreas Skovgrd Jacobsen


Department of Science, Systems and Models, Roskilde University, Denmark
June 2013

Masters thesis

Supervisors:
Associate Professor Hvard Jenssen (Roskilde University)
Professor Karen Angeliki Krogfelt (Statens Serum Institut)

Abstract
The ongoing development of antibiotic resistant infections is a major obstacle in ensuring the future
health and wellbeing. Even today, many people die from infections, which are caused by hospitalacquired multidrug resistant bacteria. This development predicts an imminent inadequacy of
applicable antibiotics. Biofilms are bacteria that stick together, forming a community, which is
embedded within a self-produced matrix. In urinary tract infections, this way of life is very
common. The formation of a biofilm creates resistance and leads to the use of even more
antibiotics. Since an overly and extensive use of antibiotics have adverse effects, such as inducible
antibiotic resistance, there is a major demand for new antibiotics.
Antimicrobial and host-defense peptides have been proposed as a new anti-infective therapeutic
strategy. The research of antimicrobial peptides has primarily two motives: understanding the host
defense and developing novel antibiotics to fight infections in an increasingly antibiotic resistant
world.
This masters thesis focuses on the only human cathelicidin, LL-37, which has a major role in
human host defence. LL-37, as well as the novel peptide 1037, possesses strong antibiofilm effects
at sub-MIC concentrations.
The results show that both LL-37 and 1037 kills clinical strains isolated from the urinary tract. Both
tobramycin and tetracycline induce P. aeruginosa biofilm formation close to MIC. LL-37 and 1037
are both active against P. aeruginosa biofilms at 2 g/ml, inhibiting the biofilm formation 21.9 %
and 28.2 %, respectively.
LL-37 and 1037 only inhibited K. pneumoniae strain C3091 attachment after two hours, but for LL37, the decrease of attachment was indicated to be due to less bacterial growth. Interestingly, 1037
did not inhibit K. pneumoniae growth at 4x MIC.
The effects on biofilms of two E. coli urinary tract isolates were inconsistent. LL-37 was effective
against E. coli CFT073 at 2 g/ml inhibiting the biofilm formation 29.3 %, and 1037 was effective
against E. coli 536 at 2 g/ml inhibiting the biofilm formation 26.9 %.
The experiments indicate that LL-37 and 1037 possess antibiofilm activity towards urinary tract
isolates.

Danish summary (Dansk sammendrag)


Den igangvrende udvikling af antibiotikaresistente infektioner er en stor forhindring for at
garantere fremtidens helbred og velvre. Selv i dag dr mange mennesker af hospitalsinfektioner
forrsaget af multiresistente bakterier. Denne udvikling varsler om en kommende mangel af
brugbare antibiotika. Biofilme er sammenklbede bakterier som har dannet et samfund, indkapslet i
en selvproduceret matrix. I urinvejsinfektioner er denne levemde almindelig. Dannelsen af en
biofilm skaber resistens og frer til endnu mere brug af antibiotika. Da et overdrevent og
omfattende forbrug af antibiotika har bivirkninger, ssom induceret antibiotikaresistens, er der en
stor eftersprgsel af nye antibiotika.
Antimikrobielle og vrtsforsvars peptider er foreslet som en ny antiinfektions terapeutisk strategi.
Forskningen af antimikrobielle peptider har primrt to motiver: forst vrtsforsvaret og udvikle
nye antibiotika til at bekmpe infektioner i en stigende antibiotisk resistent verden.
Dette speciale fokuserer p det eneste menneske-cathelicidin, LL-37, hvilket har en strre rolle i
menneskets immunforsvar. LL-37, samt det nye peptid, 1037, har vist at have antibiofilm effekter
ved sub-MIC koncentrationer.
Resultaterne viser at bde LL-37 og 1037 drber kliniske stammer isoleret fra urinvejsinfektioner.
Bde tobramycin og tetracycline inducerer P. aeruginosa biofilmdannelsen tt p MIC. LL-37 og
1037 er begge aktive mod P. aeruginosa biofilme ved 2 g/ml, hvor de inhiberer biofilmdannelsen
henholdsvis 21.9 % og 28.2 %.
LL-37 og 1037 inhiberede kun K. pneumoniae stamme C3091 biofilme efter to timer, men for LL37, blev mindskelsen af biofilm indikeret til at skyldes mindre bakterievkst. Interessant nok
inhiberede LL-37 ikke K. pneumoniae vkst ved 4x MIC
Effekterne p biofilme af to E. coli isolater fra urinvejsinfektioner var inkonsekvente. LL-37 var
effektiv mod E. coli CFT073 ved 2 g/ml og inhiberede biofilmdannelsen 29.3 %, og 1037 var
effektiv mod E. coli 536 ved 2 g/ml hvor biofilmdannelsen blev inhiberet med 26.9 %.
Disse eksperimenter indikerer, at LL-37 og 1037 har en antibiofilm effekt mod isolater fra
urinvejsinfektioner.

Preface
This thesis was written as a 60 ECTS point interdisciplinary masters thesis in medical and
molecular biology. All experiments were conducted at Roskilde University in the laboratory of
Hvard Jenssen, Anders Lbner-Olesen and Ole Skovgrd.
My supervisors for this project were Associate Professor Hvard Jenssen, Roskilde University, and
Professor Karen Angeliki Krogfelt, Statens Serum Institut.

Since high school I have been interested in antibiotic resistance. I was introduced to antimicrobial
peptides at Roskilde University, where I did a project on plectasin. I was also caught by the fact that
we are not using antimicrobial peptides therapeutically. Further investigation has showed me how
important antimicrobial peptides can be for human health, and I see the great opportunity in
optimizing the peptides and prepare them for therapeutic use.

I would like to express my appreciation and sincere gratitude to my supervisors Associate Professor
Hvard Jenssen and Professor Karen Angeliki Krogfelt.
Hvard Jenssen introduced me to the world of antimicrobial peptides and I am very grateful for this.
I would also like to thank Karen Angeliki Krogfelt for being an inspiration and introducing the
world of microbes and giving my project a clinical aspect.

I also want to thank the laboratory technicians Christa Persson and Kirsten Olesen from Roskilde
University, who advised and assisted me in the laboratory.
The project would not have been the same without the presence of the many wonderful people of
Hvard Jenssen, Anders Lbner-Olesen and Ole Skovgrds laboratory, Troels Godballe, Biljana
Mojsoska, Godefroid Charbon, Jakob Frimodt-Mller, Louise Bjrn and Andreas Thymann.

List of Abbreviations
ACN

Acetonitrile

AHL

Acyl homoserine lactones

AMP

Antimicrobial peptide

BM2

Basal medium 2

BSA

Bovine serum albumin

CFU

Colony-forming unit

DMEM

Dulbecco's Modified Eagle Medium

EPS

Extracellular polymeric substance

ESI

Electrospray ionization

hCAP

Human cationic antimicrobial protein

HPLC

High-pressure liquid chromatography

LB

Lysogeny broth

LPS

Lipopolysaccharide

MH

Mueller Hinton

MIC

Minimum inhibitory concentration

MRSA

Methicillin-resistant Staphylococcus aureus

MS

Mass spectrometry

OD

Optical density

RNA

Ribonucleic acid

TFA

Trifluoroacetic acid

Microgram

Micromolar

UTI

Urinary tract infection

UV

Ultraviolet

VRSA

Vancomycin-resistant Staphylococcus aureus

Table of content
Content
Abstract ................................................................................................................................................ 3
Danish summary (Dansk sammendrag) ............................................................................................... 4
Preface .................................................................................................................................................. 5
List of Abbreviations ........................................................................................................................... 6
Table of content ................................................................................................................................... 8
1.

Motivation and aim .................................................................................................................... 10

2.

Introduction ................................................................................................................................ 12

3.

2.1

Antibiotics ........................................................................................................................... 12

2.2

Bacterial antibiotic resistance .............................................................................................. 14

2.3

Bacterial biofilms ................................................................................................................ 17

2.4

P. aeruginosa biofilms ........................................................................................................ 18

2.5

Quorum sensing controls biofilm formation ....................................................................... 20

2.6

Biofilm inhibitors as potential drugs ................................................................................... 21

2.7

Urinary Tract infections ...................................................................................................... 22

2.8

Antimicrobial peptides ........................................................................................................ 23

2.9

Human cathelicidin LL-37 .................................................................................................. 26

2.10

Antibiofilm activity of LL-37 .......................................................................................... 30

2.11

The potential of LL-37 and 1037 as UTI inhibitors ........................................................ 31

2.12

Structure activity relationship analysis ............................................................................ 32

Materials and methods ............................................................................................................... 34


3.1

Bacterial strains ................................................................................................................... 34

3.2

Bacterial growth conditions................................................................................................. 34

3.3

Peptides and antibiotics ....................................................................................................... 34

3.4

Purification of LL-37 with HPLC ....................................................................................... 35

4.

3.5

Identification of LL-37 with HPLC/MS.............................................................................. 35

3.6

MIC determination .............................................................................................................. 36

3.7

Biofilm experiments ............................................................................................................ 36

3.8

Time-kill kinetics ................................................................................................................ 37

3.9

Statistical analysis ............................................................................................................... 38

Results ........................................................................................................................................ 39
4.1

HPLC purification of LL-37................................................................................................ 39

4.2

MS identification of peptides .............................................................................................. 40

4.3

MIC determination of antibiotics, LL-37 and LL-37 12-residue fragments ....................... 41

4.4

MIC determination of 1037 and synthetic analogues .......................................................... 43

4.5

Effect of LL-37, 1037 and antibiotics on P. aeruginosa biofilms ...................................... 45

4.6

Effect of LL-37 and 1037 on K. pneumoniae biofilms ....................................................... 48

4.7

Effect of LL-37 and 1037 on E. coli biofilms ..................................................................... 52

5.

Discussion .................................................................................................................................. 53

6.

Conclusion ................................................................................................................................. 65

7.

Future perspectives .................................................................................................................... 66

8.

References .................................................................................................................................. 67

Appendix I.......................................................................................................................................... 80
Effect of LL-37, 1037 and antibiotics on P. aeruginosa biofilms ................................................. 80
Effect of LL-37 and 1037 on K. pneumoniae biofilms .................................................................. 83
Effect of LL-37 and 1037 on E. coli biofilms ................................................................................ 87

1. Motivation and aim


Urinary tract infections (UTI) are very common, especially in young women. Urinary catheters are
commonly used in hospital and health care facilities and their use make UTI the most common
infection acquired in both hospitals and health care facilities. With long term urinary
catheterization, the presence of bacteria in the urine, bacteriuria, is developed in all patients and 25
% for patients who have catheters placed from 2 to 10 days (Saint, 2000).
These infections can in severe cases lead to death after developing pyelonephritis, urinary stones or
perinephric abscesses, yet the contribution of catheter-associated UTI to mortality is unclear
(Warren, 2001). Due to the time correlation and infection rate, it has been suggested to limit the use
of urinary catheters (Jain et al., 1995). Additional hospital days associated with these infections
have great economic and physical costs and thus it is important that we understand how these types
of infections develop, how we prevent and cure them. This is thought to be achieved by coating the
urinary catheters with antimicrobials. For example, silver coating of catheters has been
commercialized for many years (Johnson et al., 1990, Karchmer et al., 2000).
A wide variety of infection organisms are isolated from UTIs, but Escherichia coli, stands out as the
far most frequent infecting organism. Klebsiella pneumoniae are also a common infecting organism
(Nicolle, 2005). A catheter serves as an eminent surface for bacterial attachment, which then
connects the bacteria with the lumen of the urethra. Both E. coli and K. pneumoniae can attach to
the catheter and form a subsequent biofilm.
A biofilm often complicates the clearance of infections. This is experienced with the lung disease
cystic fibrosis which causes many (often fatal) Pseudomonas aeruginosa infections. These types of
infections are eradicated by intensive antibiotic therapies, but these antibiotics affect the bacteria
itself and are not directed towards biofilms. Biofilm also prevents the antibiotics and immune
defense from acting on the bacteria. Not only can it be difficult to cure biofilm infections with
antibiotics, but the often excessive amount of antibiotics used for treatment will eventually increase
the chances of developing bacterial antibiotic resistance, and there are often severe side effects
associated with the use of antibiotics.

Antibiotic resistant strains are often isolated from UTI. One of these strains is E. coli sequence type
ST131, which was first reported in 2008 (Johnson et al., 2012, Nicolas-Chanoine et al., 2008).
Since the discovery of this resistant strain, it has spread and continued to adapt new types of
antibiotic resistance and is increasing among older adults and residents of nursing homes (Banerjee

10

et al., 2013). This example of resistance development is dangerous and threatens how we will be
able to treat urinary tract infections in the future, and it calls for new strategies focusing on
preventing and treating UTIs. There are two major approaches to solving this problem: to improve
or discover new antibiotics and to develop urinary catheters that cannot serve as a link to infection.

Antimicrobial peptides (AMP) are an evolutionarily conserved component of the immune defense
in most complex organisms. Many AMPs have been discovered during the last two decades, but in
spite of the lack of discovery of new antibiotics (Figure 1, p.12) no newly discovered AMP has
passed clinical trials. But peptide drugs are becoming more and more popular in the pharmaceutical
industry and currently, there are 60-70 approved peptide drugs on the market and it is assumed that
100-200 peptide drugs are in clinical trials (Sun, 2013).
AMPs have been tried immobilized to biomaterial surfaces to lower side effects and decomposition
and they have been clinically trialed as peptide antibiotics to kill multi drug resistant strains (Costa
et al., 2011, Mygind et al., 2005). Considering their effects against microbes and with the
experience with peptides in medication, peptides are highly qualified as new antibiotics.

This masters thesis will focus on two AMPs: a highly conserved human AMP, LL-37, and the
synthetic AMP, 1037, which are both antimicrobial and antibiofilm against P. aeruginosa. P.
aeruginosa is in this study used for evaluation of the antimicrobial potency of the peptides and
serves as a model biofilm producing organism for further evaluation of antibiofilm effect of LL-37
and 1037.
This antibiofilm effect will finally be examined on a number of UTI pathogens.
The aims for this masters thesis are as listed:
1. Determining the antimicrobial potential of 8 LL-37 derivatives and for 1037, and studying
the effect of single residue substitution on antimicrobial activity.
2. Confirming and expanding the knowledge of the antimicrobial and antibiofilm activity of
LL-37 and 1037 on P. aeruginosa.
3. Studying the effects on biofilm growth of LL-37 and 1037 on clinical E. coli and K.
pneumoniae UTI strains

11

2. Introduction
The following chapters will explain the historical aspects of antibiotics and how the use of
antibiotics consequently has induced antibiotic resistance. Following, bacterial biofilms, quorumsensing and urinary tract infections chapters will argue why we need more and different working
antibiotics. Then there will be an introduction to antimicrobial peptides and the two antimicrobial
peptides, LL-37 and 1037. The introduction will finally provide an introduction and overview of the
conducted experiments.

2.1 Antibiotics
Since the 1940s, bacterial infections have become a less considerable problem and many infections
which today easily can be treated, would back then often result in death. The decreased number of
deaths from bacterial infections is most importantly due to the discovery of antibiotics.
With the discovery of the first commercially available antibacterial drugs, Salvarsan (1910) by Paul
Ehrlich, the sulfonamide Prontosil (1935) by Gerhard Domagk, and later with the introduction of
Penicillin as a therapeutic agent in 1942 by Alexander Fleming, the mortality rate of bacterial
infections rapidly decreased and an era had begun. Those three drugs was the beginning of a drug
discovery era. Salvarsan and Prontosil was both later replaced by more non-toxic drugs, while
penicillin antibiotics still are highly used today. With discovery of penicillin, the foundation of drug
discovery the next 20 years had been set. A great number of antibiotics, which are still in use today
in either their original or modified form, were commercially introduced shortly afterwards (Figure
1).

Figure 1: Timeline of antibiotic deployment (top) and antibiotic resistance (bottom) (Clatworthy et al., 2007).

12

Antibiotics are classified according to their mode of action, chemical structure and spectrum of
activity. Per definition these chemotherapeutic agents should be either bactericidal (killing) or
bacteriostatic (inhibiting growth) towards pathogens while having limited side effects when
administered to the host. The target(s) of antibiotics should ideally be essential for the viability of
the infecting pathogen and should not be similar to structures found in host cells. Often, these
essential structures are highly conserved but still contain an evolutionary distance to human
counterparts, consequently some antibiotics act against a wide range of pathogenic bacteria and are
therefore called broad-spectrum antibiotics. The disruption of these essential mechanisms leads to
cell cycle arrest or even better, cell death.
Due to the major impact of penicillins they are often used as a model antibiotic. Penicillins are all of
the -lactams antibiotic class which is named due to the presence of a structural -lactam ring. The
target of penicillins is the cell wall of gram-positive and gram-negative bacteria. This cell wall is of
critical importance for the bacteria and keeps the integrity and shape. The fundamental feature of
the cell wall is the cross-linked polymer, peptidoglycan (Scheffers and Pinho, 2005). Penicillins
disrupt the final synthesizing step of peptidoglycan by irreversibly binding to DD-transpeptidases
(also named Penicillin binding proteins) which lowers the enzyme activity and disrupts the
formation of the cell wall which leads to cell death (Lange et al., 2007). Since the bacterial cell wall
differ from the plasma membrane of human eukaryotic cells, the targeted enzymes are conveniently
not present in eukaryotic cells wand builds the whole foundation of a suitable antibiotic. Other
targets in antibiotic treatment are other targets within the bacterial cell wall, in protein synthesis,
DNA replication, transcription and other essential pathways that make the bacteria viable (BrotzOesterhelt and Brunner, 2008, Morar and Wright, 2010), for examples see Table 1. Whereas
antibiotics that target cell wall and DNA replication is commonly bactericidal, antibiotics which
target the translation are commonly bacteriostatic (except Aminoglycosides) (Table 1).

13

Table 1: Simplified diagram of selected classes of antibiotic, their target, mechanism of action and effect (Brotz-Oesterhelt
and Brunner, 2008, Morar and Wright, 2010).

Antibiotic class

Examples

Target

Mechanism of action

Effect

-lactams

Penicillins
Cephalosporins

Peptidoglycan
biosynthesis

Inhibits the formation of peptidoglycan crosslinking though binding to DD-transpeptidases

Bactericidal

Glycopeptides

Vancomycin
Teicoplanin

Peptidoglycan
biosynthesis

Inhibits the formation of peptidoglycan by


binding to D-Ala-D-Ala.

Bactericidal

Aminoglycosides

Tobramycin
Kanamycin

Translation

Protein synthesis and proofreading are


disturbed by binding to 16S rRNA near A-site
of 30S ribosomal subunit.

Bactericidal

Tetracyclines

Tetracycline
Minocycline

Translation

Binds to 16S RNA at the A-site which inhibits


the binding of aminoacyl-tRNA which blocks
protein synthesis.

Bacteriostatic

Macrolides

Erythromycin
Azithromycin

Translation

Binds to the domain V of 23S rRNA of 50S


subunit which blocks peptide chain elongation.

Bacteriostatic

Quinolones

Ciprofoxacin

DNA
replication

Inhibits and topoisomerases II (DNA gyrase)


and IV.

Bactericidal

Polypeptides

Colistin

Outer
membrane

Destabilizes the outer membrane

Bactericidal

Bacitracin

2.2 Bacterial antibiotic resistance


After the discovery and commercial introduction of antibiotics it was soon observed that usually
treatable infections were not affected by the administration of antibiotics and had adapted
mechanisms of resistance. The cause for this was lack of experience from other antibiotics and
though Alexander Fleming predicted that too low doses would lead to development of penicillin
resistance, and though bacterial resistance towards penicillin was actually discovered before it was
made available on the market, the usage of penicillin was not restricted by any means (Wright,
2012).
After a period in the mid-20th century, it was finally recognized that there should be some antibiotic
control, and penicillin and other antibiotics have to some extent been restricted since.
Antibiotic resistance is a result of the evolutionary pressure that bacteria undergo. Antibiotic
resistance has existed even before the widespread of antibiotics which there are several reasons for
(Wright, 2007).

14

The environment, in which the microorganisms exist, is not a monoculture but consist of a complex
mixture of organisms, where some live in symbiosis, but where many are combatting each other
with many toxic compounds. Many of the surrounding compounds would be lethal to the bacteria if
consumed, and to overcome the toxicity, the organisms must develop strategic countermeasures, a
sort of an environmental selective pressure. Since many antibiotics are uncovered from
microorganisms, the microorganisms found in the same environment could potentially have
developed specific mechanisms to make the antibiotics less harmful, for example by mutating the
antibiotic target or developing antibiotic paralyzing pathways.
Resistance occurs for all antibiotics after they are clinical deployed and there is a limit to the
number of natural antibiotic substances which fulfill all pharmacokinetic demands. Because of this,
much of the antibiotic work done after the 1960s was focused on chemically modifying existing
antibiotics to make them more susceptible to resistant pathogens and to improve pharmacokinetics.

Staphylococcus aureus is a major cause of hospital and community-acquired infections and due to
drug resistance they are becoming an even bigger health care problem. The antibiotic resistance of
this species is well documented, and is in this chapter used to describe how bacteria acquire
resistance.
The main cause of the antibiotic resistance problem with S. aureus lies within the extensive use of
penicillin. When introduced on the market, penicillin was life-changing and extremely efficient in
combating bacterial infections, so it became widespread and as a result of bad administering
strategies, strains of S. aureus was observed to gain resistance towards penicillin. Today some S.
aureus infections are resistant to most antibiotics in the clinic due to their number of adopted
resistance genes.
When penicillin allergies were reported, antibiotics such as macrolides, lincosamides and
streptogramins, which target the bacterial 50S ribosomal subunit, were tried as an alternative
treatment. These treatments led to the fast adaption of resistance since strains already carried two
form of resistance genes to these antibiotic classes. These two resistance mechanisms are
implemented by the methylation of 23S rRNA by ermA, ermB or ermC and by the active efflux
mediated by msrA (Schito, 2006).
Quinolones, which target topoisomerase II and IV, were only in use for a short time before
spontaneous chromosomal mutations in the GrlA subunit of topoisomerase IV and GyrA subunit of

15

topoisomerase II occurred (Figure 1). This resulted in reduced quinolone-protein affinity (Lowy,
2003).
-lactams, such as penicillins, bind very effectively to DD-transpeptidases which play an essential
role in the formation of peptidoglycan cross-linking. Penicillin resistant S. aureus have either
acquired a mutated DD-transpeptidase or a -lactamase which hydrolyses the -lactam ring of the
drug, thus inactivating it (Sauvage et al., 2008).
As for many other antibiotics that faced resistance, natural penicillin was chemical modified to gain
an improved antibiotic, methicillin, which was less susceptible to resistance. But as experienced
with all other notable antibiotics, there was soon to become methicillin-resistant S. aureus
infections, MRSA (Figure 1, p. 12). The resistance gene in MRSA, mecA, is carried by the mobile
SCCmec element and encodes a methicillin-resistant penicillin-binding protein which most likely
has been exogenously acquired from a distantly related Staphylococcus species (Hiramatsu et al.,
2001, Tsubakishita et al., 2010)
MRSAs are increasingly observed in infections among persons without established risk factors and
they are the most common identifiable cause of skin and soft-tissue infections in several
metropolitan areas across the United States (Moran et al., 2006) where they are associated with a
higher mortality than methicillin-susceptible S. aureus, MSSA, infections (Cosgrove et al., 2003).
Unfortunately, MRSA infections are no longer only hospital and healthcare-associated but an
increasing number of infections are seen to be community acquired (Klevens et al., 2007, Naimi et
al., 2003).
When MSSAs develop into a MRSA, they are only susceptible to intravenously administered
glycopeptides antibiotics such as vancomycin. The problem with antibiotic resistance in S. aureus
seems to become even more severe since some MRSAs have developed decreased susceptibility
towards vancomycin, so called vancomycin-resistant S. aureus (VRSA) strains (Hiramatsu et al.,
1997, Sieradzki and Tomasz, 1997).
Today, the search for expansion of new antibiotics is focusing on optimizing preexisting antibiotics,
discovering new drug targets and investigate new potential groups of antibiotics (Donadio et al.,
2010). One qualified target that has been extensively studied is biofilms and will be described in the
section below.

16

2.3 Bacterial biofilms


Bacteria can live in two disparate ways: as single, free-floating cells (planktonic) or in sessile
aggregates, so-called biofilms where the bacteria live in organized communities.
The production of a biofilm originates with the initial adherence of the bacteria to a surface. Here
the bacteria are embedded within a self-produced matrix of extracellular polymeric substance
(EPS), which mainly contain polysaccharides, nucleic acids, lipids and proteins (Costerton et al.,
1978). Biofilms consist mostly of EPS, 90 %, whereas the cells account for only 10 %. The
formation of a biofilm gives the bacteria several advantages: it immobilizes the cells while
maintaining a comfortable architecture allowing the cells to communicate, it creates a reservoir of
nutrients from lysed cells including DNA, which makes horizontal gene transfer more likely to
occur. Of most clinical importance, it also protects the cells from the surroundings such as host
immune defence, many antibiotics, ultraviolet radiation and oxidizing or charged biocides
(Flemming and Wingender, 2010). The biofilm mode of life is a central infection mechanism and is
recognized as the causing or exacerbating feature in many medical infections including dental
caries, nosocomial infections, pneumonia, cystic fibrosis, urinary tract infections, and infections
related to catheters and medical implants (Costerton et al., 1999, Vuong and Otto, 2002). According
to the US National Institutes of Health, biofilms are medically important and they estimate biofilms
to account for 80 % of human bacterial infections.

The typical biofilm development is divided into 5 stages: (I) reversible attachment, (II) irreversible
attachment, (III) maturation-1, (IV) maturation-2, and (V) dispersion (Figure 2) (Stoodley et al.,
2002). Most research has been conducted in P. aeruginosa which is often used as a model biofilm
producing organism and is also done so below.

17

Figure 2: Biofilm development shown as a five-stage process. (I) initial attachment of cells to a surface (II) EPS production which
leads to irreversible attachment (III) early development of biofilm architecture (IV) maturation of biofilm architecture (V) dispersal
of cells (Stoodley et al., 2002).

2.4 P. aeruginosa biofilms


P. aeruginosa is a gram-negative bacterium which lives in the soil and water, but is also an
opportunistic human pathogen causing pneumonia in cystic fibrosis patients, urinary tract infections
by contaminating urinary catheters and skin and soft tissue infections after breach of the skin.
To form a biofilm, P. aeruginosa has the ability to get to and move across a surface while
overcoming the hydrodynamic boundary layer and repulsive forces as the bacteria approach the
surface. A single flagellum drives swimming motility in P. aeruginosa and is thought to enable the
bacteria to reach the surface where it can attach and form a biofilm (O'Toole and Kolter, 1998,
Sauer et al., 2002). Type IV pili contribute to the generation of motile forces, so called twitching
motility, which makes the bacteria move across the surface of which the bacteria attaches itself to
(Mattick, 2002). In contrast too much twitching movement will make it difficult for the bacteria to
settle and form a biofilm (Picioreanu et al., 2007, Singh et al., 2002). These motion effectors are
also involved in the initial attachment phase (I) of P. aeruginosa to biotic and abiotic surfaces as
they also function as adhesins (Giltner et al., 2006, Jin et al., 2011, O'Toole and Kolter, 1998) .
However, the precise nature of that role and its impact on biofilm development varies with
18

environmental conditions (Klausen et al., 2003). Another factor that is related in attachment to
surfaces is the CupA fimbriae (Vallet et al., 2001). Other factors are often linked to the attachment
but there is a lack of knowledge of how these factors plays a role in the attachment.
The irreversible attachment (II) is followed by microcolony formation, but is still not fully
described. Though several c-di-GMP associated proteins, such as SadB, SadC and BifA have shown
to be involved in transition from reversible to irreversible attachment and microcolony formation
(Caiazza and O'Toole, 2004, Merritt et al., 2007, Kuchma et al., 2007, Caiazza et al., 2007). These
proteins are thought to be part of a complex system that regulates swarming motility, which
regulates the reversible attachment and thus continuation of the process of biofilm formation
(Murray et al., 2010), but also regulates EPS production and modulation of flagellar reversal rates
via the chemotaxis cluster IV (Petrova and Sauer, 2012).
After the formation of reversible attachment and microcolonies, a large amount of EPS is produced
(III) and (IV). P. aeruginosa produces three exopolysaccharides, alginate, Psl, Pel, which all are
differently expressed depending on strain (Ryder et al., 2007). For example, mucoid strains isolated
from Cystic fibrosis patients produce alginate in large amounts but in strains such as PAO1 and
PA14 alginate are not a major constituent of EPS (Wozniak et al., 2003).
It is known that Psl is a great constitutor to biofilm formation in PAO1 (Jackson et al., 2004) and is
also required for the attachment to both biotic and abiotic surfaces in many other strains (Ma et al.,
2006, Colvin et al., 2012, Ma et al., 2009). Pel is required for biofilm formation in PA14 (Friedman
and Kolter, 2004) where it also is crucial for the cell-cell contact. This does not account for PAO1,
where Psl seems to be the primary structural polysaccharide for biofilm maturity (Colvin et al.,
2011).
The production of these polysaccharides is highly regulated but the precise role of the different
components of the biofilm matrix remains to be determined. Overall, it is thought that Pel and Psl
are involved in the initial stages of biofilm formation and that alginate serves as the stress response
polysaccharide associated with chronic stages of infection (Schurr, 2013, Ghafoor et al., 2011). The
relationship between Pel and Psl is not determined and there is significant strain-to-strain variability
in the contribution of Pel and Psl to the mature biofilm structure (Colvin et al., 2012).
Dispersal (V) enables the bacteria in a biofilm to spread and colonize new surfaces if it is nutrient
advantageous (Sauer et al., 2004). It involves a phenotypic change of the bacteria so that bacteria in
the dispersion stage are more similar to planktonic bacteria than to maturation-2 stage bacteria
(Sauer et al., 2002). One theory propose that programmed cell death, autolysis and reduced

19

synthesis of Psl matrix are used to create a center in the biofilm matrix with planktonic-like-bacteria
for dispersal (Ma et al., 2009). Clumping dispersal and surface dispersal have also been suggested
as possible dispersal mechanisms (Hall-Stoodley and Stoodley, 2005) but the precise mechanism is
unknown in P. aeruginosa. In Actinobacillus actinomycetemcomitans, Dispersin B degrades polyN-acetylglucosamine, a biofilm matrix polysaccharide, and mutant colonies fail to release cells and
disperse (Kaplan et al., 2003). It is likely that there is a similar enzyme present in P. aeruginosa but
it remains to be discovered.

2.5 Quorum sensing controls biofilm formation


The conversion of a planktonic cell into the biofilm mode of growth is a drastic event and involves
a coordinated phenotypic change of the bacteria in every stage of biofilm formation (Sauer et al.,
2002). The reason for this phenotypic change is still an evolving matter of research but is for
example known to be influenced by nutritional changes (Hancock et al., 2011, Shrout et al., 2006,
Petrova and Sauer, 2012). In the last decade, the knowledge of this feature of bacterial pathogenesis
has been improved. We now consider most bacteria to communicate and react coordinated by
making use of autoinducers, a system known as quorum sensing (Kjelleberg and Molin, 2002). In
many bacteria quorum sensing plays a role in biofilm establishment, growth and maintenance (Irie
and Parsek, 2008). P. aeruginosa are one of the best described bacteria which quorum sensing
system is linked to biofilm formation. Two systems have been described in P. aeruginosa: las and
rhl. Quorum-sensing systems respond to a class of autoinducers named acyl homoserine lactones
(AHLs) (Ng and Bassler, 2009). The AHL autoinducer in las quorum-sensing system is synthesized
by LasI and is regulated by LasR, a transcriptional activator protein (Pearson et al., 1994). The AHL
autoinducer in rhl quorum sensing system is synthesized by RhlI and is regulated by RhlR (Fuqua
et al., 2001). rhl quorum sensing system is furthermore controlled by las quorum sensing system
(Latifi et al., 1996, Pesci et al., 1997, Medina et al., 2003).
Compared to the P. aeruginosa wild type and frequently used laboratory strain, PAO1, lasI mutants
form thin biofilms whereas rhlI is unaffected (Davies et al., 1998). Since then, a number of reports
have shown that rhl quorum sensing system affects biofilm formation and it can might be explained
by different experimental setups (de Kievit, 2009).
In PAO1, both las and rhl are also important for eDNA release to the matrix (Allesen-Holm et al.,
2006). Inactivation of las compromise late stages of biofilm formation but not earlier stages. This
effect, however, might be influenced by environmental conditions (Sauer et al., 2002, Shrout et al.,

20

2006). Both systems are also required for Type IV pilus-dependent twitching motility (Glessner et
al., 1999) and in PA14 they regulate pel expression (Sakuragi and Kolter, 2007).

2.6 Biofilm inhibitors as potential drugs


Some of the advantages of biofilms are that they aid to overcome immune system and create
resistance towards antibiotics. When the bacteria are organized in a biofilm they are up to 1,000
times less susceptible to antimicrobial agents compared to the planktonic state (Smith, 2005, Olson
et al., 2002). For example the hypochlorite resistance increased 600-fold when S. aureus was grown
as a biofilm on an abiotic surface (Luppens et al., 2002). A lower metabolic/slow growing rate
(persisters), difficult penetration of biofilm matrix, up regulation of efflux pumps and stress
response regulons are often mentioned as the reason (Davies, 2003, Lewis, 2010, Costerton et al.,
1999, Stewart, 2002, Stewart and Costerton, 2001). The clinical consequences are that biofilm
infections often develop into a chronic infection which is difficult to eradicate (Hoiby et al., 2011,
Hoiby et al., 2010).
Due to the lack of discovery of new and effective antibiotics, and because biofilms are regarded as a
determent for chronic infections and decrease of antibiotic susceptibility, many alternative strategies
have been focusing on reducing biofilm formation in infections and preventively incorporate
antibiotics into in-dwelling medical devices.
There are numerous ways to inhibit biofilm formation but when it comes to an infection where the
biofilm has already formed, quorum sensing might turn out to be the most strategic advantageous
mechanism to inhibit (de Kievit, 2009). Garlic extracts have for example shown inhibitory effect on
quorum sensing and have a synergistic effect with tobramycin (Bjarnsholt et al., 2005b, Rasmussen
et al., 2005).
Other strategies have focused on decreasing the attachment onto in-dwelling medical devices such
as catheters. This is attempted primarily by coating catheters with antibiotics such as minocycline,
rifampin and Ampicillin (Darouiche et al., 1999, Raad et al., 1997, Liu et al., 2012) and a wide
range of other antimicrobials, for example polymeric materials (Hook et al., 2012), benzalkonium
chloride (Jaramillo et al., 2012), sodium fluoride and chlorhexidine (Liu et al., 2012), and many
more (Smith, 2005).

21

2.7 Urinary Tract infections


Urinary tract infections (UTI) occur through the invasion of microorganisms in the urinary tract,
most commonly in women. In women they are the most prevalent hospital acquired type of
infection accounting for an estimated 25 % - 40 % of all infections and in elderly women, UTIs are
the second most prevalent community acquired infection (Matthews and Lancaster, 2011). It is
estimated that 40 % 60 % of women will get one or more UTI during their lifetime (Salvatore et
al., 2011). Recurrences are common in young women and are associated with sexual intercourse,
use of spermicidal products, having a first UTI at an early age and having a maternal history of
UTIs. In hospitals and long-term care facilities the uropathogen is often introduced to the urinary
tract through contaminated urinary catheters which increases the risk of developing a UTI
significantly.
The infecting pathogens usually arise from ascending infection from the urethra to the bladder. The
most prevalent uropathogen is E. coli (80-90 %) and in healthy women they are suggested to
primarily originate from the gastrointestinal flora. Other uropathogens such as Klebsiella species,
Enterococcus species and P. aeruginosa are less common and are reported to be related to the usage
of urinary catheters (Matthews and Lancaster, 2011).
The establishment and maintenance of the infection are mediated through the formation of a biofilm
by the infecting pathogen (Salo et al., 2009), which in many cases creates hard-to-treat infections
(Tenke et al., 2006). To initiate infection and oppose the wash out effect uropathogenic E. coli
express several virulence factors (Svanborg and Godaly, 1997), but most important is the Dmannose binding by type 1 fimbriae FimH adhesin which mediates attachment to uroepithelial cells
and abiotic surfaces (Finer and Landau, 2004, Klemm and Schembri, 2000). Another virulence
factor in E. coli is the expression of proteinaceous cell surface filaments, curli fimbriae. It is thought
that curli fimbriae are involved in cell-cell and cell-surface interactions and have been linked to
catheter associated infections (Hatt and Rather, 2008).
K. pneumoniae form a different form of biofilm than E. coli, since it is a urease producing species.
Urease gives the ability to the bacteria to form a crystalline biofilm on urinary catheters. In K.
pneumoniae there are two types of fimbriae, type-1 and type-3, both types were equally important
for biofilm formation in a catheterized bladder model (Stahlhut et al., 2012) but only type-3
fimbriae have shown to be important for biofilm formation in flow chamber (Schroll et al., 2010).

22

Since catheter associated urinary tract infections are quite common, there is a need for modified or
coated catheters to lower the occurrence of UTIs. Many strategies have attempted to solve the issue:
integrating antibiotics on urinary catheters is a popular strategy, and it has been attempted many
times, such as with sparfoxacin (Kowalczuk et al., 2012), ciprofloxacin, norfloxacin, and ofloxacin
(Reid et al., 1994), silver (Liedberg and Lundeberg, 1989) and a fish muscle protein,-tropomyosin
(Vejborg and Klemm, 2008). Some have been clinically introduced and have had a positive effect
on reducing catheter associated infections but since infections only are being limited researchers are
still looking for better alternatives (Johnson et al., 2006).

2.8 Antimicrobial peptides


Antimicrobial peptides (AMPs) or host defense peptides are evolutionarily highly conserved
components of the innate immune system and are produced by all complex organisms (Ganz, 2003).
Their importance in host defense is indicated in plants and insects which live in non-bacteria free
environments without the ability to produce lymphocytes and antibodies. In humans and other
mammalians, the significance of peptides in host defense are especially demonstrated by the low
risk of infection in the cornea of the eye where they serve as a first line of defense like they do
throughout the human body, e.g. in epithelia cells of human colon mucosa (Tollin et al., 2003) and
at the skin surface due to sweat glands peptide secretion (Schittek et al., 2001). They are also found
in large amounts in granulocytes where they are part of degranulation (Ganz, 2003).
AMPs are polypeptides of usually 10-50 amino acids of which the majority of them, due to the
positively charged amino acids arginine and lysine, are cationic with an overall charge of +2 to +9.
Hydrophobic residues contribute to 30% of the peptide which gives the peptides an amphipathic
nature with the clustering of cationic and hydrophobic amino acids into distinct domains (Hancock
and Lehrer, 1998, Zasloff, 2002). Many antimicrobial peptides have a wide range of activities and
they often have broad spectrum antimicrobial activity and some even kills multi drug resistant
bacteria at low concentrations. They were originally known for their antifungal, antiviral,
antiparasitic and antibacterial properties (Jenssen et al., 2006).
Most antimicrobial peptides disrupt the bacterial cell wall by forming pores (Figure 3) and therefore
many AMPs show highest activity against gram-positive bacteria. Some AMPs target lipid II or
other cell wall biosynthetic processes to disturb peptidoglycan synthesis and translocation (Yount
and Yeaman, 2013), other targets are DNA, transcription, translation, replication and essential
enzymatic activity Figure 3(Brogden, 2005, Sato and Feix, 2006). There is evidence showing that

23

antimicrobial peptides are not only involved in direct antimicrobial action but also serve as
immunomodulatory peptides, functioning as chemokines and/or inducing chemokine production,
inhibiting LPS induced pro-inflammatory cytokine production, promoting wound healing, and
modulating the responses cells of the adaptive immune response (Bowdish et al., 2005, Oppenheim
and Yang, 2005).

Figure 3: Three pore forming models that explain the mechanism of action of membrane-active antimicrobial peptides. (A)
Carpet model, (B) toroidal pore model & (C) barrel-stave model (Brogden, 2005).

An ongoing discussion is that the activity of antimicrobial peptides at physiological concentrations


is too weak to single-handedly function as mono antibiotics and that the immune stimulating
activities are of greater importance. Though many antimicrobial peptides have been tested in
clinical trials, no antimicrobial drugs have yet been commercially introduced except cyclic
polymyxins. But several non-bacteria combatting immunostimulating peptides and many other
peptide drugs, which are not related to the immune system, are already an established part of the
pharmaceutical industry (Hancock and Sahl, 2006, Sun, 2013).
The advantages of using AMPs as antibiotics are their wide spectrum of activity, target selectivity,
high efficacy at low concentrations, anti-LPS activity, often synergistic action with classical
antibiotics, and low probability for developing resistance. The reason why AMPs often fail clinical
trials is their short half-life, which demands higher concentrations in treatment which leads to side
effects, such as lysis of anionic red blood cells. The problem can be overcome by redistributing the
AMP only at the site of infection. This has led to the idea, that incorporation of AMP in implants

24

could decrease systemic distribution of AMPs and therefore decrease side effects. There are mainly
two ways to achieve this: by leach- or release based systems or covalent attachment.
Many covalent immobilization strategies have been suggested (Costa et al., 2011) but it is important
that every peptide is to be evaluated individually. The covalent immobilization can increase stability
of the peptide while being located at the site of interest for a longer period of time while decreasing
toxic side effects. Many things can influence the activity: peptide orientation, secondary structural
changes and length, flexibility and type of spacer used to link the AMP and substrate (Hilpert et al.,
2009). The covalent linking of AMPs to a substrate also varies a lot, from contact lenses (Willcox et
al., 2008), different types of resin beads (Bagheri et al., 2009) to titanium surfaces (Gabriel et al.,
2006). Leakage system such as a calcium phosphate-coated assay on titanium surfaces have also
been conducted with positive results (Kazemzadeh-Narbat et al., 2010).

The increasing problem with antibiotic resistance demands that we engage these biofilm infections
from a new perspective. Antibiofilm drugs have a great potential but will possibly show not to be
broad spectrum drugs since the molecular composition of the biofilm seems not to be highly
conserved. Since an antibiofilm effect is below the killing concentration, resistance to the peptide as
an antibiotic could be a minor risk. In contrast, highly conserved AHLs could be a broad-spectrum
target (Bassler, 2002).
A new approach to study these AMPs is to study their inhibition of biofilms which have been
shown to be accomplished by a numerous number of antimicrobial peptides (Batoni et al., 2011,
Jorge et al., 2012) and in many biofilm producing organisms which are of high clinical relevance,
such as oral streptococci (da Silva et al., 2013), S. aureus (Hochbaum et al., 2011), E. coli (Hou et
al., 2010) and P. aeruginosa (Kapoor et al., 2011). The most studied human AMP, LL-37, is from a
family of peptides named cathelicidins. It has shown antibiofilm effect against most of the biofilm
producing species mentioned above.

25

2.9 Human cathelicidin LL-37


In mammals, there are mainly two major groups of antimicrobial peptides, defensins and
cathelicidins. Cathelicidins are a heterogeneous group and vary in length from 12-80 amino acids
but the majority is 23-37 amino acid residues long which, in contact with a biological membrane,
folds into an amphipathic -helix. Cathelicidin are classified based on the presence of a highly
conserved cathelin domain (Pro-region) consisting of approximately 100 amino acids, flanked by
an N-terminal signal peptide (Pre-region) and the antimicrobial peptide on the C-terminal end
which become active upon cleavage of the prepro-peptide (Zanetti et al., 1995, Gennaro and
Zanetti, 2000).
Cathelicidins were initially discovered in myeloid stem cells of bone marrow (Bagella et al., 1995,
Zanetti et al., 1995) and they are most important of all stored in granules of neutrophils (Cowland et
al., 1995). Later, observations of constitutive and inducible expression of cathelicidins has been
observed in broad range of mammalian organs but one of the biggest focuses of the research has
been on the expression of human cathelicidin in epithelial cells (Zanetti, 2005, Kosciuczuk et al.,
2012). Cathelicidin mode of actions includes a broad range of properties such as direct
antimicrobial activity, LPS binding, chemoattraction of immune cells, stimulating release of
histamine from mast cells, immune stimulating and induction of angiogenesis (Bals and Wilson,
2003, Kosciuczuk et al., 2012).

Some mammals generate various cathelicidins (Zanetti, 2004), but in humans only one member has
been identified, hCAP-18. The CAMP gene expresses a prepropeptide consisting of a 30 amino acid
signal peptide and hCAP-18 which consists of a 103 amino acid cathelin domain and a 37 amino
acid C-terminal peptide, LL-37 (Figure 4) (Larrick et al., 1996, Gudmundsson et al., 1996).

Figure 4: The structure of CAMP product. The product consists of a 30 amino acids signal peptide at the N-terminal (red), a 103
amino acid pro-region (blue) and a 37 amino acid antimicrobial at the C-terminal (green).

26

LL-37 (LLGDFFRKSKEKIGKEFKRIVQRIKDFLRNLVPRTES) folds up into an amphiphilic ahelical structure (Wang, 2008) and exists in an equilibrium between monomers and oligomers both
in solution and in contact with zwitterionic lipids (Oren et al., 1999). Due to its cationicity (+6) and
hydrophobicity, it can interact with negatively charged membranes by a suggested carpet-like
mode of action (Figure 3), explaining the peptide antibacterial properties (Larrick et al., 1995b): it
has later been demonstrated that the peptide also has the potential to interact directly with bacterial
LPS (Larrick et al., 1995a, Nagaoka et al., 2001).
hCAP-18 is expressed in many types of blood cells (e.g., T cells, B cells, monocytes,
macrophages (Agerberth et al., 2000) and mast cells (Di Nardo et al., 2003)), but is mainly found in
large amounts in granules of polymorphonuclear neutrophils (Cowland et al., 1995) where it is kept
inactive until cleavage of the pro-protein by proteinase 3, which yields the active peptide LL-37
(Sorensen et al., 2001). Similarly, hCAP-18 is expressed in a variety of epithelial cells in the colon
(Hase et al., 2002) and cells of the gastric mucosa (Hase et al., 2003), urinary tract (Chromek et al.,
2006), epididymis (Malm et al., 2000), corneal (Gordon et al., 2005) and lungs (Bals et al., 1998), in
many cases resulting in secretion of LL-37 to fluids lining these epithelial layers. Additionally, LL37 has also been found in fluids and tissue from bone marrow, testes (Agerberth et al., 1995),
seminal plasma (Andersson et al., 2002), wound and blister fluid (Frohm et al., 1996), sweat
(Murakami et al., 2002b), skin (Yamasaki et al., 2006) and salivary glands (Murakami et al., 2002a,
Davidopoulou et al., 2012), among others.
Cleavage of hCAP-18 by proteinase 3 yielding LL-37 is embraced as the most prevalent fate of
hCAP-18. However, at vaginal pH hCAP-18 is cleaved by gastricsin to ALL-38 (Sorensen et al.,
2003). This is particularly interesting since hCAP-18 expression is found in seminal plasma but not
in vaginal fluid. Hence, the seminal plasma-derived hCAP-18 could become cleaved and activated
following sexual intercourse. Furthermore LL-37 have also been found to be cleaved into smaller
fragments, such as KR-20, KS-30, RK-31, LL-23, KS-27, LL-29 and KS-22, in sweat and skin cells
by kallikrein serine proteases, of which many of these fragments have shown increased
antimicrobial activity (Yamasaki et al., 2006, Murakami et al., 2004) and decreased
immunostimulatory functions (Braff et al., 2005).
The expression of LL-37 is thought to mainly be regulated by vitamin D (Wang et al., 2004, Dixon
et al., 2012) but the regulation of LL-37 during immune system activation seems to be controlled by
a more complex mechanism. LL-37 has been linked to several biological processes and

27

demonstrates importance in fighting infections during inflammation, cell differentiation and postinjury. Consequently, both elevated expression levels of hCAP-18 and increased release of the
active peptide LL-37 have been observed in response to microbial infections and various chronic
inflammation diseases. Decrease of expression has also been observed but overall, the expression
are most often induced during inflammation (Vandamme et al., 2012, Durr et al., 2006).

The antibacterial activities of LL-37 are often reported to be weak (Durr et al., 2006, Turner et al.,
1998) which are often explained by the in vivo presence of other host defence peptides which all act
in synergy. The antibacterial activity of LL-37 (and other antimicrobial peptides) is also weakened
with increasing NaCl concentrations matching physiological conditions (Bals et al., 1998, Turner et
al., 1998).
Other antimicrobial actions of LL-37 are anti-fungal and anti-virus activity (den Hertog et al., 2005,
Lopez-Garcia et al., 2005, Barlow et al., 2011, Yasin et al., 2000, Gordon et al., 2005, Wong et al.,
2011). The anti-fungal activity has not been a major focus of research whereas LL-37s antiviral
effect is often debated and similarity with the antibacterial debate LL-37 might exhibit strongest
anti-viral activity in combination with other host defense peptides.

Aside from the antimicrobial effects described above, recent data have suggested conflicting roles
for LL-37 in tumor development: LL-37 is shown to suppress colon cancer (Ren et al., 2012),
support natural killer cells in their antitumor effect (Buchau et al., 2010) but on the flip side, LL-37
have been suggested to assist breast cancer development (Heilborn et al., 2005).
LL-37 have also been connected to wound healing and is thought to assist by creating a barrier from
infection, modulating wound healing, assist in apoptosis and stimulating wound closure through reepithelialization (Heilborn et al., 2003, Huang et al., 2006, Koczulla et al., 2003).

However, the inducible expression of hCAP-18 during inflammation demonstrates an immense


immunological role for LL-37 and in the latest years this has been one of the hottest LL-37 topics.
LL-37 modulates the immune system in several ways of which many are still unclear, though most
evidence points towards LL-37 as an alarmin. LL-37 are proposed to bind to formyl peptide
receptor like 1, causing Ca2+ mobilization and chemoattraction of immune cells such as monocytes,
neutrophils and T lymphocytes (De et al., 2000). LL-37 is also chemotactic for mast cells via a
suggested Gi proteinphospholipase C signaling pathway (Niyonsaba et al., 2002), which mobilizes

28

the intracellular Ca2+ and induces histamine release (Niyonsaba et al., 2001). In the presence of
serum, LL-37 antibacterial and cytotoxic activity is decreased due to a conformational change
(Johansson et al., 1998). Furthermore LL-37 exhibits many other immune stimulatory effects such
as cytokine release and modulation of adaptive immunity (Vandamme et al., 2012, Wuerth and
Hancock, 2011). For an overview of LL-37s effects see Figure 5.

Figure 5: Multifunctional effects of LL-37. LL-37 has antimicrobial properties and can target the microbes directly at the surface
(1), inside the host (2) or by binding to LPS (3). Additionally, LL-37 also functions as an immunostimulating peptide, which counter
acts and regulates the release of pro-inflammatory cytokines from macrophages (4). LL-37 also stimulates histamine release from
mast cells (5), and recruitment of both monocytes and polymorphonuclear leukocytes from the blood (6). The polymorphonuclear
leukocytes will also promote phagocytosis of invading bacteria (7). Increased concentrations of LL-37 will also stimulate dendritic
cells (8) and indirectly drive antigen processing and expression of co-stimulatory molecules that ultimately stimulates the T-cell
population (9). Additionally, it has been demonstrated that LL-37 has an angiogenic effect (10) and also promote wound repair
through stimulation of re-epithelialization, fibroblast growth and adherence (11) (Jacobsen and Jenssen, 2012).

29

2.10

Antibiofilm activity of LL-37

Although antibacterial activity of LL-37 has been shown in several studies it is still questionable
how secreted LL-37 function in vivo. The activity of antimicrobial peptides in environments with
high concentrations of NaCl has been center of debate among scientists where many believe that
NaCl has a negative effect on antibacterial effect of peptides (Bals et al., 1998). Therefore it is
questionable if LL-37 has a significant antibacterial effect in salt containing environments where
LL-37 is secreted into. It is argued that the secretion of LL-37 into the sweat and urine could be a
way to get rid of the peptide or that LL-37 is further degraded into more active peptides and finally
it is also be argued that it functions as a persistent immune stimulator when secreted. Another
answer implicates LL-37s effect on reducing attachment and biofilm formation.
The first research that showed LL-37 to exhibit an antibiofilm was done on P.aeruginosa biofilms
and is relatively new (Overhage et al., 2008) but yet it is well documented, that LL-37 exhibit an
antibiofilm effect towards a wide range of bacterial species (Table 2) (Jacobsen and Jenssen, 2012).
The most distinctive characteristic is the antibiofilm activity LL-37 shows at sub-MIC
concentrations. For example is LL-37 active against Francisella novicida biofilms as low as at 3.8
ng/ml whereas the MIC is 250 g/ml (Amer et al., 2010) and in P.aeruginosa the inhibition was as
low as 0.5 g/ml which was 128 times lower than their detected MIC value (Overhage et al., 2008).
This sub-MIC antibiofilm activity goes again for a wide range of common pathogenic bacteria
(Table 2). LL-37 inhibits the initial attachment of the bacteria to abiotic surfaces and inhibits
preformed P.aeruginosa biofilms (Hell et al., 2010, Dean et al., 2011b, Dean et al., 2011a,
Overhage et al., 2008).
The antibiofilm mechanisms of LL-37 are not thoroughly studied but in P.aeruginosa Las and Rhl
quorum sensing system are affected by LL-37 which consequently decreases biofilm relevant gene
expression. LL-37 also increases the twitching motility in P.aeruginosa which makes the bacteria
unable to settle on a surface, but in contrast, LL-37 is not influencing the swimming and swarming
in P.aeruginosa (Overhage et al., 2008, Dean et al., 2011b).
In E. coli, the decreased attachment can be explained by LL-37's ability to inhibit the curli structure
development by preventing the polymerization of the major curli subunit (CsgA) even in low
concentrations, and thereby minimize attachment to epithelial cells (Kai-Larsen et al., 2010).
LL-37 is also able to bind to cell wall polysaccharides of the fungus Candida albicans which
prevents the adherence of the fungus in an in vivo mouse urinary bladders model (Tsai et al., 2011).

30

In addition, LL-37 also binds to EPS polysaccharide alginate of P.aeruginosa and the capsular
polysaccharide K40 of K. pneumoniae, which consequently increases the MIC of LL-37 and the
survivability of the bacteria (Herasimenka et al., 2005, Foschiatti et al., 2009, Benincasa et al.,
2009)
Table 2: In vitro effect of LL-37 on biofilm formation. The studies represent the percentage inhibition of LL-37 after 24h
incubation. Inhibition indicates the percent inhibition of bacterial biofilm formation, compared with untreated controls.

Bacteria strain Inhibition (%)

LL-37 concentration

Reference
g/ml

S. epidermidis

43

0.22

(Hell et al., 2010)

F. novicida

>80

0.2

0.05

(Amer et al., 2010)

S. aureus

>40

10

2.23

(Dean et al., 2011a)

E. coli

80

11.2

2.5

(Kai-Larsen et al., 2010)

P. aeruginosa

40

0.5

0.11

(Overhage et al., 2008)

P. aeruginosa

50

0.22

(Dean et al., 2011b)

P. aeruginosa

35

4.5

(de la Fuente-Nunez et al., 2012a)

P. aeruginosa

57

13.5

(Mohanty et al., 2012)

The antibiofilm effects of LL-37 are indicated not to be a unique feature of the cathelicidin peptide
family. Whereas Indolicidin, a bovine cathelicidin, showed to inhibit P.aeruginosa biofilm
formation, CRAMP, a mice cathelicidin, did not affect P.aeruginosa biofilm formation (Overhage
et al., 2008, Dean et al., 2011b).

2.11

The potential of LL-37 and 1037 as UTI inhibitors

The study by Kai-Larsen et al. 2010 shows that LL-37 prevents the assembly of curli fimbriae from
a UTI isolated E. coli, which indicates that LL-37 antibiofilm properties plays an active preventive
role in establishment of infection by minimizing epithelial cell attachment. Another peptide, 1037,
has shown great antibiofilm potential, inhibiting P.aeruginosa biofilms with 78 % at x MIC (MIC
304 g/ml) (de la Fuente-Nunez et al., 2012a). The major aim for these experiments is to show that
the biofilm formation of three UTI isolates, one K. pneumoniae and two E. coli isolates can be
inhibited by LL-37 and peptide 1037. The experiments are initiated by confirming and determining

31

the MIC and biofilm inhibition on P.aeruginosa by a chosen spectrum of LL-37 derivatives and
1037 analogs. The experimental setup is outlined in Figure 6.

Figure 6: Experimental outline of the conducted experiments.

Based on statements and studies from a previous experiment (Pompilio et al., 2011) tetracycline
was used as a negative control and tobramycin was used as a positive control for P.aeruginosa
biofilm inhibition.
All biofilm experiments are performed as abiotic static surface assay in 96-well microtiter plates.
Static biofilm systems make it possible to study early stages of biofilm formation, but more
importantly, the peptide amount needed for these assays are smaller than in continuous flow cell
biofilm assays, and the continuous effect to the peptide in the bacterial environment can be
determined.

2.12

Structure activity relationship analysis

To analyze the contribution of the specific residues to the structure and function of 1037, an
alanine-scanning technique was used. Alanine-scanning is a commonly used technique to determine
epitopes and identify the contribution of specific amino acids in proteins and peptides. The method
allows a wild type amino acid to be substituted with alanine, either by mutagenesis or peptide
synthesis, to deduce the roles of the wild type side chain (Weiss et al., 2000). The alanine side chain
consists of a non-reactive methyl group and is therefore non-bulky. Alanine is chosen due to its
32

simplicity, and because the simple side chain of glycine would lead to a conformational change,
thus in most cases alanine substitutions will mimic the secondary structure, but in cases where the
secondary structure is affected and need to be conserved, amino acids with larger side chains, such
as leucine and valine, can be applied (Morrison and Weiss, 2001).
When synthesized, it is easy to detect the critical positions within the peptide sequence in relation to
antimicrobial activity by measuring and comparing MIC values. An increase in MIC would mean
that the amino acid in question is of major importance for the antimicrobial function of the peptide.
This could both be due to a lack in binding contact with peptide and bacteria or the secondary
structure could be disrupted. Most amphipathic cationic antimicrobial peptides are linear and rely
on the conservation of their linear -helical structure to disrupt of the bacterial membrane (Powers
and Hancock, 2003) and this makes a substitution of hydrophobic side chains difficult to conduct
without losing its secondary structure. Alanine scanning of fallaxin have given rise to peptides with
higher antimicrobial activity but common for all of them was a higher hemolytic activity (Nielsen et
al., 2007) thus, the technique is primarily thought to result in decrease of antimicrobial activities.

33

3. Materials and methods


3.1 Bacterial strains
A panel of strains was chosen due to their popularity as laboratory strains (Table 3). P. aeruginosa
PAO1 are frequently used for peptide experiments and was mainly chosen to replicate previous
results. To my knowledge no one has conducted experiments with LL-37 and K. pneumoniae
C3091, E. coli 536 and E. coli CFT073 and these strains were all chosen due to their clinical
background.

Table 3: Bacterial strains used for the experiments

Name

Strain

Clinical data

Reference

P. aeruginosa

PAO1

Burn-wound infection (Stover et al., 2000)

K. pneumoniae C3091

UTI

(Struve et al., 2009)

E. coli

536

UTI

(Hochhut et al., 2006)

E. coli

CFT073 UTI

(Welch et al., 2002)

3.2 Bacterial growth conditions


All strains were grown overnight in lysogeny broth (LB) at 37 C in a shaking waterbath, diluted in
25 % glycerol and kept at -80 C until use. Before individual experiments, the bacteria were plated
on a LB agar plate which was kept at 5 C. The day before the experiments, single colonies was
picked from the plate and grown overnight in 5 ml of selected media for each experiment in a
shaking water bath at 37 C.
For biofilm experiments the following media was used: LB, BM2 (62 mM potassium phosphate
buffer (pH 7), 7 mM (NH4)2SO4, 2mM MgSO4, 10 M FeSO4, 0.4% (wt/vol) glucose and 0.5%
(wt/vol) Casamino Acids), M9 minimal (42mM Na2HPO4, 22 mM KH2PO4, 8.5 mM NaCl, NH4Cl
19 mM, 2 mM MgSO4, 0.1 mM CaCl2 and 0.4 % glucose) and DMEM (Cat.-No.: 31885-023
Gibco).

3.3 Peptides and antibiotics


Peptide LL-37 (LLGDFFRKSKEKIGKEFKRIVQRIKDFLRNLVPRTES) was made using Fmoc
SPPS and purchased from GenScript as a crude product. This peptide was purified with HPLC and
purity and identification was determined with HPLC and MS. The amount of HPLC purified

34

peptide was too low and therefore peptide was given as a gift from R.E.W Hancock Laboratory,
University of British Columbia.
Seven 12 residue LL-37 fragments and a 12 residue fragment with a valine substitution with Cterminal amidation were purchased from GenScript.
Peptide 1037 (KRFRIRVRV) and 9 peptides with a alanine substitution in the individual residues
(1037 A1 A9) and 7 other synthetic analogues of 1037 were synthesized with a C-terminal
amidation by Solid-phase synthesis using Fmoc chemistry by Hvard Jenssen, Roskilde University.
The peptides were stored in autoclaved MilliQ water and stored at -20 C. The peptides were
prepared as two fold dilutions containing 0.01% acetic acid and 0.2% BSA in 10 times higher
concentration than desired and stored in polypropylene 96 well microtiter plates and wrapped in
parafilm.
Tetracycline was dissolved in ethanol and stored at 5 C, and tobramycin was dissolved in
autoclaved MilliQ and stored at 5 C, both having a stock concentration of 1 mg/ml. They were
prepared in 10x test concentration with 0.01% acetic acid and 0.2% BSA in polystyrene 96 well
microtiter plates, wrapped in parafilm and stored at 5 C for maximum a week.

3.4 Purification of LL-37 with HPLC


Purchased LL-37 was purified with high performance liquid chromatography (HPLC). 20 mg LL37 crude extract dissolved in 20 % acetonitrile (ACN) was injected into the sample loop. For the
reversed-phase chromatography a 201SP C18 250x10mm 10um column was used with buffer A,
H2O and 0.1 % TFA and buffer B, acetonitrile. The flow rate was 5 ml/min and all products were
collected in fractions of 5 ml which were stored at 5 C before freeze drying.

3.5 Identification of LL-37 with HPLC/MS


Purity of the LL-37 HPLC purified product was analyzed with HPLC/MS with a Luna C18(2)
column. A small portion of each fraction of interest was dissolved in 23 % acetonitrile. The two
solvents were 1 % HCOOH, 94 % H2O and 5 % acetonitrile and methanol. The mass was
determined by ESI-MS with a Finnigan LTQ.
Due to a systemic breakdown, 1037 alanine scanning peptides was not purified by HPLC and the
purity was not determined with HPLC/MS. The mass was determined directly by ESI-MS spray of
a peptide concentration of 0.1 mg/ml.

35

The mass weight was identified by using this formula:

Where p is read as mass per charge at the x-axis (m/z), Mw is the mass weight of the peptide and z
is the charge at x protons.

3.6 MIC determination


The minimum inhibitory concentration, MIC, was determined by broth microdilution (Wiegand et
al., 2008). All strains were grown overnight in MH broth (Cat.-No.: 275730 BD Difco) in a
shaking water bath at 37 C. The culture was diluted 11 times in MH broth and grown exponentially
in a shaking water bath at 37 C until the optical density (600 nm) 0.400, which are corresponding
to 1 x 108 CFU/ml. The bacteria were then diluted 500 times yielding the desired bacterial test
concentration 5 x 105 CFU/ml. For verifying this cell number the bacteria were diluted 500 times in
0.9 % NaCl and 100 l suspension was plated in duplicates.
The plates were prepared with 10 l peptide or antibiotic solution in two fold dilutions containing
0.01% acetic acid and 0.2% BSA in 10 times higher concentration than desired test concentration.
The peptides were prepared in range of 2.56 mg/ml to 0.5 mg/ml and the antibiotics were prepared
in the range of 50 mg/ml to 0.05 mg/ml. 90 l bacteria (5 x 105 CFU/ml) were added to each well
with a multichannel pipette. Positive controls were 10 l autoclaved MilliQ water with 0.01% acetic
acid and 0.2% BSA and 90 l bacteria (5 x 105 CFU/ml). Negative controls were 10 l autoclaved
MilliQ water with 0.01% acetic acid and 0.2% BSA and 90 l medium used for dilutions. All MIC

experiments were conducted in polypropylene 96-well microtiter plates (Cat.-No.: 650201 Greiner
Bio-One) and replicated three times.
The MIC was defined as the minimal concentration needed to inhibit all visible growth after 48
hours.

3.7 Biofilm experiments


Biofilm formation was analyzed in an abiotic static surface assay in polystyrene 96-well microtiter
plates (Cat.-No.: 655101 Greiner Bio-One) and for P. aeruginosa in polypropylene 96-well
microtiter plates (Cat.-No.: 650201 Greiner Bio-One) (O'Toole et al., 1999, Merritt et al., 2005).
Single colonies was inoculated in 5 ml appropriate medium and grown overnight at 37 C in a
shaking water bath. 96-well microtiter plates were prepared with 10 l peptide/antibiotic solution in
2 fold dilutions in concentrations 10 times higher than desired final concentration. P. aeruginosa
36

biofilm experiments were conducted in LB medium, K. pneumoniae experiments were conducted in


M9 and BM2 media, and E. coli biofilm experiments were conducted in DMEM and LB media.
The overnight cultures were diluted 100 times and 90 l of the suspension was distributed in the
wells, with positive controls in three wells and negative control with medium only, which data was
deducted from all data. The suspension was diluted 105 and 100 l was plated on LB agar plates to
disconfirm contamination. The microtiter plates were closed tightly with parafilm and incubated at
37 C for 24 hours.
After 24 hours the planktonic bacteria were gently removed, and the remaining were stained with
125 l, 1 % crystal violet (0.1 % for P. aeruginosa) and incubated for 10 minutes at room
temperature. Excess crystal violet was removed and the wells were washed with 200 l phosphate
buffered saline (PBS) 2 times. For dissolvent of attached crystal violet, 200 l 95 % ethanol was
added to each well and incubated for 10 minutes at room temperature. The content was briefly
mixed by pipetting, and 125 l of the solubilized crystal violet was transferred to a separate well in
an optically clear flat-bottom 96-well polystyrene plate. The OD595 was measured by a Bio-Tek
Synergy HT Microplate Reader in all experiments except for tetracycline and tobramycin
experiments where OD600 due to an induction of biofilm formation.
The initial attachment was tested for K. pneumoniae and P. aeruginosa as described above but with
an incubation time of 2 hours.
The effects of LL-37 and 1037 on preformed K. pneumoniae biofilms were analyzed by letting the
biofilm form in 24 hours on polystyrene 96-well microtiter plates. After removal of the planktonic
bacteria, fresh medium with the desired peptide concentrations were added and incubated 3 hours at
37 C followed by staining as described above. The effect of LL-37 and 1037 on preformed P.
aeruginosa biofilms was analyzed by letting the biofilm form in 24 hours on polypropylene 96-well
microtiter plates. After removal of the planktonic bacteria, fresh medium with the desired peptide
concentrations were added and incubated 24 or 2 hours at 37 C followed by staining as described
above. All biofilm percentage means with standard deviations can be found in Appendix I.

3.8 Time-kill kinetics


Time-kill kinetics was performed in polystyrene 96-well microtiter plates with BM2 medium. The
same setup as for the biofilm experiments was used. The suspension was removed from each well at
each time points, diluted and plated on LB plates. The experiment was performed as in singles with
no replication.

37

3.9 Statistical analysis


Two tailed Student's t-test was used for the determination of statistical significance with a threshold
for statistical significance at p = 0.01. Statistical analyses were accomplished with GraphPad Prism
5.

38

4. Results
4.1 HPLC purification of LL-37
LL-37 was purchased as an unpurified crude extract and was therefore purified with HPLC, which
with the chosen HPLC column, using reverse phase, segregates the injected content according to its
hydrophobic/hydrophilic nature.
LL-37, dissolved and injected into the system, are in a mobile phase but when the reversed-phase
chromatography column is reached, the solution is in a stationary phase. In the column, the content
is allowed to run through the column based on its hydrophobic nature and based of a gradient of
acetonitrile.
The purification of LL-37 with HLPC is plotted on Figure 7. A peptide bond is detected by UV
absorption at a wavelength of 214 nm which is represented by the red line and aromatic amino acids
are detected by UV absorption at a wavelength of 280 nm which is represented by the blue line. The
area under the peaks often can be correlated with purity of the solution. In this case the solution
would seem to be very impure but taken in mind that LL-37 often forms polymers (Oren et al.,
1999), it might not be the case.

Figure 7: RPC-HPLC of LL-37 with a flow rate of 5 ml/ml in a gradient of ACN. Gradient I: 20 % to 70 % ACN over 20 column
volumes (5 ml fractions). Step II: Instant jump to 100 % ACN over 2 column volumes (12 ml fractions). Gradient III: 100 % ACN
over 2 column volumes (12 ml fractions). Step IV: 100 % - 20 % ACN. Gradient V: 20 % ACN over 5 column volumes (12 ml
fractions). Lines are as followed: UV at 280nm (Blue), UV at 214 nm (Red) and gradient of acetonitrile (Green). Red numbers at xaxis represent the fraction collection tube.

39

4.2 MS identification of peptides


HPLC/MS was carried out on HPLC fractions 33, 40, 41, 62 and 69. By using the formula
described in chapter of methods and the molecular weight of LL-37 (4493.33 g/mol), the theoretical
mass per charge (m/z) was calculated (Table 4).
Table 4: Theoretical mass per charge (m/z) of LL-37

m/z 1124.33 899.66 749.88 624.90 562.66

Figure 8 represent the trend of most of the analyzed fractions. All analyzed fractions was found to
be very pure with only a single peak, as in Figure 8A. All analyzed fractions except peak 41, which
was slight larger in size than LL-37, was found to be a more or less pure and contained LL-37. The
chosen example, Figure 8B, differ from the theoretical m/z of LL-37 at z = 4 with 0.06 which is
considered satisfactory. Fraction 41 had a difference in m/z of 3.45 and was therefore not
considered being LL-37.
The LL-37 derived peptides (Table 5, p. 42) were purchased purified and was therefore not MS
analyzed. Peptide 1037 and the 1037 single residue substitution peptides (Table 6, p. 44) were
confirmed by direct ESI-MS and, due to breakdown of machinery, the purity was not examined.

Figure 8: HPLC/MS of HPLC fraction 40. Figure A shows HPLC identified purity of the fraction. Figure B shows the mass
spectrum of LL-37 identified with ESI-MS.

40

4.3 MIC determination of antibiotics, LL-37 and LL-37 12-residue fragments


P. aeruginosa strain PAO1 is a popular laboratory strain and is frequently used in MIC
determination and biofilm experiments. To verify the functionality of the broth microdilution assay
and to compare the antimicrobial activity of peptides and antibiotics with other experiment, PAO1
was chosen as a model organism. Antibiotics, such as tobramycin and tetracycline, have been
reported to reduce or induce the biofilm formation, respectively, and they were therefore included
as controls (Linares et al., 2006, Pompilio et al., 2011).
The MIC of LL-37 was 32 g/ml for P. aeruginosa and K. pneumoniae and 16 g/ml for both E.
coli strains (Table 5). This corresponds to the MIC for P. aeruginosa which has been described
elsewhere (de la Fuente-Nunez et al., 2012a) and that LL-37 usually is more sensitive towards E.
coli than P. aeruginosa (De Smet and Contreras, 2005). The LL-37 12-mer fragments were in this
project only screened for potential antimicrobial activity against P. aeruginosa and K. pneumoniae.
It is notable that the MIC of KR-12 is decreased 8-fold compared to FK-12 which is overlapping
with the addition of N-terminal phenylalanine and deletion of the cationic C-terminal arginine. The
cationicity of antimicrobial peptides is thought to often be correlated with antimicrobial activity.
VQ-12 supports that hypothesis since the substitution of the anionic aspartic acid with a
hydrophobic valine, VQ-12V6, increases the antimicrobial activity. The relationship between
antimicrobial activity and charge of the antimicrobial peptide is illustrated on Figure 9A, and the
relationship between antimicrobial activity and peptide hydrophobicity is illustrated on Figure 9B.
The trend line of Figure 9A supports the theory (r2 = 0.5741) whereas the hydrophobicity shows no
correlation with antimicrobial activity for LL-37(r2 = 0.0161).

Figure 9: Correlation between the MIC of P. aeruginosa of LL-37 related peptides and A, Charge (r2 = 0.5741), and B,
Hydrophobicity (r2 = 0.0161). The MIC values >200 was edited to 200.

41

Table 5: MIC determination of LL-37 native fragments, a synthetic analogue of LL-37, tobramycin and tetracycline. a Generally accepted nomenclature. b Number of
the first and last residue in native LL-37 of the derivate. c Single letter amino acid code of the peptide. d MIC determination of the tested antimicrobials with least 3 replicates. e
Charge of the peptide at pH 7. f Ratio of hydrophobic/total number of residues (%) calculated with innovagen peptide calculator. g Mean relative hydrophobic moment (mH)
calculated with HydroMCalc using Eisenberg scale. Strains were P. aeruginosa PAO1, K. pneumoniae C3091, E. coli 536 and CFT073. Underscore is used to make
comparison easier. Not determined MIC is marked ND.

Name

Residue

Sequence

MIC (g/ml)d

PAO1

C3091

536

CFT073

Chargee

%f

mHg

LL-37

1-37

LLGDFFRKSKEKIGKEFKRIVQRIKDFLRNLVPRTES

32

32

16

16

35

0.44

LL-12

1-12

LLGDFFRKSKEKIGKEFKRIVQRIKDFLRNLVPRTES

>200

>200

ND

ND

33

0.38

SK-12

9-20

LLGDFFRKSKEKIGKEFKRIVQRIKDFLRNLVPRTES

>200

>200

ND

ND

25

0.49

FK-12

17-28

LLGDFFRKSKEKIGKEFKRIVQRIKDFLRNLVPRTES

100

100

ND

ND

50

0.68

KR-12

18-29

LLGDFFRKSKEKIGKEFKRIVQRIKDFLRNLVPRTES

12.5

>200

ND

ND

41

0.75

VQ-12

21-32

LLGDFFRKSKEKIGKEFKRIVQRIKDFLRNLVPRTES

>200

>200

ND

ND

58

0.6

VQ-12V6 21-32V26

LLGDFFRKSKEKIGKEFKRIVQRIKVFLRNLVPRTES

50

>200

ND

ND

58

0.47

IK-12

24-35

LLGDFFRKSKEKIGKEFKRIVQRIKDFLRNLVPRTES

>200

>200

ND

ND

50

0.43

DF-12

26-37

LLGDFFRKSKEKIGKEFKRIVQRIKDFLRNLVPRTES

>200

>200

ND

ND

41

0.4

Tobramycin

0.1

0.2

0.8

0.8

Tetracycline

25

6.2

1.6

1.6

42

4.4 MIC determination of 1037 and synthetic analogues


The MIC of 1037 against P. aeruginosa was 16 g/ml which in comparison to the original
publication of 1037 (MIC 304 g/ml), is extremely low (de la Fuente-Nunez et al., 2012a). Since a
high genomic diversity of laboratory P. aeruginosa strains is known to occur (Klockgether et al.,
2010), the MIC was replicated on the P. aeruginosa strain from the original experiment and the
MIC was confirmed to 16 g/ml. E. coli strains was more susceptible towards 1037 (MIC 8 g/ml)
which is a trend that goes again in these experiments.
To determine the amino acids that are most important for the antimicrobial activity of 1037, several
1037 synthetic analogues was synthesized including 9 alanine single-substitution peptides (Table 6
& Figure 10).

Figure 10: Overview of the MIC of 1037 alanine scanning library (A-D). Letters (x-axis) represent the amino acids of 1037 which
have been substituted with alanine at each position, starting with K1. The specific antimicrobial activity (g/ml) of the individual
peptides are plotted on the y-axis. Arrows indicate MIC >256. The MIC was determined for P. aeruginosa (A), E. coli 536 (B), K.
pneumoniae (C) and E. coli CFT073 (D).

Both K1 and R2 showed only a slight involvement in determining the MIC. This observation led to
the synthesizing of 1037 with 1 and 2 deleted N-terminal residues.
Of the 7 C-terminal residues, four were hydrophobic and three of four of these, F3, I5 and V7, were
shown to be very important or essential for antimicrobial activity. Of the three arginine residues R6
sustained most antimicrobial activity whereas the MIC of R4 and R8 increased 4-8 folds.
For further study and to support some of the hypothesis created by the alanine scan, a new list of
peptides was synthesized. Deletion of the first residue (1037 1x) showed an actual decrease of MIC
43

whereas deletion of the first two residues (1037 2x) showed a slight increase. It should be
considered that these peptides were not synthesized in parallel with 1037 alanine scanning peptides
and that these peptides have not but purified.
I suspected that the substitution of the hydrophobic residues led to a destabilization of the secondary
structure, and the I5 was substituted with V (1037 V5) or L (1037 L5). These substitutions showed
no complete disruption of antimicrobial activity but it was also showed that these amino acids were
not ideal replacements.
The insertions of a P in a peptide will lead to a break in the amino acids structure. P was substituted
for R4 and R6 P6 completely disrupted the antimicrobial activity of the peptide whereas P4 showed
only a slight decrease in antimicrobial activity compared to the respective alanine analog.
Table 6: MIC determination of 1037 and synthetic analogues. a Generally accepted nomenclature. b Single letter amino
acid code of the peptide. c MIC determination of the tested antimicrobials with least 3 replicates. d Charge of the peptide at
pH 7.

Ratio of hydrophobic/total number of residues (%) calculated with innovagen peptide calculator. f Mean relative

hydrophobic moment (mH) calculated with HydroMCalc using Eisenberg scale. Strains were P. aeruginosa PAO1, K.
pneumoniae C3091, E. coli 536 and CFT073.

Name

Sequence

MIC (g/ml)c

Charged %e mHf

PAO1

C3091

536

CFT073

1037

16

16

44

0.3

1037 A1

32

16

32

16

55

0.12

1037 A2

16

32

16

32

55

0.4

1037 A3

>256

>256

128

256

44

0.25

1037 A4

128

128

64

64

55

0.33

1037 A5

>256

>256

128

256

44

0.36

1037 A6

64

32

16

32

55

0.52

1037 A7

256

64

64

64

44

0.27

1037 A8

128

64

32

64

55

0.15

1037 A9

64

32

16

16

44

0.31

16

16

50

0.17

64

128

64

64

57

0.33

1037 1x
1037 2x
1037 P6

>256

>256

256

256

44

0.48

1037 P4

128

256

128

128

44

0.31

1037 V5

64

128

16

16

44

0.32

1037 L5

64

32

16

44

0.32

1037 Reverse

32

32

16

44

0.3

44

4.5 Effect of LL-37, 1037 and antibiotics on P. aeruginosa biofilms


P. aeruginosa formed a strong biofilm with the vast majority growing on the sides of the wells at
the air/liquid interface. After staining with crystal violet, the average untreated wells were measured
to 2.26 at OD595. It is important to note that crystal violet is not selectively binding to the biofilm
but binds to all substances in the well. This assay is therefore a measurement of biomass, including
both attached cells and EPS. The presences of EPS also influence the number of bacteria which is
able to attach, so a high value corresponds to high level of biofilm formation.
LL-37 showed to decrease the amount of P. aeruginosa biofilm by 56 % at MIC (32 g/ml) and the
inhibition was statistically significant down to a concentration of 2 g/ml (1/16x MIC) where the
biofilm was reduced by 31.9 % (Figure 11A).
1037 decreased P. aeruginosa biofilm formation with 57.1 % at MIC (16 g/ml) and a 77.8 %
inhibition at 128 g/ml. Inhibition of biofilm formation was observed at concentrations as low as 2
g/ml (1/8 of MIC) but was showed not to have an effect at 4 g/ml (Figure 11B).
Tobramycin did not inhibit P. aeruginosa biofilm formation but was increased by 42.4 % at 2x MIC
(Figure 11C). Tetracycline showed no reduction in P. aeruginosa biofilm formation but at x MIC,
tetracycline increased P. aeruginosa biofilm formation with 34. 6 % (Figure 11D).

Figure 11: Effect of LL-37, 1037, Tobramycin and Tetracycline on P. aeruginosa biofilms (A-D). Inhibition of P. aeruginosa
biofilm formation was demonstrated by LL-37 (A) and 1037 (B), while tobramycin (C) and Tetracycline (D) showed no inhibition.
The data is expressed as percentage of the positive control with mean standard derivation and is the result of at least three
independent trials. Statistical significance is indicated with asterisks (* p < 0.01, ** p < 0.001, *** p < 0.0001).

45

For further examination of LL-37 and 1037s antibiofilm effect, the incubation time was changed to
two hours to determine a potential correlation with inhibition of initial attachment. The attachment
of P. aeruginosa was decreased in the presence of LL-37 by 35.8 % at MIC (32 g/ml) but at subMIC concentrations the majority of the samples were close to 0 % (Figure 12A). 1037 was more
potent and inhibited the initial attachment by 29.5 % at MIC (16 g/ml) and no attachment was
detected at 128 g/ml.

Figure 12: Effect of LL-37 and 1037 on P. aeruginosa initial attachment (A-B) and on preformed biofilms (C-F). Inhibition of
LL-37 (A) and 1037 (B) on P. aeruginosa initial attachment (no replication). LL-37 and 1037 were added to P. aeruginosa biofilms
that have been growing for 24h and was incubated for 24h (C+E) or 2h (D+F), done in quadruplets (no replication). Statistical
significance is indicated with asterisks (*** p < 0.0001).

46

The effect of LL-37 on preformed P. aeruginosa biofilms was indicated to be statistically


significant at 16 g/ml with a biofilm reduction of 55.6 % (Figure 12C). At 4 g/ml, the amount of
biofilm was reduced 17.2 % and was statistically just above the significance level. The effect after 2
hours of incubation showed a slight increase in biofilm produced, thus LL-37 was indicated not to
be fast acting (Figure 12D).
1037 was indicated to not affect preformed biofilms, both at 2h and 24h incubation time (Figure
12E-F).
Whereas the first two residues of 1037 were shown to be less important for antimicrobial activity,
the alanine scanning library of 1037 revealed that, they were very important for the antibiofilm
effect (Table 7). This was indicated to be the case with most of the residues. R6 was indicated to
have the least importance in the antibiofilm, since 1037 A6 inhibited P. aeruginosa biofilms with
36.2 % at 8 g/ml. At some concentrations, the biofilm was even shown to increase the amount of
biofilm formed (Table 7).
a

Table 7: Effect of 1037 Alanine library against P. aeruginosa biofilm formation.


c

Generally accepted nomenclature.

Single

letter amino acid code of the peptide. MIC against P. aeruginosa. Inhibition of biofilm formation in percentage compared to the
positive control at MIC and at 8 g/ml. + indicates an increase in the biofilm formation. For MIC that exceeds the 256 g/ml
detection limit, the MIC was taken as to 256 g/ml when calculated MIC. The data is the result of a single experiment with no
duplicates.

Biofilm inhibitiond
Name

Sequence

MIC (g/ml)

MIC

8 g/ml

1037

16

46.5

46.5

1037 A1

32

6.2

6.6

1037 A2

16

9.5

9.5

1037 A3

>256

52.4

+21.7

1037 A4

128

+8.2

14.7

1037 A5

>256

54.5

6.2

1037 A6

64

25.7

36.2

1037 A7

256

39.2

20.5

1037 A8

128

20

9.3

1037 A9

64

13.7

16.7

47

4.6 Effect of LL-37 and 1037 on K. pneumoniae biofilms


The inhibition of K. pneumoniae biofilms by LL-37 was first assessed in five different media
(Figure 13A-E). K. pneumoniae C3091 have previously shown to form a biofilm in M9 minimal
medium (Struve et al., 2009), but to my knowledge this medium is not used for peptide
experiments. BM-2, TB, MH and LB media have all three been used to show an antibiofilm effect
of LL-37 and was therefore found qualified to be included for screening (Overhage et al., 2008,
Kai-Larsen et al., 2010, Dean et al., 2011b, Pompilio et al., 2011).

Figure 13: The effect of LL-37 on K. pneumoniae biofilm formation in different media (A-E). The experiments were conducted
in round-bottomed polypropylene 96-well microtiter plates as singles with at least three replications and incubated 24h.
Experiments that were considered as major outliers were removed from the graph. The different media were BM2 (A), M9 (B), LB
(C), MH (D) and TB (E).

48

Since polystyrene microtiter plates are charged, they can bind some peptides. Therefore the first
assessment of LL-37s antibiofilm activity was carried out in polypropylene microtiter plates. K.
pneumoniae formed a biofilm on the bottom of the plate. The result was that much of the bound
biofilm was removed by pipetting. The results from Figure 13A-E clearly show a great variance
within many of the samples. Some of the results were also deleted due to the positive control was
too low and was considered interrupted, which in particular was a big problem with M9 minimal
medium.
With this taken into account, the subsequent experiments were conducted in flat bottomed
polystyrene plates. With the object to analyze the time growth of K. pneumoniae biofilm formation,
K. pneumoniae was grown in BM2 and M9 media and analyzed after 2, 4, 6 and 24 hours (Figure
14A-B). BM2 medium produced the highest amount of biofilm after 2 and 24 hours (Figure 14B)
but the variation after 24 hours was rather big.

Figure 14: The time growth of K. pneumoniae biofilm formation in M9 (A) and BM2 (B) media. The experiment was run as
parallel on the same 96-well microtiter polystyrene plate and shows the amount of biofilm produced by K. pneumoniae at different
time points in M9 (A) and BM2 (B) media.

Due to this observation and due to the no-inhibition trend (Figure 13), a single experiment with LL37 and 1037 was conducted in polystyrene plates with BM2 medium to verify this trend. For LL-37
(Figure 15A), the trend seems to be no inhibition at all. The highest concentrations, 128 and 64
g/ml shows a slight increase, while the two of the lowest concentrations, 2 and 4 g/ml, show a
slight decrease. For 1037 (Figure 15B), the trend seems to be a slight decrease of biofilm at the
highest concentrations, but nothing that occur significant.

49

To see if the initial of K. pneumoniae biofilm formation is interrupted, the incubation time was
adjusted to two and four hours. K. pneumoniae was significant effected by LL-37 down to 2 g/ml
(Figure 15E), but after four hours, the amount of biofilm was completely restored (Figure 15C). The
two hour effect of 1037 was more potent (Figure 15F), and the effect was indicated to last for a
longer period of time (Figure 15D).

Figure 15: The effect of LL-37 and 1037 on K. pneumoniae biofilms in 24h (A-B), 4h (C-D) and 2h (E-F). The experiments were
conducted in flat-bottomed polypropylene 96-well microtiter plates. For 24h (A-B) and 4h (C-D) samples, no replication was carried
out. For 2h incubation (E-F), the data is the representative for 3 individual experiments. Statistical significance is indicated with
asterisks (* p < 0.01, ** p < 0.001, *** p < 0.0001).

50

Since LL-37 and 1037 seems to inhibit the biofilm formation at short time of incubation, the effect
on preformed K. pneumoniae biofilms was examined with a three hour incubation time, but the
result showed a slight increase in biofilm, Figure 16.

Figure 16: Effects of LL-37 and 1037 on preformed K. pneumoniae biofilms. The biofilm was grown for 24h and incubated with
peptide and fresh medium for 3h. Experiment was done in quadruplets but with no replication.

Too see if this temporary reduction of attached cells and biofilm produced after 2 hours, a time-kill
experiment was conducted (Figure 17). LL-37 was very efficient at 2x and 4xMIC but less efficient
at 1xMIC (Figure 17A). After 24 hours, the CFU/ml was restored and was comparable with the
positive control.
1037 showed no decrease in CFU/ml at all, not even at 4xMIC (Figure 17B). This seems strange,
and it the experiment should be reproduced.

Figure 17: Time-kill kinetics of LL-37 (A) and 1037 (B) on K. pneumoniae. The bacteria were grown in the same condition as for
the biofilm experiment. After each time point, the bacteria was isolated from the wells and diluted and plated on LB agar plates.

51

4.7 Effect of LL-37 and 1037 on E. coli biofilms


E. coli biofilms was observed, as for K. pneumoniae, to grow on the bottom of the microtiter plate.
The average growth (OD595) for E. coli strain 536 was 0.696 in LB and 0.435 in DMEM and for E.
coli strain CFT073 it was 0.937 in LB and 0.321 in DMEM. By observation, it seemed that the
effect by the peptides on the biofilm was the same for the two media. The same hurdle as for K.
pneumoniae, where the biofilm growth on the bottom of the plate lead to an assay source of error,
was observed for E. coli. The biofilm growing on the bottom made the experiment less reliable and
some data was, due to a low positive control, removed. The experiment was conducted in two
different media, LB and DMEM, which was combined to present the data (Figure 18). LL-37 was
shown to have a statistically significant effect on both E. coli strains (Figure 18A-B). The inhibitory
concentration was as low as 2 g/ml where E. coli strain CFT073 was inhibited by 29.3 %.
1037 was shown to be more efficient versus E. coli strain 536 than CFT073 (Figure 18C-D). For E.
coli strain 536 the inhibitory concentration was 2 g/ml where the inhibition was 26.9 %.

Figure 18: Effects of LL-37 (A-B) and 1037 (C-D) on E. coli biofilm formation. The antibiofilm effect of LL-37 on E. coli strains
536 (A) and CFT073 (B) and 1037 on E. coli strains 536 (C) and CFT073 (D). The data are the result of at least 3 replications in
either LB or DMEM media with incubation of 21h. Statistical significance is indicated with asterisks (* p < 0.01, ** p < 0.001, *** p
< 0.0001).

52

5. Discussion
With the lack of discovery and development of antibiotics the last decades, Figure 1, scientists are
looking for ways to expand the quantity of anti-infective drugs. This is a challenge that is becoming
increasingly crucial for keeping up with the development of resistant pathogenic bacteria. AMPs are
by many scientists presented as a future solution to this problem but their clinical potential is often
rejected or met with skepticism.
The first AMPs were discovered in the 1980ies, including a wide variety of AMPs secreted by frog
skin (Zasloff, 1987, Bevins and Zasloff, 1990, Ge et al., 1999). The most notable characteristic of
frog skin AMPs is that they are found to be secreted in much larger amounts than in mammalians.
Since frogs live in constant contact with bacteria, the secretion of AMPs serve as a strong indication
of their immunological importance. Since this observation, Dr. Michael Zasloff has put a lot of
effort in optimizing and bringing frog magainins into the pharmaceutical industry and has
developed Pexiganan, an AMP which is active against various clinical isolates (Ge et al., 1999).
Pexiganan was in the late 1990ies in clinical trials but is currently pending and the project is
seeking further funding. Meanwhile other (unrelated) peptides, such as a synthetic analog of
indolicidin, MBI 226, and a synthetic analog of an AMP (plectasin) isolated from Pseudoplectania
nigrella fungus, NZ2114, have in the latest years caught attention (Sader et al., 2004, Xiong et al.,
2011). All three peptides have in common that they have been in phase trials, they are efficient
against clinical drug resistant isolates, and interestingly they are all derivatives of original isolated
AMPs.

In these experiments LL-37 and 8 LL-37 analogs, seven derivatives and one derivative with a single
residue substitution, are screened for P. aeruginosa and K. pneumoniae antimicrobial activity.
The antimicrobial activity of LL-37 was weak against P. aeruginosa and K. pneumoniae but good
against E. coli (Table 5, p. 42). Recently, other studies have found the LL-37 to be 12.5 g/ml, 31
g/ml, 64 g/ml and 112 g/ml (25 M) (Bucki et al., 2008, Overhage et al., 2008, de la FuenteNunez et al., 2012a, Nagant et al., 2012). This indicates that the MIC value should be treated with
cautious consideration. The experimental setup can vary depending on laboratory, for example the
MIC was 31 g/ml in BM2 (de la Fuente-Nunez et al., 2012a), 12.5 g/ml and 112 g/ml in MH
(Bucki et al., 2008, Nagant et al., 2012). Taken in mind that the two studies use P. aeruginosa
PAO1, it seems unlikely that the MIC can vary from 12.5 g/ml to 112 g/ml. It is possible that the
high MIC of LL-37 is the result of the abundant use and widespread of P. aeruginosa PAO1 which

53

is causing an ongoing microevolution of genotype and phenotype (Klockgether et al., 2010). The
experiment which determinates the MIC of LL-37 to 112 g/ml should be treated with caution since
all their tested peptides show decreased antimicrobial activities (Nagant et al., 2012). It is likely
that the MIC values for LL-37 shown here are more correct, since they are in range with the
previous discussed values. But it is important to notice that the MIC value is not a static
measurement. For example did I observe some outliers in my experiments, and one replicate
showed an MIC at 64 g/ml.
My results support the conception of AMPs as cooperative components of the innate host defense
whose effects are additive and synergistic, resulting in an efficient defense against intruding
microorganisms. These additive effects are best illustrated in the saliva of the oral cavity where 45
AMPs, such as defensins, histatins and LL-37 have been identified, whereas some are present in
low concentrations (Fabian et al., 2012, Gorr, 2012).

Compared to LL-37, KR-12 (KRIVQRIKDFLR) was the only derivative which showed an
increased antimicrobial activity, but only for P. aeruginosa (12.5 g/ml) and not for K. pneumoniae
(>200 g/ml) (Table 5, p. 42). Compared to FK-12 (FKRIVQRIKDFL), which is overlapping with
only one N-terminal residue, the MIC of KR-12 is decreased 8-fold. This argues for residue R29
being important for antimicrobial activity and F17 less important. The improvement of antimicrobial
activity should be seen in contrast with KR-12 having a molecular weight more than half as low as
LL-37, and when converted into molar concentration, the MICs are almost the same, 8 M for KR12 and 7.1 M for LL-37. It also shows the limitations of the micro-broth dilution method where
the test concentrations are doubled theoretically, LL-37 could have an MIC against P. aeruginosa
at 17 g/ml in contrast to the observed 32 g/ml.
VQ-12 (VQRIKDFLRNLV) is overlapping KR-12 C-terminal wise and the antimicrobial activity is
completely lost (MIC >200). The three amino acids, KRI18-20, appear to be more important for
antimicrobial activity than previously discussed R29. These results are in agreement with novel
structural experiments of LL-37 derivative GF-17 ((G)FKRIVQRIKDFLRNLV) and GE-18
((G)EFKRIVQRIKDFLRNLV) which show that K18, R19, and R29 were important for antimicrobial
activity (Wang et al., 2012). The results from Table 5 show that fragments without K18, R19, and R29
lose antimicrobial activity. In the study by Wang et al, 2012, these residues were by alanine
scanning not indicated to be the most important residues, but only doubled the MIC. R23 was most
important in killing S. aureus whereas R23 and K25 were more important in killing E. coli. K18, R19,

54

and R29 were all shown to be on the hydrophilic surface of the peptide, and it is possible that the
deletion of these residues could alter the structure of the peptides tested in these experiments.
The observed importance of R29 to LL-37s killing mechanism was supported by another study
which identified KR-12 (KRIVQRIKDFLR) as the smallest LL-37 fragment with maintained
antimicrobial activity (Wang, 2008). KR-10 (KRIVQRIKDF), KR-11 (KRIVQRIKDFL) and RI-10
(RIVQRIKDFL) showed no or little antimicrobial activity, which supports our observations with
FK-12 (lacking R29) and VQ-12 (lacking KRI18-20).
These results also make sense in contests to naturally found derivatives of LL-37, such as KR-20
(KRIVQRIKDFLRNLVPRTES) which is processed after secretion onto the skin surface and show
higher antimicrobial activity than LL-37 (Murakami et al., 2004).
These results also indicate that D26 is less important for antimicrobial activity since VQ-12V6
(VQRIKVFLRNLV) has >4 times lower MIC for P. aeruginosa. This is likely to be because VQ-12
(VQRIKDFLRNLV) is less cationic and a small fragment of LL-37 and replacement of anionic D26
increases the cationicity, thus increasing antimicrobial activity. Since the activity of many AMPs is
based on their amphipathic nature, correlations between cationicity and hydrophobicity and
antimicrobial activity have been suggested (Hilpert et al., 2009). The measured MICs (Table 5) are
mostly very high, based on very few peptides and it is therefore difficult to make anything but weak
indications based on these results. This taken in mind, the correlation between cationicity and MIC
for P. aeruginosa cannot be completely denied (r2 = 0.5741), and in contrast, the correlation
between hydrophobicity, the mean relative hydrophobic moment and the tested LL-37 derivatives
are non-exciting (r2 = 0.0161) (Eisenberg et al., 1982), see Figure 9A-B and Table 5.

The MIC of the 9-residue peptide, 1037, was determined to 16 g/ml, which is in striking contrast
to the originally determined MIC value at 304 g/ml. As mentioned above, the MIC at 16 g/ml
was confirmed using the same strains as in the original experiment (de la Fuente-Nunez et al.,
2012a). Their study was based on a screening of a peptide library which originated with the
consensus peptide, FRIRVRV. 1037 (KRFRIRVRV) was by far the worst peptide, but interestingly,
peptide 1029 (KQFRIRVRV) had a MIC of 10 g/ml, peptide HH10 (KRFRIRVAVRRA) had a
MIC of 0.8 g/ml and HH15 (KRFRIRVRVIRK) had a MIC of 12 g/ml. 1037 and 1029 differ in
sequence only in R2 which would mean that Q2 of 1029 was of crucial importance to antimicrobial
activity. In contrast it was shown that HH15 (KRFRIRVRVIRK), which was 1037 with the Nterminal addition of IRK, and the KR sequences is thereby indicated not to negatively affect

55

antimicrobial activity. The only plausible reasons for this observation could be that the peptides
coincidently have antimicrobial activity by two difference mechanisms, or that this peptide is
structural dependent and that 1037 is more affected by the BM2 medium, though it seems very
unlikely. This study has already shown that peptides with a single overlapping residue can have
completely different antimicrobial activities (Table 5), and it needs to be taken into consideration
that the original study determined the MIC of all peptides in BM2 medium. Even this taken into
account, it is unlikely that only one peptide of a set of related peptides shows strong media
dependence.
The alanine scanning of 1037 shows that K1 and R2 residues of 1037 are of minor importance, and
even with the removal of K1, 1037 1x (RFRIRVRV), the antimicrobial activity was preserved, and
this further questions the previous results, since R2 show not to decrease antimicrobial activity.
The only apparent explanation to these crucial differences is that, one of the studies has weighted
the peptide wrongfully. Their studies show that 1037 decrease P. aeruginosa biofilm formation with
78 % at 152 g/ml, and the peptide was identified with the highest biofilm inhibition at their stated
x MIC. This is exactly the same inhibition observed in this study (Figure 11). Thus it is most
likely that the original study of 1037 wrongfully determined the MIC of 1037 and that the biofilm
inhibition potential of 1037 at x MIC is worse than originally measured.
The results from Figure 10 and Table 6 show that the hydrophobic residues F3, I5 and V7 are crucial
for antimicrobial activity of 1037 and that the hydrophobic V9 is of less importance. This could be
because that 1037 depends on its secondary structure, and -helical peptides are often helical in the
center and not organized in the ends, which in this case probably would mean a single helical turn
in the center.
The positively charged residues, K1, R2, R4, R6 and R8 are not affected as much as the hydrophobic
residues when replaced with alanine. The decrease in antimicrobial activity can be explained by a
decreased electrostatic attraction between the positively charged peptides and negatively charged
bacterial membrane which contains anionic molecules such as LPS in Gram-negative bacteria, and
teichoic acids in Gram-positive bacteria (Powers and Hancock, 2003). But since the cationicity is
high for 1037, it is more likely to be explained by loss of structure by making the peptide structure
unstable.
An insertion of a proline will make a break and completely disrupt peptide structure. The MIC of
1037 A4 and 1037 A6 is increased when R4 and R6 are replaced with proline, with the R6 being

56

affected most. This argues for the structure being important for the peptide function and for a
structure, which is not highly dependent on hydrophilic residues. The antimicrobial function was
almost restored when I5 was replaced with leucine and valine.
The hypothetical helical wheel (Figure 19) shows that the hydrophobic residues cluster and that the
1037 peptide forms into an amphipathic nature. Often smaller hydrophobic amino acids are
expected to adopt an -helical conformation with 3.6 amino acids pr. turn, and larger residues such
as phenylalanine, isoleucine and valine can adopt both -sheet and -helical conformations (Munoz
and Serrano, 1994). Even though 1037 consist of large hydrophobic residues that often adopts a sheet conformation, it is unlikely that 1037 forms a -sheet since -sheets form adjacent hydrogen
bonds and would essentially need to make a -turn, which often requires a proline or glycine.
However, since a related peptide from the same peptide library, 1018 (VRLIVAVRIWRR), was
indicated to form an -helical conformation (Wieczorek et al., 2010), I suggest that 1037 is also has
the ability to form an -helix.

Figure 19: Helical wheel diagram for 1037. Orange circles mark hydrophobic residues and green circles mark hydrophilic residues.
Constructed by using the helical wheel plotting program from http://www-nmr.cabm.rutgers.edu.

The formation of biofilms has a huge clinical impact on the medical capability to treat infections.
A biofilm protects the bacteria from polymorphonuclear leukocytes, adds a protective layer that
restricts the access of antibiotics and up-regulates resistance genes (Hoiby et al., 2010, Whiteley et
al., 2001). More importantly, cells deep within a biofilm are dormant, and since many antibiotics
only are able to target metabolic active cells, these cells are called persisters. Consequently, cells
living in a biofilm need a higher antibiotic dosage to be eradicated (Hoiby et al., 2011).

57

The argument to study the antibiotics tobramycin and tetracycline, were based on results from an
independent study (Pompilio et al., 2011), which showed tobramycin to be active against P.
aeruginosa and S. aureus biofilms at x MIC.
In this study, tobramycin and tetracycline were not observed to reduce P. aeruginosa biofilms, in
contrast the biofilm formation was induced by 42.4 % at 2x MIC tobramycin and 34 % by
tetracycline at x MIC (Figure 11C-D). In the biofilm experiments a higher bacteria load and
different media were used compared to the MIC experiments, which make it possible to study the
antibiotics and peptides in concentrations above MIC. Tetracycline is a bacteriostatic drug and a
gradient of pellet size indicated that the samples differed in cell number, but there was not,
however, conducted a time-kill kinetics experiment to verify inhibition of cell growth.
It is evident that different strains respond differently to the subinhibitory concentrations of
tetracycline and tobramycin. In a study, only half of the P. aeruginosa clinical isolates from cystic
fibrosis patients showed sub-MIC inducible biofilm formation by tobramycin (Elliott et al., 2010).
In an environmental P. aeruginosa strain, RhlI/R quorum sensing system was inhibited by
tobramycin (Babic et al., 2010) and with P. aeruginosa PAO1, tobramycin induced biofilm
formation by inducing aminoglycoside response regulator gene (arr) gene which effects cell
adhesion (Hoffman 2005). Tetracycline activates rhlA and lasB quorum sensing in P. aeruginosa at
subinhibitory concentrations (Liang et al., 2008) and at sub-MIC levels, -Lactam antibiotics induce
S. aureus biofilm formation (Kaplan et al., 2012).

For LL-37 P. aeruginosa biofilms were inhibited 58.4 % at x MIC (16 g/ml) and statistically
significant down to 2 g/ml where the biofilm formation was reduced 32.1 % (Figure 11A). These
results are comparable with the previous determined antibiofilm activity of LL-37 (Table 2, p. 31).
There was no effect of LL-37 on P. aeruginosa initial attachment below MIC (32 g/ml) (Figure
12A). The effect of LL-37 on pregrown biofilms at 16 g/ml was 55.6 % inhibition (Figure 12C).
Previous studies have shown inhibition of pregrown biofilms down to concentration of 4 g/ml, but
these experiments were performed in flow-cells (Overhage et al., 2008). Pregrown biofilms were
shown not to be affected by the presence of LL-37 after 2 hours of incubation (Figure 12D). This
indicates that LL-37 does not affect P. aeruginosa biofilms in the short term. In contrast a study
shows that LL-37 inhibited the 1 hour attachment and connected it to the LL-37 mediated induction
of twitching motility (Overhage et al., 2008).

58

1037 have previously been shown to inhibit P. aeruginosa PAO1 and PA14, Burkholderia
cenocepacia and Listeria monocytogenes biofilms by inhibiting the flagella-dependent swarming
motility and the twitching motility (de la Fuente-Nunez et al., 2012a). At concentrations 10 g/ml,
15 g/ml and 25 g/ml, P. aeruginosa PAO1 was reduced by 60 % and 78 % at 152 g/ml. These
results were somewhat confirmed in this study. At concentration 8 g/ml and 16 g/ml the amount
of biofilm was reduced by 46.5 % and 57.1 %, respectively, and the inhibition was even greater at
32 g/ml and 64 g/ml (Figure 11B). The initial attachment was indicated not to be inhibited to the
same extend and it was indicated that 1037 had no effect on pregrown biofilms at 2h and 24h after
(Figure 12BDF).
All alanine scanning peptides showed less antibiofilm potency at 8 g/ml. The cationic residues
were shown to be of less importance and the hydrophobic residues were of greater importance in
antimicrobial activity (Figure 10A-D). In contrast there was no similar trend for the antibiofilm
activity (Table 7).

Many antimicrobial peptides and proteins are secreted from urinary tract epithelial cells producing
an antibacterial shield of the human urinary tract (Zasloff, 2007, Spencer et al., 2013).
LL-37 is constituently expressed in urinary tract epithelial cells and can be detected in urine with
median concentration at 0.3 ng/ml but the level is increased 8 times in children with pyelonephritis
or cystitis (Chromek et al., 2006). Nevertheless, it is thought that the concentration of LL-37 and
other AMPs is significantly higher in the fluid layer which is in direct contact with the luminal
surface. The rapid increase of LL-37 mRNA expression after E. coli exposure suggests that this
peptide is highly regulated by bacteria (Chromek et al., 2006). The same study also showed that
CRAMP-deficient mice have more bacterial bladder attachment.

The study of K. pneumoniae biofilms were shown to be highly affected by the assay itself. In
polypropylene plates, the biofilm was very unstable and caused high standard deviations (Figure
13A-E). This consequently changed the setup of the assay, and a polystyrene microtiter plate had to
be employed instead. These plates are not optimal for studying cationic AMPs, since polystyrene
and tissue-culture treated plates are negatively charged and can be a deciding determinant on
antimicrobial assays (Wiegand et al., 2008). Another affected deciding determinant is the medium
used for the experiments. Previously K. pneumoniae biofilms were observed to grow well in M9
medium (Struve et al., 2009) but this medium is not popular for peptide effect determination. Since

59

BM2 have been used to identify LL-37 and 1037 inhibition of P. aeruginosa biofilms (de la FuenteNunez et al., 2012a), and since biofilms were observed to grow faster (Figure 14A-B), this medium
was chosen for further experiments.
In polystyrene microtiter plates, the 24 hour inhibition of 1037 and LL-37 on K. pneumoniae
biofilms was initially denied (Figure 15A-B). LL-37 was indicated to moderately inhibit the biofilm
at 2 g/ml and 4 g/ml but remarkably, the biofilm formation was increased with the concentration
of LL-37. 1037 showed low or no inhibition of the biofilm formation.
In relation to the polypropylene experiment, there might be an effect of LL-37 on K. pneumoniae
biofilms (Figure 13B), but this trend was not confirmed in polystyrene plates. Since K. pneumoniae
C3091 is an UTI isolate, and taken in mind that LL-37 is expressed in the urinary tract, it is possible
that the cells have developed resistance towards AMPs. Consequently, these experiments focused
on the 2 and 4 hour effect on the initial attachment and biofilm formation (Figure 15C-F). Both LL37 and 1037 strongly inhibited the initial attachment, both as low as 2 g/ml (Figure 15E-F). This
effect was somewhat neutralized for LL-37 after 4 hours at low concentrations (Figure 15C), but it
was maintained for 1037 (Figure 15D). This indicated that these peptides could be efficient against
K. pneumoniae, but only at the initial stages of infection. In contrast, there was no effect when the
biofilm were pregrown and incubated for 3 hours with LL-37 or 1037 (Figure 16). It was
hypothesized that these AMPs would have an effect on the cell growth, even close to MIC. Since
the medium and bacterial load were different, the inhibition on cell growth were assessed over time
at 1x, 2x and 4x MIC (Figure 17A-B). LL-37 was shown to initially inhibit the cell growth, but after
24 hours the number of cells was restored (Figure 17A). 1037 showed surprisingly no inhibition of
planktonic growth at all, even at 4x MIC (Figure 17B). This is truly hard to believe, but it should be
taken into account, that this result was not replicated and needs to be replicated for further analysis.
It might be due to the peptide being strongly influenced by the salt content of the medium, which
could explain the MIC of 304 g/ml from the previous mentioned experiment (de la Fuente-Nunez
et al., 2012a), but it seems unlikely.

LL-37 slightly inhibited E. coli biofilms (Figure 18A-B). This effect was dose dependent for strain
536 but seemed random for strain CFT073 biofilms. 1037 was more effective than LL-37 (Figure
18C-D), but only against 536 biofilms.
Another study has shown an 80 % inhibition at 11.2 g/ml on an E. coli isolate from a child with
pyelonephritis (Kai-Larsen et al., 2010). This inhibition was in fact not found to be the case with E.

60

coli 536 and CFT073. LL-37 could potentially be almost good against E. coli biofilm, but in these
experiments the standard deviations were too big, and the biofilm inhibition trend was often not
comparable between the experiments.

The biofilm results expose the inadequacy of this type of experiment. Evidently, the results from
this type of experiment are reflected by the type of abiotic material used for bacterial attachment
and peptide disablement, the medium, bacterial load and an essential thing such as pipetting.
The pipetting for these experiments makes up for the biggest source of error, since it only took little
force to the well, before the biofilm fell off. My biggest mistake for these experiments was that I
initially washed the biofilms too much which consequently removed most of the biofilm from the
wells. As a result, the protocol for these biofilm experiments was changed so the washing procedure
became as simple as possible. The succeeding experiments (the results presented) still showed a big
variation. Most important was the positive control, since bad pipetting of the positive control would
affect the experiment as a whole. Therefore most experiments were performed with three positive
controls, but mistakenly no replica of the test wells. This is something I would improve if the
experiments were to be reproduced. The focus should only be on sub-MIC concentrations, maybe
even as low as MIC, since the peptide should not affect the reproduction of the cells. On the other
hand, another study has shown big variations in an in vivo attachment assay indicating that the
standard deviations in these experiments are common (Tsai et al., 2011).

These results do not strongly support an antibiofilm mode of action by LL-37 and 1037 against the
tested strains. All three strains are clinical isolates and it is possible that they have developed
resistance towards LL-37 and other AMPs since they are required to tolerate exposure antimicrobial
peptides at the site of infection. The MIC of E. coli strains correlates with more invasive infection
of the upper urinary tract, pyelonephritis, compared to infection of the lower urinary tract, cystitis
(Chromek et al., 2006). Resistance of the tested strains towards LL-37 and 1037 seems unlikely
since the MICs are relatively low (Table 5 & Table 6). Contrarily, if LL-37 was proven to
efficiently fight the bacteria in vitro, it could be the result of the immune system suffering from
local stagnation as a result of local temperature decrease or decreased expression of LL-37, which
in fact has been linked to some types of infections (Vandamme et al., 2012).

61

To sum up, LL-37 significantly inhibits E. coli CFT073 biofilms down to a concentration of 2
g/ml and E. coli 536 at 32 g/ml. 1037 inhibits E. coli 536 biofilms at 2 g/ml but does not affect
E. coli CFT073 at any concentration. K. pneumoniae C3091 was inhibited at 2 g/ml by LL-37 and
1037 after 2 hours but this effect could for LL-37 be explained by a reduced cell number. No timekill kinetics experiments were conducted on E. coli but the effect of LL-37 and 1037 on cell
replication at 2 g/ml after 24 hours would most likely be low and it cannot explain the reduced
biofilm production.
As for P. aeruginosa, it could be strategically clever to target quorum sensing in E. coli and K.
pneumoniae but those species have been observed to secrete low or no AHLs when forming a
biofilm on indwelling urethral catheters (Stickler et al., 1998)
Another effective strategy would be to block the adhesion to the epithelial surface or the indwelling
catheter (Klemm et al., 2010), by blocking the mannose binding pocket of FimH to prevent
adhesion of type 1 fimbriae (Chen et al., 2009, Wang et al., 2010) or by blocking the type 1
fimbriae. The curli fimbriae major subunit (CsgA), have previously been shown to be targeted by
LL-37 but it is likely that LL-37 and 1037 have several targets (Kai-Larsen et al., 2010).
The essential question to answer is how big the therapeutically potential of AMPs, such as LL-37
and 1037 are. This question encompasses important factors such as resistance, resistance
development, pharmacokinetics and pharmacodynamics.

The biofilm results presented here also show that the AMPs differ in level of activity when tested
on different bacterial species and strains. The elementary component of the bacterial cytoplasmic
membrane is the amphipathic phospholipid bilayer which provides a target for the AMP. The net
charge of the bacterial membrane is based upon its composition and architecture which differ
among bacteria species. A well-known resistance strategy is to alter the bilayer chemically, for
example altering Lipid A of LPS, which makes the membrane less anionic and therefore the peptide
less likely to interact with the membrane (Yeaman and Yount, 2003, Gutsmann et al., 2005).
Other resistance strategies have had direct effect on peptides from this study. For example is the Slayer of Caulobacter crescentus providing a barrier from LL-37 and 1037 which increases the MIC
(de la Fuente-Nunez et al., 2012b). In K. pneumoniae, LL-37 is specifically bound to the capsular
K40 polymer and EPS polysaccharide alginate of P.aeruginosa (Herasimenka et al., 2005,
Foschiatti et al., 2009, Benincasa et al., 2009). This is an example of the fact that peptides, such as
LL-37, needs selectivity for their specific target. Since strong biofilm forming bacteria will secrete a

62

lot of polysaccharides, it is not very convenient to target that part of the biofilm. The peptide
concentration is stagnant or decreasing and cannot prevent the biofilm to form at increasing
polysaccharide secretion. The interaction with polysaccharides could explain LL-37s weak activity
towards multidrug resistant clinical strains that produce a strong biofilm (Pompilio et al., 2011).
But the most crucial factor is whether the peptide is stable when administered, considering that the
peptide stability is challenges both by the host and pathogen responses.
For example, OmpT, an outer membrane protease of E. coli , cleaves LL-37 between R7-K8 and
K18-R19, which according to my results in Table 5 would inactivate the peptide (Thomassin et al.,
2012). The expression levels of OmpT were furthermore shown to be high in some E. coli strains
and low in others. Also human host proteases have been mentioned to cleave LL-37, for example in
the skin, but it is not known whether this is a type of inactivation and if this type of bacteria-peptide
dynamic taking place anywhere else in the human body. In the liver, many AMPs would also be
neutralized by proteases, especially peptides containing L-amino acids.
Another downside is that many AMPs have been reported to be hemolytic. This problem arises
from the negative charge of the red blood cells of which positively charged amino acids are likely to
interact with and consequently lyse. Therefore many scientific experiments determine the hemolytic
activity of AMPs to evaluate the therapeutic potential. However, this assay is misleading since
peptides can bind to many components of the blood. For example, LL-37 have shown not to be
cytotoxic to red blood cells when tested in serum due to the binding of lipoproteins, but this
consequently also disabled the antibacterial activity of LL-37 (Johansson et al., 1998, Larrick et al.,
1995a). It is also convenient that lysed LL-37 yielding KR-20 show no hemolytic activity,
indicating that KR-12 would be less toxic (Murakami et al., 2004).

On the contrary AMPs have many advantages that antibiotics often lack. They can be used as single
antibiotics or in combination with other antibiotics. Evidently it would be effective to combine
treatments with antibiofilm and conventional drugs. Studies on P. aeruginosa have shown
promising results for the future. Inhibiting of las quorum sensing would be the most efficient way to
decrease the biofilm formation since las quorum sensing controls the rhl system (Latifi et al., 1996,
Pesci et al., 1997, Medina et al., 2003). P. aeruginosa tolerance to tobramycin, hydrogen peroxide
and polymorphonuclear leukocytes is quorum sensing dependent (Bjarnsholt et al., 2005a), and
when quorum sensing is inhibited in P. aeruginosa, tobramycin are more efficient (Christensen et
al., 2012, Brackman et al., 2011).

63

AMPs could ideally also function as solely administered peptide antibiotics and especially against
biofilm infections because they are often not dependent on the cells being metabolic active, which
in contrast many conventional antibiotics are (Batoni et al., 2011). LL-37 has for example shown to
effectively kill MRSA at low concentrations (Larrick et al., 1995b, De Smet and Contreras, 2005,
Turner et al., 1998) and other AMPs have shown to have an synergistic effect with antibiotics on
MRSA (Mataraci and Dosler, 2012) .
The biggest challenge blocking AMPs for therapeutically use is the administration challenges
mentioned above. Basically, three ways to overcome these challenges have been suggested. Coating
of indwelling catheters with peptides, topical administration of the AMP or chemically modifying
the AMP to increase the half-life while maintaining antibacterial activity.
Peptides such as Pexiganan and MBI 226, which was thought to prevent catheter infections, (Sader
et al., 2004, Ge et al., 1999) were/are both drug candidates for topical deployment, but they have
both failed FDA approval due to various problems with the experimental setup.
Topical administration in the urinary tract would be rather inconvenient and impossible in many
complicated infections but prevention of attachment would be more evident.

Since most UTI

strains form a biofilm (Subramanian et al., 2012), it would be most strategic to target the adherence
to the indwelling catheter and the following biofilm formation. This also solves the problem with
per oral and intravenous absorption and turnover.
An approach to optimize peptide antibiotics so it can be administered orally or intravenously is to
alter the amino acids chemically so that proteases or gastric acid in the stomach cannot neutralize
the peptide. This could be altered by using D-amino acids to target the biofilm formation. This has
in fact already been shown in a study that used D-amino acids in LL-37 to maintain the
antibacterial potential while being stable to trypsin (Dean et al., 2011b).

64

6. Conclusion
The results show that both peptide LL-37 and 1037 kill clinical strains isolated from UTI. The low
antimicrobial activity of KR-12 for P. aeruginosa was not observed for K. pneumoniae.
The activity study of 1037 is not completely consistent the theoretical expectation of an optimal helix, but since the peptide is too small to form a -strands, I suggest that 1037 form in a helical
matter.
Both tobramycin and tetracycline showed to induce P. aeruginosa biofilm formation close to MIC.
LL-37 and 1037 were both active against P. aeruginosa biofilms at 2 g/ml inhibiting the biofilm
formation 21.9 % and 28.2 %, respectively. 1037 showed no effect on against pregrown P.
aeruginosa biofilms, and for LL-37, the effect was shown not to occur in the first two hours of
incubation.
LL-37 and 1037 showed only to inhibit K. pneumoniae biofilms after two hours of incubation and
for LL-37, the decrease of biofilm was indicated to be due to less bacterial growth. Interestingly,
1037 showed not to inhibit K. pneumoniae growth at 4x MIC.
The effects on E. coli strains were inconsistent. LL-37 was effective against E. coli CFT073 at 2
g/ml inhibiting the biofilm formation 29.3 %, and 1037 was effective against E. coli 536 at 2
g/ml inhibiting the biofilm formation 26.9 %.
The experiments indicate that LL-37 and 1037 are active against several strains isolated from
urinary tract.

65

7. Future perspectives
There are some challenges in developing AMPs as future antibiotics that need to be addressed. The
poor bioavailability when a peptide drug is administered orally limits the commercial applications.
In a chemically optimized peptide, the advancement away from the originally structure might
increase with more chemically adjustments and as discussed above, the structure is important for
maintaining the antimicrobial effect of AMPs.
In 2012, Pergamum, a biopharmaceutical company, began phase I/II trials of LL-37 for the
treatment of patients with hard-to-heal venous leg ulcers. This brings hope to the possibility that
LL-37 at some point can be exploited for its antimicrobial effects. Further analysis of LL-37 could
show that the KR-12 fragment of LL-37 could be more effective than LL-37 while lowering the
production cost.

Regarding the prevention of bacterial attachment in the urinary tract, screening more strains could
reveal if the isolates have developed a tolerance threshold to overcome the effect of LL-37. It could
in particular be interesting to measure the content of LL-37 fragments in the urine of infected
individuals.
Coating of urinary tract catheters could ideally prevent a number of UTIs but it would be too
optimistic to expect all catheter associated infections to be prevented. The cost of coating
indwelling catheters with peptides would also be high and complicated. If proven efficient, the
usage could be focused on individuals in the risk group.
Further studies could be applying circular dichroism to determine whether 1037 forms in an helical matter or as a random coil. The potential of 1037 as an antibiotic could be even better than
LL-37, but future studies should determine whether 1037 is able to inhibit multi drug resistant
strains.

66

8. References
AGERBERTH, B., CHARO, J., WERR, J., OLSSON, B., IDALI, F., LINDBOM, L., KIESSLING, R., JORNVALL, H.,
WIGZELL, H. & GUDMUNDSSON, G. H. 2000. The human antimicrobial and chemotactic peptides LL37 and alpha-defensins are expressed by specific lymphocyte and monocyte populations. Blood, 96,
3086-93.
AGERBERTH, B., GUNNE, H., ODEBERG, J., KOGNER, P., BOMAN, H. G. & GUDMUNDSSON, G. H. 1995. FALL39, a putative human peptide antibiotic, is cysteine-free and expressed in bone marrow and testis.
Proc Natl Acad Sci U S A, 92, 195-9.
ALLESEN-HOLM, M., BARKEN, K. B., YANG, L., KLAUSEN, M., WEBB, J. S., KJELLEBERG, S., MOLIN, S.,
GIVSKOV, M. & TOLKER-NIELSEN, T. 2006. A characterization of DNA release in Pseudomonas
aeruginosa cultures and biofilms. Mol Microbiol, 59, 1114-28.
AMER, L. S., BISHOP, B. M. & VAN HOEK, M. L. 2010. Antimicrobial and antibiofilm activity of cathelicidins
and short, synthetic peptides against Francisella. Biochem Biophys Res Commun, 396, 246-51.
ANDERSSON, E., SORENSEN, O. E., FROHM, B., BORREGAARD, N., EGESTEN, A. & MALM, J. 2002. Isolation of
human cationic antimicrobial protein-18 from seminal plasma and its association with prostasomes.
Hum Reprod, 17, 2529-34.
BABIC, F., VENTURI, V. & MARAVIC-VLAHOVICEK, G. 2010. Tobramycin at subinhibitory concentration
inhibits the RhlI/R quorum sensing system in a Pseudomonas aeruginosa environmental isolate.
BMC Infect Dis, 10, 148.
BAGELLA, L., SCOCCHI, M. & ZANETTI, M. 1995. cDNA sequences of three sheep myeloid cathelicidins. FEBS
Lett, 376, 225-8.
BAGHERI, M., BEYERMANN, M. & DATHE, M. 2009. Immobilization reduces the activity of surface-bound
cationic antimicrobial peptides with no influence upon the activity spectrum. Antimicrob Agents
Chemother, 53, 1132-41.
BALS, R., WANG, X., ZASLOFF, M. & WILSON, J. M. 1998. The peptide antibiotic LL-37/hCAP-18 is expressed
in epithelia of the human lung where it has broad antimicrobial activity at the airway surface. Proc
Natl Acad Sci U S A, 95, 9541-6.
BALS, R. & WILSON, J. M. 2003. Cathelicidins--a family of multifunctional antimicrobial peptides. Cell Mol
Life Sci, 60, 711-20.
BANERJEE, R., JOHNSTON, B., LOHSE, C., PORTER, S. B., CLABOTS, C. & JOHNSON, J. R. 2013. Escherichia coli
sequence type 131 is a dominant, antimicrobial-resistant clonal group associated with healthcare
and elderly hosts. Infect Control Hosp Epidemiol, 34, 361-9.
BARLOW, P. G., SVOBODA, P., MACKELLAR, A., NASH, A. A., YORK, I. A., POHL, J., DAVIDSON, D. J. & DONIS,
R. O. 2011. Antiviral activity and increased host defense against influenza infection elicited by the
human cathelicidin LL-37. PLoS One, 6, e25333.
BASSLER, B. L. 2002. Small talk. Cell-to-cell communication in bacteria. Cell, 109, 421-4.
BATONI, G., MAISETTA, G., BRANCATISANO, F. L., ESIN, S. & CAMPA, M. 2011. Use of antimicrobial peptides
against microbial biofilms: advantages and limits. Curr Med Chem, 18, 256-79.
BENINCASA, M., MATTIUZZO, M., HERASIMENKA, Y., CESCUTTI, P., RIZZO, R. & GENNARO, R. 2009. Activity
of antimicrobial peptides in the presence of polysaccharides produced by pulmonary pathogens. J
Pept Sci, 15, 595-600.
BEVINS, C. L. & ZASLOFF, M. 1990. Peptides from frog skin. Annu Rev Biochem, 59, 395-414.
BJARNSHOLT, T., JENSEN, P. O., BURMOLLE, M., HENTZER, M., HAAGENSEN, J. A., HOUGEN, H. P., CALUM,
H., MADSEN, K. G., MOSER, C., MOLIN, S., HOIBY, N. & GIVSKOV, M. 2005a. Pseudomonas
aeruginosa tolerance to tobramycin, hydrogen peroxide and polymorphonuclear leukocytes is
quorum-sensing dependent. Microbiology, 151, 373-83.
BJARNSHOLT, T., JENSEN, P. O., RASMUSSEN, T. B., CHRISTOPHERSEN, L., CALUM, H., HENTZER, M.,
HOUGEN, H. P., RYGAARD, J., MOSER, C., EBERL, L., HOIBY, N. & GIVSKOV, M. 2005b. Garlic blocks

67

quorum sensing and promotes rapid clearing of pulmonary Pseudomonas aeruginosa infections.
Microbiology, 151, 3873-80.
BOWDISH, D. M., DAVIDSON, D. J. & HANCOCK, R. E. 2005. A re-evaluation of the role of host defence
peptides in mammalian immunity. Curr Protein Pept Sci, 6, 35-51.
BRACKMAN, G., COS, P., MAES, L., NELIS, H. J. & COENYE, T. 2011. Quorum sensing inhibitors increase the
susceptibility of bacterial biofilms to antibiotics in vitro and in vivo. Antimicrob Agents Chemother,
55, 2655-61.
BRAFF, M. H., HAWKINS, M. A., DI NARDO, A., LOPEZ-GARCIA, B., HOWELL, M. D., WONG, C., LIN, K., STREIB,
J. E., DORSCHNER, R., LEUNG, D. Y. & GALLO, R. L. 2005. Structure-function relationships among
human cathelicidin peptides: dissociation of antimicrobial properties from host immunostimulatory
activities. J Immunol, 174, 4271-8.
BROGDEN, K. A. 2005. Antimicrobial peptides: pore formers or metabolic inhibitors in bacteria? Nat Rev
Microbiol, 3, 238-50.
BROTZ-OESTERHELT, H. & BRUNNER, N. A. 2008. How many modes of action should an antibiotic have?
Curr Opin Pharmacol, 8, 564-73.
BUCHAU, A. S., MORIZANE, S., TROWBRIDGE, J., SCHAUBER, J., KOTOL, P., BUI, J. D. & GALLO, R. L. 2010.
The host defense peptide cathelicidin is required for NK cell-mediated suppression of tumor
growth. J Immunol, 184, 369-78.
BUCKI, R., NAMIOT, D. B., NAMIOT, Z., SAVAGE, P. B. & JANMEY, P. A. 2008. Salivary mucins inhibit
antibacterial activity of the cathelicidin-derived LL-37 peptide but not the cationic steroid CSA-13. J
Antimicrob Chemother, 62, 329-35.
CAIAZZA, N. C., MERRITT, J. H., BROTHERS, K. M. & O'TOOLE, G. A. 2007. Inverse regulation of biofilm
formation and swarming motility by Pseudomonas aeruginosa PA14. J Bacteriol, 189, 3603-12.
CAIAZZA, N. C. & O'TOOLE, G. A. 2004. SadB is required for the transition from reversible to irreversible
attachment during biofilm formation by Pseudomonas aeruginosa PA14. J Bacteriol, 186, 4476-85.
CHEN, S. L., HUNG, C. S., PINKNER, J. S., WALKER, J. N., CUSUMANO, C. K., LI, Z., BOUCKAERT, J., GORDON, J.
I. & HULTGREN, S. J. 2009. Positive selection identifies an in vivo role for FimH during urinary tract
infection in addition to mannose binding. Proc Natl Acad Sci U S A, 106, 22439-44.
CHRISTENSEN, L. D., VAN GENNIP, M., JAKOBSEN, T. H., ALHEDE, M., HOUGEN, H. P., HOIBY, N.,
BJARNSHOLT, T. & GIVSKOV, M. 2012. Synergistic antibacterial efficacy of early combination
treatment with tobramycin and quorum-sensing inhibitors against Pseudomonas aeruginosa in an
intraperitoneal foreign-body infection mouse model. J Antimicrob Chemother, 67, 1198-206.
CHROMEK, M., SLAMOVA, Z., BERGMAN, P., KOVACS, L., PODRACKA, L., EHREN, I., HOKFELT, T.,
GUDMUNDSSON, G. H., GALLO, R. L., AGERBERTH, B. & BRAUNER, A. 2006. The antimicrobial
peptide cathelicidin protects the urinary tract against invasive bacterial infection. Nat Med, 12,
636-41.
CLATWORTHY, A. E., PIERSON, E. & HUNG, D. T. 2007. Targeting virulence: a new paradigm for antimicrobial
therapy. Nat Chem Biol, 3, 541-8.
COLVIN, K. M., GORDON, V. D., MURAKAMI, K., BORLEE, B. R., WOZNIAK, D. J., WONG, G. C. & PARSEK, M.
R. 2011. The pel polysaccharide can serve a structural and protective role in the biofilm matrix of
Pseudomonas aeruginosa. PLoS Pathog, 7, e1001264.
COLVIN, K. M., IRIE, Y., TART, C. S., URBANO, R., WHITNEY, J. C., RYDER, C., HOWELL, P. L., WOZNIAK, D. J. &
PARSEK, M. R. 2012. The Pel and Psl polysaccharides provide Pseudomonas aeruginosa structural
redundancy within the biofilm matrix. Environ Microbiol, 14, 1913-28.
COSGROVE, S. E., SAKOULAS, G., PERENCEVICH, E. N., SCHWABER, M. J., KARCHMER, A. W. & CARMELI, Y.
2003. Comparison of mortality associated with methicillin-resistant and methicillin-susceptible
Staphylococcus aureus bacteremia: a meta-analysis. Clin Infect Dis, 36, 53-9.
COSTA, F., CARVALHO, I. F., MONTELARO, R. C., GOMES, P. & MARTINS, M. C. 2011. Covalent
immobilization of antimicrobial peptides (AMPs) onto biomaterial surfaces. Acta Biomater, 7, 143140.

68

COSTERTON, J. W., GEESEY, G. G. & CHENG, K. J. 1978. How bacteria stick. Sci Am, 238, 86-95.
COSTERTON, J. W., STEWART, P. S. & GREENBERG, E. P. 1999. Bacterial biofilms: a common cause of
persistent infections. Science, 284, 1318-22.
COWLAND, J. B., JOHNSEN, A. H. & BORREGAARD, N. 1995. hCAP-18, a cathelin/pro-bactenecin-like protein
of human neutrophil specific granules. FEBS Lett, 368, 173-6.
DA SILVA, B. R., DE FREITAS, V. A., CARNEIRO, V. A., ARRUDA, F. V., LORENZON, E. N., DE AGUIAR, A. S.,
CILLI, E. M., CAVADA, B. S. & TEIXEIRA, E. H. 2013. Antimicrobial activity of the synthetic peptide
Lys-a1 against oral streptococci. Peptides.
DAROUICHE, R. O., RAAD, II, HEARD, S. O., THORNBY, J. I., WENKER, O. C., GABRIELLI, A., BERG, J.,
KHARDORI, N., HANNA, H., HACHEM, R., HARRIS, R. L. & MAYHALL, G. 1999. A comparison of two
antimicrobial-impregnated central venous catheters. Catheter Study Group. N Engl J Med, 340, 1-8.
DAVIDOPOULOU, S., DIZA, E., MENEXES, G. & KALFAS, S. 2012. Salivary concentration of the antimicrobial
peptide LL-37 in children. Arch Oral Biol, 57, 865-9.
DAVIES, D. 2003. Understanding biofilm resistance to antibacterial agents. Nat Rev Drug Discov, 2, 114-22.
DAVIES, D. G., PARSEK, M. R., PEARSON, J. P., IGLEWSKI, B. H., COSTERTON, J. W. & GREENBERG, E. P. 1998.
The involvement of cell-to-cell signals in the development of a bacterial biofilm. Science, 280, 2958.
DE KIEVIT, T. R. 2009. Quorum sensing in Pseudomonas aeruginosa biofilms. Environ Microbiol, 11, 279-88.
DE LA FUENTE-NUNEZ, C., KOROLIK, V., BAINS, M., NGUYEN, U., BREIDENSTEIN, E. B., HORSMAN, S.,
LEWENZA, S., BURROWS, L. & HANCOCK, R. E. 2012a. Inhibition of bacterial biofilm formation and
swarming motility by a small synthetic cationic peptide. Antimicrob Agents Chemother, 56, 2696704.
DE LA FUENTE-NUNEZ, C., MERTENS, J., SMIT, J. & HANCOCK, R. E. 2012b. The bacterial surface layer
provides protection against antimicrobial peptides. Appl Environ Microbiol, 78, 5452-6.
DE SMET, K. & CONTRERAS, R. 2005. Human antimicrobial peptides: defensins, cathelicidins and histatins.
Biotechnol Lett, 27, 1337-47.
DE, Y., CHEN, Q., SCHMIDT, A. P., ANDERSON, G. M., WANG, J. M., WOOTERS, J., OPPENHEIM, J. J. &
CHERTOV, O. 2000. LL-37, the neutrophil granule- and epithelial cell-derived cathelicidin, utilizes
formyl peptide receptor-like 1 (FPRL1) as a receptor to chemoattract human peripheral blood
neutrophils, monocytes, and T cells. J Exp Med, 192, 1069-74.
DEAN, S. N., BISHOP, B. M. & VAN HOEK, M. L. 2011a. Natural and synthetic cathelicidin peptides with antimicrobial and anti-biofilm activity against Staphylococcus aureus. BMC Microbiol, 11, 114.
DEAN, S. N., BISHOP, B. M. & VAN HOEK, M. L. 2011b. Susceptibility of Pseudomonas aeruginosa Biofilm to
Alpha-Helical Peptides: D-enantiomer of LL-37. Front Microbiol, 2, 128.
DEN HERTOG, A. L., VAN MARLE, J., VAN VEEN, H. A., VAN'T HOF, W., BOLSCHER, J. G., VEERMAN, E. C. &
NIEUW AMERONGEN, A. V. 2005. Candidacidal effects of two antimicrobial peptides: histatin 5
causes small membrane defects, but LL-37 causes massive disruption of the cell membrane.
Biochem J, 388, 689-95.
DI NARDO, A., VITIELLO, A. & GALLO, R. L. 2003. Cutting edge: mast cell antimicrobial activity is mediated by
expression of cathelicidin antimicrobial peptide. J Immunol, 170, 2274-8.
DIXON, B. M., BARKER, T., MCKINNON, T., CUOMO, J., FREI, B., BORREGAARD, N. & GOMBART, A. F. 2012.
Positive correlation between circulating cathelicidin antimicrobial peptide (hCAP18/LL-37) and 25hydroxyvitamin D levels in healthy adults. BMC Res Notes, 5, 575.
DONADIO, S., MAFFIOLI, S., MONCIARDINI, P., SOSIO, M. & JABES, D. 2010. Antibiotic discovery in the
twenty-first century: current trends and future perspectives. J Antibiot (Tokyo), 63, 423-30.
DURR, U. H., SUDHEENDRA, U. S. & RAMAMOORTHY, A. 2006. LL-37, the only human member of the
cathelicidin family of antimicrobial peptides. Biochim Biophys Acta, 1758, 1408-25.
EISENBERG, D., WEISS, R. M. & TERWILLIGER, T. C. 1982. The helical hydrophobic moment: a measure of
the amphiphilicity of a helix. Nature, 299, 371-4.

69

ELLIOTT, D., BURNS, J. L. & HOFFMAN, L. R. 2010. Exploratory study of the prevalence and clinical
significance of tobramycin-mediated biofilm induction in Pseudomonas aeruginosa isolates from
cystic fibrosis patients. Antimicrob Agents Chemother, 54, 3024-6.
FABIAN, T. K., HERMANN, P., BECK, A., FEJERDY, P. & FABIAN, G. 2012. Salivary defense proteins: their
network and role in innate and acquired oral immunity. Int J Mol Sci, 13, 4295-320.
FINER, G. & LANDAU, D. 2004. Pathogenesis of urinary tract infections with normal female anatomy. Lancet
Infect Dis, 4, 631-5.
FLEMMING, H. C. & WINGENDER, J. 2010. The biofilm matrix. Nat Rev Microbiol, 8, 623-33.
FOSCHIATTI, M., CESCUTTI, P., TOSSI, A. & RIZZO, R. 2009. Inhibition of cathelicidin activity by bacterial
exopolysaccharides. Mol Microbiol, 72, 1137-46.
FRIEDMAN, L. & KOLTER, R. 2004. Genes involved in matrix formation in Pseudomonas aeruginosa PA14
biofilms. Mol Microbiol, 51, 675-90.
FROHM, M., GUNNE, H., BERGMAN, A. C., AGERBERTH, B., BERGMAN, T., BOMAN, A., LIDEN, S., JORNVALL,
H. & BOMAN, H. G. 1996. Biochemical and antibacterial analysis of human wound and blister fluid.
Eur J Biochem, 237, 86-92.
FUQUA, C., PARSEK, M. R. & GREENBERG, E. P. 2001. Regulation of gene expression by cell-to-cell
communication: acyl-homoserine lactone quorum sensing. Annu Rev Genet, 35, 439-68.
GABRIEL, M., NAZMI, K., VEERMAN, E. C., NIEUW AMERONGEN, A. V. & ZENTNER, A. 2006. Preparation of
LL-37-grafted titanium surfaces with bactericidal activity. Bioconjug Chem, 17, 548-50.
GANZ, T. 2003. Defensins: antimicrobial peptides of innate immunity. Nat Rev Immunol, 3, 710-20.
GE, Y., MACDONALD, D. L., HOLROYD, K. J., THORNSBERRY, C., WEXLER, H. & ZASLOFF, M. 1999. In vitro
antibacterial properties of pexiganan, an analog of magainin. Antimicrob Agents Chemother, 43,
782-8.
GENNARO, R. & ZANETTI, M. 2000. Structural features and biological activities of the cathelicidin-derived
antimicrobial peptides. Biopolymers, 55, 31-49.
GHAFOOR, A., HAY, I. D. & REHM, B. H. 2011. Role of exopolysaccharides in Pseudomonas aeruginosa
biofilm formation and architecture. Appl Environ Microbiol, 77, 5238-46.
GILTNER, C. L., VAN SCHAIK, E. J., AUDETTE, G. F., KAO, D., HODGES, R. S., HASSETT, D. J. & IRVIN, R. T. 2006.
The Pseudomonas aeruginosa type IV pilin receptor binding domain functions as an adhesin for
both biotic and abiotic surfaces. Mol Microbiol, 59, 1083-96.
GLESSNER, A., SMITH, R. S., IGLEWSKI, B. H. & ROBINSON, J. B. 1999. Roles of Pseudomonas aeruginosa las
and rhl quorum-sensing systems in control of twitching motility. J Bacteriol, 181, 1623-9.
GORDON, Y. J., HUANG, L. C., ROMANOWSKI, E. G., YATES, K. A., PROSKE, R. J. & MCDERMOTT, A. M. 2005.
Human cathelicidin (LL-37), a multifunctional peptide, is expressed by ocular surface epithelia and
has potent antibacterial and antiviral activity. Curr Eye Res, 30, 385-94.
GORR, S. U. 2012. Antimicrobial peptides in periodontal innate defense. Front Oral Biol, 15, 84-98.
GUDMUNDSSON, G. H., AGERBERTH, B., ODEBERG, J., BERGMAN, T., OLSSON, B. & SALCEDO, R. 1996. The
human gene FALL39 and processing of the cathelin precursor to the antibacterial peptide LL-37 in
granulocytes. Eur J Biochem, 238, 325-32.
GUTSMANN, T., HAGGE, S. O., DAVID, A., ROES, S., BOHLING, A., HAMMER, M. U. & SEYDEL, U. 2005. Lipidmediated resistance of Gram-negative bacteria against various pore-forming antimicrobial
peptides. J Endotoxin Res, 11, 167-73.
HALL-STOODLEY, L. & STOODLEY, P. 2005. Biofilm formation and dispersal and the transmission of human
pathogens. Trends Microbiol, 13, 7-10.
HANCOCK, R. E. & LEHRER, R. 1998. Cationic peptides: a new source of antibiotics. Trends Biotechnol, 16,
82-8.
HANCOCK, R. E. & SAHL, H. G. 2006. Antimicrobial and host-defense peptides as new anti-infective
therapeutic strategies. Nat Biotechnol, 24, 1551-7.
HANCOCK, V., WITSO, I. L. & KLEMM, P. 2011. Biofilm formation as a function of adhesin, growth medium,
substratum and strain type. Int J Med Microbiol, 301, 570-6.

70

HASE, K., ECKMANN, L., LEOPARD, J. D., VARKI, N. & KAGNOFF, M. F. 2002. Cell differentiation is a key
determinant of cathelicidin LL-37/human cationic antimicrobial protein 18 expression by human
colon epithelium. Infect Immun, 70, 953-63.
HASE, K., MURAKAMI, M., IIMURA, M., COLE, S. P., HORIBE, Y., OHTAKE, T., OBONYO, M., GALLO, R. L.,
ECKMANN, L. & KAGNOFF, M. F. 2003. Expression of LL-37 by human gastric epithelial cells as a
potential host defense mechanism against Helicobacter pylori. Gastroenterology, 125, 1613-25.
HATT, J. K. & RATHER, P. N. 2008. Role of bacterial biofilms in urinary tract infections. Curr Top Microbiol
Immunol, 322, 163-92.
HEILBORN, J. D., NILSSON, M. F., JIMENEZ, C. I., SANDSTEDT, B., BORREGAARD, N., THAM, E., SORENSEN, O.
E., WEBER, G. & STAHLE, M. 2005. Antimicrobial protein hCAP18/LL-37 is highly expressed in breast
cancer and is a putative growth factor for epithelial cells. Int J Cancer, 114, 713-9.
HEILBORN, J. D., NILSSON, M. F., KRATZ, G., WEBER, G., SORENSEN, O., BORREGAARD, N. & STAHLEBACKDAHL, M. 2003. The cathelicidin anti-microbial peptide LL-37 is involved in re-epithelialization
of human skin wounds and is lacking in chronic ulcer epithelium. J Invest Dermatol, 120, 379-89.
HELL, E., GISKE, C. G., NELSON, A., ROMLING, U. & MARCHINI, G. 2010. Human cathelicidin peptide LL37
inhibits both attachment capability and biofilm formation of Staphylococcus epidermidis. Lett Appl
Microbiol, 50, 211-5.
HERASIMENKA, Y., BENINCASA, M., MATTIUZZO, M., CESCUTTI, P., GENNARO, R. & RIZZO, R. 2005.
Interaction of antimicrobial peptides with bacterial polysaccharides from lung pathogens. Peptides,
26, 1127-32.
HILPERT, K., ELLIOTT, M., JENSSEN, H., KINDRACHUK, J., FJELL, C. D., KORNER, J., WINKLER, D. F., WEAVER, L.
L., HENKLEIN, P., ULRICH, A. S., CHIANG, S. H., FARMER, S. W., PANTE, N., VOLKMER, R. &
HANCOCK, R. E. 2009. Screening and characterization of surface-tethered cationic peptides for
antimicrobial activity. Chem Biol, 16, 58-69.
HIRAMATSU, K., CUI, L., KURODA, M. & ITO, T. 2001. The emergence and evolution of methicillin-resistant
Staphylococcus aureus. Trends Microbiol, 9, 486-93.
HIRAMATSU, K., HANAKI, H., INO, T., YABUTA, K., OGURI, T. & TENOVER, F. C. 1997. Methicillin-resistant
Staphylococcus aureus clinical strain with reduced vancomycin susceptibility. J Antimicrob
Chemother, 40, 135-6.
HOCHBAUM, A. I., KOLODKIN-GAL, I., FOULSTON, L., KOLTER, R., AIZENBERG, J. & LOSICK, R. 2011.
Inhibitory effects of D-amino acids on Staphylococcus aureus biofilm development. J Bacteriol, 193,
5616-22.
HOCHHUT, B., WILDE, C., BALLING, G., MIDDENDORF, B., DOBRINDT, U., BRZUSZKIEWICZ, E., GOTTSCHALK,
G., CARNIEL, E. & HACKER, J. 2006. Role of pathogenicity island-associated integrases in the
genome plasticity of uropathogenic Escherichia coli strain 536. Mol Microbiol, 61, 584-95.
HOIBY, N., BJARNSHOLT, T., GIVSKOV, M., MOLIN, S. & CIOFU, O. 2010. Antibiotic resistance of bacterial
biofilms. Int J Antimicrob Agents, 35, 322-32.
HOIBY, N., CIOFU, O., JOHANSEN, H. K., SONG, Z. J., MOSER, C., JENSEN, P. O., MOLIN, S., GIVSKOV, M.,
TOLKER-NIELSEN, T. & BJARNSHOLT, T. 2011. The clinical impact of bacterial biofilms. Int J Oral Sci,
3, 55-65.
HOOK, A. L., CHANG, C. Y., YANG, J., LUCKETT, J., COCKAYNE, A., ATKINSON, S., MEI, Y., BAYSTON, R.,
IRVINE, D. J., LANGER, R., ANDERSON, D. G., WILLIAMS, P., DAVIES, M. C. & ALEXANDER, M. R. 2012.
Combinatorial discovery of polymers resistant to bacterial attachment. Nat Biotechnol, 30, 868-75.
HOU, S., LIU, Z., YOUNG, A. W., MARK, S. L., KALLENBACH, N. R. & REN, D. 2010. Effects of Trp- and Argcontaining antimicrobial-peptide structure on inhibition of Escherichia coli planktonic growth and
biofilm formation. Appl Environ Microbiol, 76, 1967-74.
HUANG, L. C., PETKOVA, T. D., REINS, R. Y., PROSKE, R. J. & MCDERMOTT, A. M. 2006. Multifunctional roles
of human cathelicidin (LL-37) at the ocular surface. Invest Ophthalmol Vis Sci, 47, 2369-80.
IRIE, Y. & PARSEK, M. R. 2008. Quorum sensing and microbial biofilms. Curr Top Microbiol Immunol, 322,
67-84.

71

JACKSON, K. D., STARKEY, M., KREMER, S., PARSEK, M. R. & WOZNIAK, D. J. 2004. Identification of psl, a
locus encoding a potential exopolysaccharide that is essential for Pseudomonas aeruginosa PAO1
biofilm formation. J Bacteriol, 186, 4466-75.
JACOBSEN, A. S. & JENSSEN, H. 2012. Human cathelicidin LL-37 prevents bacterial biofilm formation. Future
Med Chem, 4, 1587-99.
JAIN, P., PARADA, J. P., DAVID, A. & SMITH, L. G. 1995. Overuse of the indwelling urinary tract catheter in
hospitalized medical patients. Arch Intern Med, 155, 1425-9.
JARAMILLO, D. E., ARRIOLA, A., SAFAVI, K. & CHAVEZ DE PAZ, L. E. 2012. Decreased bacterial adherence and
biofilm growth on surfaces coated with a solution of benzalkonium chloride. J Endod, 38, 821-5.
JENSSEN, H., HAMILL, P. & HANCOCK, R. E. 2006. Peptide antimicrobial agents. Clin Microbiol Rev, 19, 491511.
JIN, F., CONRAD, J. C., GIBIANSKY, M. L. & WONG, G. C. 2011. Bacteria use type-IV pili to slingshot on
surfaces. Proc Natl Acad Sci U S A, 108, 12617-22.
JOHANSSON, J., GUDMUNDSSON, G. H., ROTTENBERG, M. E., BERNDT, K. D. & AGERBERTH, B. 1998.
Conformation-dependent antibacterial activity of the naturally occurring human peptide LL-37. J
Biol Chem, 273, 3718-24.
JOHNSON, J. R., KUSKOWSKI, M. A. & WILT, T. J. 2006. Systematic review: antimicrobial urinary catheters to
prevent catheter-associated urinary tract infection in hospitalized patients. Ann Intern Med, 144,
116-26.
JOHNSON, J. R., NICOLAS-CHANOINE, M. H., DEBROY, C., CASTANHEIRA, M., ROBICSEK, A., HANSEN, G.,
WEISSMAN, S., URBAN, C., PLATELL, J., TROTT, D., ZHANEL, G., CLABOTS, C., JOHNSTON, B. D. &
KUSKOWSKI, M. A. 2012. Comparison of Escherichia coli ST131 pulsotypes, by epidemiologic traits,
1967-2009. Emerg Infect Dis, 18, 598-607.
JOHNSON, J. R., ROBERTS, P. L., OLSEN, R. J., MOYER, K. A. & STAMM, W. E. 1990. Prevention of catheterassociated urinary tract infection with a silver oxide-coated urinary catheter: clinical and
microbiologic correlates. J Infect Dis, 162, 1145-50.
JORGE, P., LOURENCO, A. & PEREIRA, M. O. 2012. New trends in peptide-based anti-biofilm strategies: a
review of recent achievements and bioinformatic approaches. Biofouling, 28, 1033-61.
KAI-LARSEN, Y., LUTHJE, P., CHROMEK, M., PETERS, V., WANG, X., HOLM, A., KADAS, L., HEDLUND, K. O.,
JOHANSSON, J., CHAPMAN, M. R., JACOBSON, S. H., ROMLING, U., AGERBERTH, B. & BRAUNER, A.
2010. Uropathogenic Escherichia coli modulates immune responses and its curli fimbriae interact
with the antimicrobial peptide LL-37. PLoS Pathog, 6, e1001010.
KAPLAN, J. B., IZANO, E. A., GOPAL, P., KARWACKI, M. T., KIM, S., BOSE, J. L., BAYLES, K. W. & HORSWILL, A.
R. 2012. Low levels of beta-lactam antibiotics induce extracellular DNA release and biofilm
formation in Staphylococcus aureus. MBio, 3, e00198-12.
KAPLAN, J. B., RAGUNATH, C., RAMASUBBU, N. & FINE, D. H. 2003. Detachment of Actinobacillus
actinomycetemcomitans biofilm cells by an endogenous beta-hexosaminidase activity. J Bacteriol,
185, 4693-8.
KAPOOR, R., WADMAN, M. W., DOHM, M. T., CZYZEWSKI, A. M., SPORMANN, A. M. & BARRON, A. E. 2011.
Antimicrobial peptoids are effective against Pseudomonas aeruginosa biofilms. Antimicrob Agents
Chemother, 55, 3054-7.
KARCHMER, T. B., GIANNETTA, E. T., MUTO, C. A., STRAIN, B. A. & FARR, B. M. 2000. A randomized
crossover study of silver-coated urinary catheters in hospitalized patients. Arch Intern Med, 160,
3294-8.
KAZEMZADEH-NARBAT, M., KINDRACHUK, J., DUAN, K., JENSSEN, H., HANCOCK, R. E. & WANG, R. 2010.
Antimicrobial peptides on calcium phosphate-coated titanium for the prevention of implantassociated infections. Biomaterials, 31, 9519-26.
KJELLEBERG, S. & MOLIN, S. 2002. Is there a role for quorum sensing signals in bacterial biofilms? Curr Opin
Microbiol, 5, 254-8.

72

KLAUSEN, M., HEYDORN, A., RAGAS, P., LAMBERTSEN, L., AAES-JORGENSEN, A., MOLIN, S. & TOLKERNIELSEN, T. 2003. Biofilm formation by Pseudomonas aeruginosa wild type, flagella and type IV pili
mutants. Mol Microbiol, 48, 1511-24.
KLEMM, P. & SCHEMBRI, M. A. 2000. Bacterial adhesins: function and structure. Int J Med Microbiol, 290,
27-35.
KLEMM, P., VEJBORG, R. M. & HANCOCK, V. 2010. Prevention of bacterial adhesion. Appl Microbiol
Biotechnol, 88, 451-9.
KLEVENS, R. M., MORRISON, M. A., NADLE, J., PETIT, S., GERSHMAN, K., RAY, S., HARRISON, L. H., LYNFIELD,
R., DUMYATI, G., TOWNES, J. M., CRAIG, A. S., ZELL, E. R., FOSHEIM, G. E., MCDOUGAL, L. K., CAREY,
R. B. & FRIDKIN, S. K. 2007. Invasive methicillin-resistant Staphylococcus aureus infections in the
United States. JAMA, 298, 1763-71.
KLOCKGETHER, J., MUNDER, A., NEUGEBAUER, J., DAVENPORT, C. F., STANKE, F., LARBIG, K. D., HEEB, S.,
SCHOCK, U., POHL, T. M., WIEHLMANN, L. & TUMMLER, B. 2010. Genome diversity of
Pseudomonas aeruginosa PAO1 laboratory strains. J Bacteriol, 192, 1113-21.
KOCZULLA, R., VON DEGENFELD, G., KUPATT, C., KROTZ, F., ZAHLER, S., GLOE, T., ISSBRUCKER, K.,
UNTERBERGER, P., ZAIOU, M., LEBHERZ, C., KARL, A., RAAKE, P., PFOSSER, A., BOEKSTEGERS, P.,
WELSCH, U., HIEMSTRA, P. S., VOGELMEIER, C., GALLO, R. L., CLAUSS, M. & BALS, R. 2003. An
angiogenic role for the human peptide antibiotic LL-37/hCAP-18. J Clin Invest, 111, 1665-72.
KOSCIUCZUK, E. M., LISOWSKI, P., JARCZAK, J., STRZALKOWSKA, N., JOZWIK, A., HORBANCZUK, J.,
KRZYZEWSKI, J., ZWIERZCHOWSKI, L. & BAGNICKA, E. 2012. Cathelicidins: family of antimicrobial
peptides. A review. Mol Biol Rep, 39, 10957-70.
KOWALCZUK, D., GINALSKA, G., PIERSIAK, T. & MIAZGA-KARSKA, M. 2012. Prevention of biofilm formation
on urinary catheters: comparison of the sparfloxacin-treated long-term antimicrobial catheters
with silver-coated ones. J Biomed Mater Res B Appl Biomater, 100B, 1874-82.
KUCHMA, S. L., BROTHERS, K. M., MERRITT, J. H., LIBERATI, N. T., AUSUBEL, F. M. & O'TOOLE, G. A. 2007.
BifA, a cyclic-Di-GMP phosphodiesterase, inversely regulates biofilm formation and swarming
motility by Pseudomonas aeruginosa PA14. J Bacteriol, 189, 8165-78.
LANGE, R. P., LOCHER, H. H., WYSS, P. C. & THEN, R. L. 2007. The targets of currently used antibacterial
agents: lessons for drug discovery. Curr Pharm Des, 13, 3140-54.
LARRICK, J. W., HIRATA, M., BALINT, R. F., LEE, J., ZHONG, J. & WRIGHT, S. C. 1995a. Human CAP18: a novel
antimicrobial lipopolysaccharide-binding protein. Infect Immun, 63, 1291-7.
LARRICK, J. W., HIRATA, M., ZHONG, J. & WRIGHT, S. C. 1995b. Anti-microbial activity of human CAP18
peptides. Immunotechnology, 1, 65-72.
LARRICK, J. W., LEE, J., MA, S., LI, X., FRANCKE, U., WRIGHT, S. C. & BALINT, R. F. 1996. Structural, functional
analysis and localization of the human CAP18 gene. FEBS Lett, 398, 74-80.
LATIFI, A., FOGLINO, M., TANAKA, K., WILLIAMS, P. & LAZDUNSKI, A. 1996. A hierarchical quorum-sensing
cascade in Pseudomonas aeruginosa links the transcriptional activators LasR and RhIR (VsmR) to
expression of the stationary-phase sigma factor RpoS. Mol Microbiol, 21, 1137-46.
LEWIS, K. 2010. Persister cells. Annu Rev Microbiol, 64, 357-72.
LIANG, H., LI, L., DONG, Z., SURETTE, M. G. & DUAN, K. 2008. The YebC family protein PA0964 negatively
regulates the Pseudomonas aeruginosa quinolone signal system and pyocyanin production. J
Bacteriol, 190, 6217-27.
LIEDBERG, H. & LUNDEBERG, T. 1989. Silver coating of urinary catheters prevents adherence and growth of
Pseudomonas aeruginosa. Urol Res, 17, 357-8.
LINARES, J. F., GUSTAFSSON, I., BAQUERO, F. & MARTINEZ, J. L. 2006. Antibiotics as intermicrobial signaling
agents instead of weapons. Proc Natl Acad Sci U S A, 103, 19484-9.
LIU, J., LING, J. Q., ZHANG, K., HUO, L. J. & NING, Y. 2012. Effect of sodium fluoride, ampicillin, and
chlorhexidine on Streptococcus mutans biofilm detachment. Antimicrob Agents Chemother, 56,
4532-5.

73

LOPEZ-GARCIA, B., LEE, P. H., YAMASAKI, K. & GALLO, R. L. 2005. Anti-fungal activity of cathelicidins and
their potential role in Candida albicans skin infection. J Invest Dermatol, 125, 108-15.
LOWY, F. D. 2003. Antimicrobial resistance: the example of Staphylococcus aureus. J Clin Invest, 111, 126573.
LUPPENS, S. B., REIJ, M. W., VAN DER HEIJDEN, R. W., ROMBOUTS, F. M. & ABEE, T. 2002. Development of a
standard test to assess the resistance of Staphylococcus aureus biofilm cells to disinfectants. Appl
Environ Microbiol, 68, 4194-200.
MA, L., CONOVER, M., LU, H., PARSEK, M. R., BAYLES, K. & WOZNIAK, D. J. 2009. Assembly and development
of the Pseudomonas aeruginosa biofilm matrix. PLoS Pathog, 5, e1000354.
MA, L., JACKSON, K. D., LANDRY, R. M., PARSEK, M. R. & WOZNIAK, D. J. 2006. Analysis of Pseudomonas
aeruginosa conditional psl variants reveals roles for the psl polysaccharide in adhesion and
maintaining biofilm structure postattachment. J Bacteriol, 188, 8213-21.
MALM, J., SORENSEN, O., PERSSON, T., FROHM-NILSSON, M., JOHANSSON, B., BJARTELL, A., LILJA, H.,
STAHLE-BACKDAHL, M., BORREGAARD, N. & EGESTEN, A. 2000. The human cationic antimicrobial
protein (hCAP-18) is expressed in the epithelium of human epididymis, is present in seminal plasma
at high concentrations, and is attached to spermatozoa. Infect Immun, 68, 4297-302.
MATARACI, E. & DOSLER, S. 2012. In vitro activities of antibiotics and antimicrobial cationic peptides alone
and in combination against methicillin-resistant Staphylococcus aureus biofilms. Antimicrob Agents
Chemother, 56, 6366-71.
MATTHEWS, S. J. & LANCASTER, J. W. 2011. Urinary tract infections in the elderly population. Am J Geriatr
Pharmacother, 9, 286-309.
MATTICK, J. S. 2002. Type IV pili and twitching motility. Annu Rev Microbiol, 56, 289-314.
MEDINA, G., JUAREZ, K., DIAZ, R. & SOBERON-CHAVEZ, G. 2003. Transcriptional regulation of Pseudomonas
aeruginosa rhlR, encoding a quorum-sensing regulatory protein. Microbiology, 149, 3073-81.
MERRITT, J. H., BROTHERS, K. M., KUCHMA, S. L. & O'TOOLE, G. A. 2007. SadC reciprocally influences biofilm
formation and swarming motility via modulation of exopolysaccharide production and flagellar
function. J Bacteriol, 189, 8154-64.
MERRITT, J. H., KADOURI, D. E. & O'TOOLE, G. A. 2005. Growing and analyzing static biofilms. Curr Protoc
Microbiol, Chapter 1, Unit 1B 1.
MOHANTY, S., MISHRA, S., JENA, P., JACOB, B., SARKAR, B. & SONAWANE, A. 2012. An investigation on the
antibacterial, cytotoxic, and antibiofilm efficacy of starch-stabilized silver nanoparticles.
Nanomedicine, 8, 916-24.
MORAN, G. J., KRISHNADASAN, A., GORWITZ, R. J., FOSHEIM, G. E., MCDOUGAL, L. K., CAREY, R. B. &
TALAN, D. A. 2006. Methicillin-resistant S. aureus infections among patients in the emergency
department. N Engl J Med, 355, 666-74.
MORAR, M. & WRIGHT, G. D. 2010. The genomic enzymology of antibiotic resistance. Annu Rev Genet, 44,
25-51.
MORRISON, K. L. & WEISS, G. A. 2001. Combinatorial alanine-scanning. Curr Opin Chem Biol, 5, 302-7.
MUNOZ, V. & SERRANO, L. 1994. Intrinsic secondary structure propensities of the amino acids, using
statistical phi-psi matrices: comparison with experimental scales. Proteins, 20, 301-11.
MURAKAMI, M., LOPEZ-GARCIA, B., BRAFF, M., DORSCHNER, R. A. & GALLO, R. L. 2004. Postsecretory
processing generates multiple cathelicidins for enhanced topical antimicrobial defense. J Immunol,
172, 3070-7.
MURAKAMI, M., OHTAKE, T., DORSCHNER, R. A. & GALLO, R. L. 2002a. Cathelicidin antimicrobial peptides
are expressed in salivary glands and saliva. J Dent Res, 81, 845-50.
MURAKAMI, M., OHTAKE, T., DORSCHNER, R. A., SCHITTEK, B., GARBE, C. & GALLO, R. L. 2002b. Cathelicidin
anti-microbial peptide expression in sweat, an innate defense system for the skin. J Invest
Dermatol, 119, 1090-5.

74

MURRAY, T. S., LEDIZET, M. & KAZMIERCZAK, B. I. 2010. Swarming motility, secretion of type 3 effectors
and biofilm formation phenotypes exhibited within a large cohort of Pseudomonas aeruginosa
clinical isolates. J Med Microbiol, 59, 511-20.
MYGIND, P. H., FISCHER, R. L., SCHNORR, K. M., HANSEN, M. T., SONKSEN, C. P., LUDVIGSEN, S., RAVENTOS,
D., BUSKOV, S., CHRISTENSEN, B., DE MARIA, L., TABOUREAU, O., YAVER, D., ELVIG-JORGENSEN, S.
G., SORENSEN, M. V., CHRISTENSEN, B. E., KJAERULFF, S., FRIMODT-MOLLER, N., LEHRER, R. I.,
ZASLOFF, M. & KRISTENSEN, H. H. 2005. Plectasin is a peptide antibiotic with therapeutic potential
from a saprophytic fungus. Nature, 437, 975-80.
NAGANT, C., PITTS, B., NAZMI, K., VANDENBRANDEN, M., BOLSCHER, J. G., STEWART, P. S. & DEHAYE, J. P.
2012. Identification of peptides derived from the human antimicrobial peptide LL-37 active against
biofilms formed by Pseudomonas aeruginosa using a library of truncated fragments. Antimicrob
Agents Chemother, 56, 5698-708.
NAGAOKA, I., HIROTA, S., NIYONSABA, F., HIRATA, M., ADACHI, Y., TAMURA, H. & HEUMANN, D. 2001.
Cathelicidin family of antibacterial peptides CAP18 and CAP11 inhibit the expression of TNF-alpha
by blocking the binding of LPS to CD14(+) cells. J Immunol, 167, 3329-38.
NAIMI, T. S., LEDELL, K. H., COMO-SABETTI, K., BORCHARDT, S. M., BOXRUD, D. J., ETIENNE, J., JOHNSON, S.
K., VANDENESCH, F., FRIDKIN, S., O'BOYLE, C., DANILA, R. N. & LYNFIELD, R. 2003. Comparison of
community- and health care-associated methicillin-resistant Staphylococcus aureus infection.
JAMA, 290, 2976-84.
NG, W. L. & BASSLER, B. L. 2009. Bacterial quorum-sensing network architectures. Annu Rev Genet, 43, 197222.
NICOLAS-CHANOINE, M. H., BLANCO, J., LEFLON-GUIBOUT, V., DEMARTY, R., ALONSO, M. P., CANICA, M.
M., PARK, Y. J., LAVIGNE, J. P., PITOUT, J. & JOHNSON, J. R. 2008. Intercontinental emergence of
Escherichia coli clone O25:H4-ST131 producing CTX-M-15. J Antimicrob Chemother, 61, 273-81.
NICOLLE, L. E. 2005. Catheter-related urinary tract infection. Drugs Aging, 22, 627-39.
NIELSEN, S. L., FRIMODT-MOLLER, N., KRAGELUND, B. B. & HANSEN, P. R. 2007. Structure--activity study of
the antibacterial peptide fallaxin. Protein Sci, 16, 1969-76.
NIYONSABA, F., IWABUCHI, K., SOMEYA, A., HIRATA, M., MATSUDA, H., OGAWA, H. & NAGAOKA, I. 2002. A
cathelicidin family of human antibacterial peptide LL-37 induces mast cell chemotaxis. Immunology,
106, 20-6.
NIYONSABA, F., SOMEYA, A., HIRATA, M., OGAWA, H. & NAGAOKA, I. 2001. Evaluation of the effects of
peptide antibiotics human beta-defensins-1/-2 and LL-37 on histamine release and prostaglandin
D(2) production from mast cells. Eur J Immunol, 31, 1066-75.
O'TOOLE, G. A. & KOLTER, R. 1998. Flagellar and twitching motility are necessary for Pseudomonas
aeruginosa biofilm development. Mol Microbiol, 30, 295-304.
O'TOOLE, G. A., PRATT, L. A., WATNICK, P. I., NEWMAN, D. K., WEAVER, V. B. & KOLTER, R. 1999. Genetic
approaches to study of biofilms. Methods Enzymol, 310, 91-109.
OLSON, M. E., CERI, H., MORCK, D. W., BURET, A. G. & READ, R. R. 2002. Biofilm bacteria: formation and
comparative susceptibility to antibiotics. Can J Vet Res, 66, 86-92.
OPPENHEIM, J. J. & YANG, D. 2005. Alarmins: chemotactic activators of immune responses. Curr Opin
Immunol, 17, 359-65.
OREN, Z., LERMAN, J. C., GUDMUNDSSON, G. H., AGERBERTH, B. & SHAI, Y. 1999. Structure and
organization of the human antimicrobial peptide LL-37 in phospholipid membranes: relevance to
the molecular basis for its non-cell-selective activity. Biochem J, 341 ( Pt 3), 501-13.
OVERHAGE, J., CAMPISANO, A., BAINS, M., TORFS, E. C., REHM, B. H. & HANCOCK, R. E. 2008. Human host
defense peptide LL-37 prevents bacterial biofilm formation. Infect Immun, 76, 4176-82.
PEARSON, J. P., GRAY, K. M., PASSADOR, L., TUCKER, K. D., EBERHARD, A., IGLEWSKI, B. H. & GREENBERG, E.
P. 1994. Structure of the autoinducer required for expression of Pseudomonas aeruginosa virulence
genes. Proc Natl Acad Sci U S A, 91, 197-201.

75

PESCI, E. C., PEARSON, J. P., SEED, P. C. & IGLEWSKI, B. H. 1997. Regulation of las and rhl quorum sensing in
Pseudomonas aeruginosa. J Bacteriol, 179, 3127-32.
PETROVA, O. E. & SAUER, K. 2012. Sticky situations: key components that control bacterial surface
attachment. J Bacteriol, 194, 2413-25.
PICIOREANU, C., KREFT, J. U., KLAUSEN, M., HAAGENSEN, J. A., TOLKER-NIELSEN, T. & MOLIN, S. 2007.
Microbial motility involvement in biofilm structure formation--a 3D modelling study. Water Sci
Technol, 55, 337-43.
POMPILIO, A., SCOCCHI, M., POMPONIO, S., GUIDA, F., DI PRIMIO, A., FISCARELLI, E., GENNARO, R. & DI
BONAVENTURA, G. 2011. Antibacterial and anti-biofilm effects of cathelicidin peptides against
pathogens isolated from cystic fibrosis patients. Peptides, 32, 1807-14.
POWERS, J. P. & HANCOCK, R. E. 2003. The relationship between peptide structure and antibacterial
activity. Peptides, 24, 1681-91.
RASMUSSEN, T. B., BJARNSHOLT, T., SKINDERSOE, M. E., HENTZER, M., KRISTOFFERSEN, P., KOTE, M.,
NIELSEN, J., EBERL, L. & GIVSKOV, M. 2005. Screening for quorum-sensing inhibitors (QSI) by use of
a novel genetic system, the QSI selector. J Bacteriol, 187, 1799-814.
REID, G., SHARMA, S., ADVIKOLANU, K., TIESZER, C., MARTIN, R. A. & BRUCE, A. W. 1994. Effects of
ciprofloxacin, norfloxacin, and ofloxacin on in vitro adhesion and survival of Pseudomonas
aeruginosa AK1 on urinary catheters. Antimicrob Agents Chemother, 38, 1490-5.
REN, S. X., CHENG, A. S., TO, K. F., TONG, J. H., LI, M. S., SHEN, J., WONG, C. C., ZHANG, L., CHAN, R. L.,
WANG, X. J., NG, S. S., CHIU, L. C., MARQUEZ, V. E., GALLO, R. L., CHAN, F. K., YU, J., SUNG, J. J., WU,
W. K. & CHO, C. H. 2012. Host immune defense peptide LL-37 activates caspase-independent
apoptosis and suppresses colon cancer. Cancer Res, 72, 6512-23.
RYDER, C., BYRD, M. & WOZNIAK, D. J. 2007. Role of polysaccharides in Pseudomonas aeruginosa biofilm
development. Curr Opin Microbiol, 10, 644-8.
RAAD, I., DAROUICHE, R., DUPUIS, J., ABI-SAID, D., GABRIELLI, A., HACHEM, R., WALL, M., HARRIS, R., JONES,
J., BUZAID, A., ROBERTSON, C., SHENAQ, S., CURLING, P., BURKE, T. & ERICSSON, C. 1997. Central
venous catheters coated with minocycline and rifampin for the prevention of catheter-related
colonization and bloodstream infections. A randomized, double-blind trial. The Texas Medical
Center Catheter Study Group. Ann Intern Med, 127, 267-74.
SADER, H. S., FEDLER, K. A., RENNIE, R. P., STEVENS, S. & JONES, R. N. 2004. Omiganan pentahydrochloride
(MBI 226), a topical 12-amino-acid cationic peptide: spectrum of antimicrobial activity and
measurements of bactericidal activity. Antimicrob Agents Chemother, 48, 3112-8.
SAINT, S. 2000. Clinical and economic consequences of nosocomial catheter-related bacteriuria. Am J Infect
Control, 28, 68-75.
SAKURAGI, Y. & KOLTER, R. 2007. Quorum-sensing regulation of the biofilm matrix genes (pel) of
Pseudomonas aeruginosa. J Bacteriol, 189, 5383-6.
SALO, J., SEVANDER, J. J., TAPIAINEN, T., IKAHEIMO, I., POKKA, T., KOSKELA, M. & UHARI, M. 2009. Biofilm
formation by Escherichia coli isolated from patients with urinary tract infections. Clin Nephrol, 71,
501-7.
SALVATORE, S., CATTONI, E., SIESTO, G., SERATI, M., SORICE, P. & TORELLA, M. 2011. Urinary tract
infections in women. Eur J Obstet Gynecol Reprod Biol, 156, 131-6.
SATO, H. & FEIX, J. B. 2006. Peptide-membrane interactions and mechanisms of membrane destruction by
amphipathic alpha-helical antimicrobial peptides. Biochim Biophys Acta, 1758, 1245-56.
SAUER, K., CAMPER, A. K., EHRLICH, G. D., COSTERTON, J. W. & DAVIES, D. G. 2002. Pseudomonas
aeruginosa displays multiple phenotypes during development as a biofilm. J Bacteriol, 184, 114054.
SAUER, K., CULLEN, M. C., RICKARD, A. H., ZEEF, L. A., DAVIES, D. G. & GILBERT, P. 2004. Characterization of
nutrient-induced dispersion in Pseudomonas aeruginosa PAO1 biofilm. J Bacteriol, 186, 7312-26.
SAUVAGE, E., KERFF, F., TERRAK, M., AYALA, J. A. & CHARLIER, P. 2008. The penicillin-binding proteins:
structure and role in peptidoglycan biosynthesis. FEMS Microbiol Rev, 32, 234-58.

76

SCHEFFERS, D. J. & PINHO, M. G. 2005. Bacterial cell wall synthesis: new insights from localization studies.
Microbiol Mol Biol Rev, 69, 585-607.
SCHITO, G. C. 2006. The importance of the development of antibiotic resistance in Staphylococcus aureus.
Clin Microbiol Infect, 12 Suppl 1, 3-8.
SCHITTEK, B., HIPFEL, R., SAUER, B., BAUER, J., KALBACHER, H., STEVANOVIC, S., SCHIRLE, M., SCHROEDER,
K., BLIN, N., MEIER, F., RASSNER, G. & GARBE, C. 2001. Dermcidin: a novel human antibiotic peptide
secreted by sweat glands. Nat Immunol, 2, 1133-7.
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 Microbiol, 10, 179.
SCHURR, M. J. 2013. Which Bacterial Biofilm Exopolysaccharide is Preferred, Psl or Alginate? J Bacteriol.
SHROUT, J. D., CHOPP, D. L., JUST, C. L., HENTZER, M., GIVSKOV, M. & PARSEK, M. R. 2006. The impact of
quorum sensing and swarming motility on Pseudomonas aeruginosa biofilm formation is
nutritionally conditional. Mol Microbiol, 62, 1264-77.
SIERADZKI, K. & TOMASZ, A. 1997. Inhibition of cell wall turnover and autolysis by vancomycin in a highly
vancomycin-resistant mutant of Staphylococcus aureus. J Bacteriol, 179, 2557-66.
SINGH, P. K., PARSEK, M. R., GREENBERG, E. P. & WELSH, M. J. 2002. A component of innate immunity
prevents bacterial biofilm development. Nature, 417, 552-5.
SMITH, A. W. 2005. Biofilms and antibiotic therapy: is there a role for combating bacterial resistance by the
use of novel drug delivery systems? Adv Drug Deliv Rev, 57, 1539-50.
SORENSEN, O. E., FOLLIN, P., JOHNSEN, A. H., CALAFAT, J., TJABRINGA, G. S., HIEMSTRA, P. S. &
BORREGAARD, N. 2001. Human cathelicidin, hCAP-18, is processed to the antimicrobial peptide LL37 by extracellular cleavage with proteinase 3. Blood, 97, 3951-9.
SORENSEN, O. E., GRAM, L., JOHNSEN, A. H., ANDERSSON, E., BANGSBOLL, S., TJABRINGA, G. S., HIEMSTRA,
P. S., MALM, J., EGESTEN, A. & BORREGAARD, N. 2003. Processing of seminal plasma hCAP-18 to
ALL-38 by gastricsin: a novel mechanism of generating antimicrobial peptides in vagina. J Biol Chem,
278, 28540-6.
SPENCER, J. D., SCHWADERER, A. L., WANG, H., BARTZ, J., KLINE, J., EICHLER, T., DESOUZA, K. R., SIMSLUCAS, S., BAKER, P. & HAINS, D. S. 2013. Ribonuclease 7, an antimicrobial peptide upregulated
during infection, contributes to microbial defense of the human urinary tract. Kidney Int, 83, 61525.
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 Immunol Med
Microbiol, 65, 350-9.
STEWART, P. S. 2002. Mechanisms of antibiotic resistance in bacterial biofilms. Int J Med Microbiol, 292,
107-13.
STEWART, P. S. & COSTERTON, J. W. 2001. Antibiotic resistance of bacteria in biofilms. Lancet, 358, 135-8.
STICKLER, D. J., MORRIS, N. S., MCLEAN, R. J. & FUQUA, C. 1998. Biofilms on indwelling urethral catheters
produce quorum-sensing signal molecules in situ and in vitro. Appl Environ Microbiol, 64, 3486-90.
STOODLEY, P., SAUER, K., DAVIES, D. G. & COSTERTON, J. W. 2002. Biofilms as complex differentiated
communities. Annu Rev Microbiol, 56, 187-209.
STOVER, C. K., PHAM, X. Q., ERWIN, A. L., MIZOGUCHI, S. D., WARRENER, P., HICKEY, M. J., BRINKMAN, F. S.,
HUFNAGLE, W. O., KOWALIK, D. J., LAGROU, M., GARBER, R. L., GOLTRY, L., TOLENTINO, E.,
WESTBROCK-WADMAN, S., YUAN, Y., BRODY, L. L., COULTER, S. N., FOLGER, K. R., KAS, A., LARBIG,
K., LIM, R., SMITH, K., SPENCER, D., WONG, G. K., WU, Z., PAULSEN, I. T., REIZER, J., SAIER, M. H.,
HANCOCK, R. E., LORY, S. & OLSON, M. V. 2000. Complete genome sequence of Pseudomonas
aeruginosa PAO1, an opportunistic pathogen. Nature, 406, 959-64.
STRUVE, C., BOJER, M. & KROGFELT, K. A. 2009. Identification of a conserved chromosomal region encoding
Klebsiella pneumoniae type 1 and type 3 fimbriae and assessment of the role of fimbriae in
pathogenicity. Infect Immun, 77, 5016-24.

77

SUBRAMANIAN, P., SHANMUGAM, N., SIVARAMAN, U., KUMAR, S. & SELVARAJ, S. 2012. Antiobiotic
resistance pattern of biofilm-forming uropathogens isolated from catheterised patients in
Pondicherry, India. Australas Med J, 5, 344-8.
SUN, L. 2013. Peptide-Based Drug Development. Mod Chem appl 1.
SVANBORG, C. & GODALY, G. 1997. Bacterial virulence in urinary tract infection. Infect Dis Clin North Am,
11, 513-29.
TENKE, P., KOVACS, B., JACKEL, M. & NAGY, E. 2006. The role of biofilm infection in urology. World J Urol,
24, 13-20.
THOMASSIN, J. L., BRANNON, J. R., GIBBS, B. F., GRUENHEID, S. & LE MOUAL, H. 2012. OmpT outer
membrane proteases of enterohemorrhagic and enteropathogenic Escherichia coli contribute
differently to the degradation of human LL-37. Infect Immun, 80, 483-92.
TOLLIN, M., BERGMAN, P., SVENBERG, T., JORNVALL, H., GUDMUNDSSON, G. H. & AGERBERTH, B. 2003.
Antimicrobial peptides in the first line defence of human colon mucosa. Peptides, 24, 523-30.
TSAI, P. W., YANG, C. Y., CHANG, H. T. & LAN, C. Y. 2011. Human antimicrobial peptide LL-37 inhibits
adhesion of Candida albicans by interacting with yeast cell-wall carbohydrates. PLoS One, 6,
e17755.
TSUBAKISHITA, S., KUWAHARA-ARAI, K., SASAKI, T. & HIRAMATSU, K. 2010. Origin and molecular evolution
of the determinant of methicillin resistance in staphylococci. Antimicrob Agents Chemother, 54,
4352-9.
TURNER, J., CHO, Y., DINH, N. N., WARING, A. J. & LEHRER, R. I. 1998. Activities of LL-37, a cathelinassociated antimicrobial peptide of human neutrophils. Antimicrob Agents Chemother, 42, 2206-14.
VALLET, I., OLSON, J. W., LORY, S., LAZDUNSKI, A. & FILLOUX, A. 2001. The chaperone/usher pathways of
Pseudomonas aeruginosa: identification of fimbrial gene clusters (cup) and their involvement in
biofilm formation. Proc Natl Acad Sci U S A, 98, 6911-6.
VANDAMME, D., LANDUYT, B., LUYTEN, W. & SCHOOFS, L. 2012. A comprehensive summary of LL-37, the
factotum human cathelicidin peptide. Cell Immunol, 280, 22-35.
VEJBORG, R. M. & KLEMM, P. 2008. Blocking of bacterial biofilm formation by a fish protein coating. Appl
Environ Microbiol, 74, 3551-8.
VUONG, C. & OTTO, M. 2002. Staphylococcus epidermidis infections. Microbes Infect, 4, 481-9.
WANG, G. 2008. Structures of human host defense cathelicidin LL-37 and its smallest antimicrobial peptide
KR-12 in lipid micelles. J Biol Chem, 283, 32637-43.
WANG, G., EPAND, R. F., MISHRA, B., LUSHNIKOVA, T., THOMAS, V. C., BAYLES, K. W. & EPAND, R. M. 2012.
Decoding the functional roles of cationic side chains of the major antimicrobial region of human
cathelicidin LL-37. Antimicrob Agents Chemother, 56, 845-56.
WANG, T. T., NESTEL, F. P., BOURDEAU, V., NAGAI, Y., WANG, Q., LIAO, J., TAVERA-MENDOZA, L., LIN, R.,
HANRAHAN, J. W., MADER, S. & WHITE, J. H. 2004. Cutting edge: 1,25-dihydroxyvitamin D3 is a
direct inducer of antimicrobial peptide gene expression. J Immunol, 173, 2909-12.
WANG, X., LUNSDORF, H., EHREN, I., BRAUNER, A. & ROMLING, U. 2010. Characteristics of biofilms from
urinary tract catheters and presence of biofilm-related components in Escherichia coli. Curr
Microbiol, 60, 446-53.
WARREN, J. W. 2001. Catheter-associated urinary tract infections. Int J Antimicrob Agents, 17, 299-303.
WEISS, G. A., WATANABE, C. K., ZHONG, A., GODDARD, A. & SIDHU, S. S. 2000. Rapid mapping of protein
functional epitopes by combinatorial alanine scanning. Proc Natl Acad Sci U S A, 97, 8950-4.
WELCH, R. A., BURLAND, V., PLUNKETT, G., 3RD, REDFORD, P., ROESCH, P., RASKO, D., BUCKLES, E. L., LIOU,
S. R., BOUTIN, A., HACKETT, J., STROUD, D., MAYHEW, G. F., ROSE, D. J., ZHOU, S., SCHWARTZ, D. C.,
PERNA, N. T., MOBLEY, H. L., DONNENBERG, M. S. & BLATTNER, F. R. 2002. Extensive mosaic
structure revealed by the complete genome sequence of uropathogenic Escherichia coli. Proc Natl
Acad Sci U S A, 99, 17020-4.
WHITELEY, M., BANGERA, M. G., BUMGARNER, R. E., PARSEK, M. R., TEITZEL, G. M., LORY, S. & GREENBERG,
E. P. 2001. Gene expression in Pseudomonas aeruginosa biofilms. Nature, 413, 860-4.

78

WIECZOREK, M., JENSSEN, H., KINDRACHUK, J., SCOTT, W. R., ELLIOTT, M., HILPERT, K., CHENG, J. T.,
HANCOCK, R. E. & STRAUS, S. K. 2010. Structural studies of a peptide with immune modulating and
direct antimicrobial activity. Chem Biol, 17, 970-80.
WIEGAND, I., HILPERT, K. & HANCOCK, R. E. 2008. Agar and broth dilution methods to determine the
minimal inhibitory concentration (MIC) of antimicrobial substances. Nat Protoc, 3, 163-75.
WILLCOX, M. D., HUME, E. B., ALIWARGA, Y., KUMAR, N. & COLE, N. 2008. A novel cationic-peptide coating
for the prevention of microbial colonization on contact lenses. J Appl Microbiol, 105, 1817-25.
WONG, J. H., LEGOWSKA, A., ROLKA, K., NG, T. B., HUI, M., CHO, C. H., LAM, W. W., AU, S. W., GU, O. W. &
WAN, D. C. 2011. Effects of cathelicidin and its fragments on three key enzymes of HIV-1. Peptides,
32, 1117-22.
WOZNIAK, D. J., WYCKOFF, T. J., STARKEY, M., KEYSER, R., AZADI, P., O'TOOLE, G. A. & PARSEK, M. R. 2003.
Alginate is not a significant component of the extracellular polysaccharide matrix of PA14 and PAO1
Pseudomonas aeruginosa biofilms. Proc Natl Acad Sci U S A, 100, 7907-12.
WRIGHT, G. D. 2007. The antibiotic resistome: the nexus of chemical and genetic diversity. Nat Rev
Microbiol, 5, 175-86.
WRIGHT, G. D. 2012. The origins of antibiotic resistance. Handb Exp Pharmacol, 13-30.
WUERTH, K. & HANCOCK, R. E. 2011. New insights into cathelicidin modulation of adaptive immunity. Eur J
Immunol, 41, 2817-9.
XIONG, Y. Q., HADY, W. A., DESLANDES, A., REY, A., FRAISSE, L., KRISTENSEN, H. H., YEAMAN, M. R. &
BAYER, A. S. 2011. Efficacy of NZ2114, a novel plectasin-derived cationic antimicrobial peptide
antibiotic, in experimental endocarditis due to methicillin-resistant Staphylococcus aureus.
Antimicrob Agents Chemother, 55, 5325-30.
YAMASAKI, K., SCHAUBER, J., CODA, A., LIN, H., DORSCHNER, R. A., SCHECHTER, N. M., BONNART, C.,
DESCARGUES, P., HOVNANIAN, A. & GALLO, R. L. 2006. Kallikrein-mediated proteolysis regulates
the antimicrobial effects of cathelicidins in skin. FASEB J, 20, 2068-80.
YASIN, B., PANG, M., TURNER, J. S., CHO, Y., DINH, N. N., WARING, A. J., LEHRER, R. I. & WAGAR, E. A. 2000.
Evaluation of the inactivation of infectious Herpes simplex virus by host-defense peptides. Eur J Clin
Microbiol Infect Dis, 19, 187-94.
YEAMAN, M. R. & YOUNT, N. Y. 2003. Mechanisms of antimicrobial peptide action and resistance.
Pharmacol Rev, 55, 27-55.
YOUNT, N. Y. & YEAMAN, M. R. 2013. Peptide antimicrobials: cell wall as a bacterial target. Ann N Y Acad
Sci, 1277, 127-38.
ZANETTI, M. 2004. Cathelicidins, multifunctional peptides of the innate immunity. J Leukoc Biol, 75, 39-48.
ZANETTI, M. 2005. The role of cathelicidins in the innate host defenses of mammals. Curr Issues Mol Biol, 7,
179-96.
ZANETTI, M., GENNARO, R. & ROMEO, D. 1995. Cathelicidins: a novel protein family with a common
proregion and a variable C-terminal antimicrobial domain. FEBS Lett, 374, 1-5.
ZASLOFF, M. 1987. Magainins, a class of antimicrobial peptides from Xenopus skin: isolation,
characterization of two active forms, and partial cDNA sequence of a precursor. Proc Natl Acad Sci
U S A, 84, 5449-53.
ZASLOFF, M. 2002. Antimicrobial peptides of multicellular organisms. Nature, 415, 389-95.
ZASLOFF, M. 2007. Antimicrobial peptides, innate immunity, and the normally sterile urinary tract. J Am Soc
Nephrol, 18, 2810-6.

79

Appendix I
Below, the plotted data from the biofilm experiments is listed.

Effect of LL-37, 1037 and antibiotics on P. aeruginosa biofilms


Table 8: Effect of LL-37 on P. aeruginosa biofilms. The mean is given as the percentage of untreated samples with standard
deviation (SD) and number of replicates (N).

Concentration
Mean (%)
(g/ml)

SD

100.00

16.65

15

0.5

110.60

13.26

78.14

23.53

67.84

24.37

53.73

23.00

52.69

9.01

16

41.64

25.26

32

44.08

7.00

64

36.39

27.91

128

51.94

31.33

Table 9: Effect of 1037 on P. aeruginosa biofilms. The mean is given as the percentage of untreated samples with standard
deviation (SD) and number of replicates (N).

Concentration
Mean (%)
(g/ml)

SD

100.00

9.21

71.83

22.43

93.56

28.02

53.48

19.67

16

42.86

33.80

32

31.20

20.27

64

28.06

35.84

128

22.15

27.48

80

Table 10: Effect of tobramycin on P. aeruginosa biofilms. The mean is given as the percentage of untreated samples with standard
deviation (SD) and number of replicates (N).

Concentration
Mean (%)
(g/ml)

SD

100.00

21.43

15

0.01

97.42

30.52

0.025

110.20

24.22

0.05

91.96

35.95

0.1

92.42

29.21

0.2

142.40

40.31

Table 11 Effect of tetracycline on P. aeruginosa biofilms. The mean is given as the percentage of untreated samples with standard
deviation (SD) and number of replicates (N).

Concentration
Mean (%)
(g/ml)

SD

100.00

14.16

15

0.1

118.20

25.45

0.2

107.10

17.44

0.4

88.13

16.56

0.8

93.29

30.98

1.6

104.20

46.96

3.1

97.09

41.61

6.2

107.00

44.22

12.5

134.60

36.57

Table 12: Effect of LL-37 on P. aeruginosa initial attachment. The inhibition shown as % of one untreated sample.

Concentration (g/ml)

16

32

64

128

% of control

143.84

99.29

118.72

117.77

105.45

64.22

29.14

9.24

Table 13: Effect of 1037 on P. aeruginosa initial attachment. The inhibition shown as % of one untreated sample.

Concentration (g/ml)

16

32

64

128

% of control

108.53

91.79

78.58

70.53

70.85

52.82

81

Table 14: The 24h effect of LL-37 on pregrown P. aeruginosa biofilms. The mean is given as the percentage of untreated samples
with standard deviation (SD) and number of replicates (N).

Concentration
Mean (%)
(g/ml)

SD

100

10.24

82.77

15.32

16

44.36

12.54

Table 15: The 2h effect of LL-37 on pregrown P. aeruginosa biofilms. The mean is given as the percentage of untreated samples
with standard deviation (SD) and number of replicates (N).

Concentration
Mean (%)
(g/ml)

SD

100

32.77

114.8

14.86

16

120.8

28.98

Table 16: The 24h effect of 1037 on pregrown P. aeruginosa biofilms. The mean is given as the percentage of untreated samples
with standard deviation (SD) and number of replicates (N).

Concentration
Mean (%)
(g/ml)

SD

100

12.91

116.7

9.97

91.74

3.861

Table 17: The 2h effect of 1037 on pregrown P. aeruginosa biofilms. The mean is given as the percentage of untreated samples
with standard deviation (SD) and number of replicates (N).

Concentration
Mean (%)
(g/ml)

SD

100

29.70

96.31

30.13

83.90

34.14

82

Effect of LL-37 and 1037 on K. pneumoniae biofilms


Table 18: The effect of LL-37 in BM2 medium on K. pneumoniae biofilms in polypropylene microtiter plates. The mean is
given as the percentage of untreated samples with standard deviation (SD) and number of replicates (N).

Concentration (g/ml)

16

32

64

128

Mean (%)

155.8

113.8

139.3

171.1

161.9

206.1

SD

110.5

67.96

79.21

113.0

103.8

208.0

Table 19: The effect of LL-37 in M9 minimal medium on K. pneumoniae biofilms in polypropylene microtiter plates. The
mean is given as the percentage of untreated samples with standard deviation (SD) and number of replicates (N).

Concentration (g/ml)

16

32

64

128

Mean (%)

53.32

57.43

51.88

86.18

89.07

57.44

SD

6.723

20.34

27.66

36.25

53.15

23.90

Table 20: The effect of LL-37 in LB medium on K. pneumoniae biofilms in polypropylene microtiter plates. The mean is given
as the percentage of untreated samples with standard deviation (SD) and number of replicates (N).

Concentration (g/ml)

16

32

64

128

Mean (%)

92.46

123.0

97.64

171.4

158.5

186.4

SD

32.37

81.68

39.89

149.2

118.5

98.26

Table 21: The effect of LL-37 in MH medium on K. pneumoniae biofilms in polypropylene microtiter plates. The mean is given
as the percentage of untreated samples with standard deviation (SD) and number of replicates (N).

Concentration (g/ml)

16

32

Mean (%)

81.83

89.64

74.10

98.36

SD

11.56

89.60

27.29

60.23

83

Table 22: The effect of LL-37 in TB medium on K. pneumoniae biofilms in polypropylene microtiter plates. The mean is given
as the percentage of untreated samples with standard deviation (SD) and number of replicates (N).

Concentration (g/ml)

16

32

64

128

Mean (%)

107.2

102.9

109.2

120.9

101.7

90.94

SD

13.98

25.15

51.34

56.19

42.57

39.91

Table 23: The time growth of K. pneumoniae biofilm formation in M9. The mean is given as the OD595 with standard deviation
(SD) and number of replicates (N).

Time (h)

24

Mean

0.113

0.293

0.293

0.225

SD

0.018

0.085

0.033

0.072

Table 24: The time growth of K. pneumoniae biofilm formation in BM2. The mean is given as the OD595 with standard deviation
(SD) and number of replicates (N).

Time (h)

24

Mean

0.173

0.228

0.182

0.507

SD

0.040

0.042

0.037

0.176

Table 25: The 24h effect of LL-37 on K. pneumoniae biofilms. The inhibition shown as % of one untreated sample.

Concentration (g/ml)

16

32

64

128

% of control

107.04

71.36

71.95

101.98

89.79

92.77

114.77

121.31

Table 26: The 24h effect of 1037 on K. pneumoniae biofilms. The inhibition shown as % of one untreated sample.

Concentration (g/ml)

16

32

64

128

% of control

85.23

126.83

114.32

89.97

88.61

88.95

75.76

Table 27: The 4h effect of LL-37 on K. pneumoniae biofilms. The inhibition shown as % of one untreated sample.

Concentration (g/ml)
% of control

1
90.90

2
110.33

84

98.30

98.76

16
126.52

32
79.10

64
29.37

128
23.13

Table 28: The 4h effect of 1037 on K. pneumoniae biofilms. The inhibition shown as % of one untreated sample.

Concentration (g/ml)

16

32

64

128

% of control

59.67

55.28

78.64

61.98

67.30

41.17

19.89

Table 29: Effect of LL-37 on K. pneumoniae initial attachment. The mean is given as the percentage of untreated samples with
standard deviation (SD) and number of replicates (N).

Concentration
Mean (%)
(g/ml)

SD

100.00

10.83

12

73.41

23.68

72.20

17.80

71.33

12.19

62.90

17.20

16

34.97

23.61

32

14.66

19.94

64

7.29

14.41

128

0.17

5.94

Table 30: Effect of 1037 on K. pneumoniae initial attachment. The mean is given as the percentage of untreated samples with
standard deviation (SD) and number of replicates (N).

Concentration
Mean (%)
(g/ml)

SD

100.00

16.80

12

54.43

13.12

49.08

19.33

39.60

18.57

16

46.96

24.51

32

28.55

14.48

64

19.85

15.26

128

13.04

11.59

Table 31: The 3h effect of LL-37 on pregrown K. pneumoniae biofilms. The mean is given as the percentage of untreated samples
with standard deviation (SD) and number of replicates (N).

Concentration
Mean (%)
(g/ml)

SD

116.2

12.15

16

115.0

6.375

85

Table 32: The 3h effect of 1037 on pregrown K. pneumoniae biofilms. The mean is given as the percentage of untreated samples
with standard deviation (SD) and number of replicates (N).

Concentration
Mean (%)
(g/ml)

SD

103.3

12.45

32

125.4

6.642

86

Effect of LL-37 and 1037 on E. coli biofilms


Table 33: Effect of LL-37 on E. coli biofilms. The mean is given as the percentage of untreated samples with standard deviation
(SD) and number of replicates (N).

Concentration
(g/ml)

E. coli 536

E. coli CFT073

Mean

SD

Mean

SD

100.00

14.80

15

100.00

16.22

14

113.30

16.73

94.68

42.87

102.50

10.18

70.66

10.55

92.33

6.89

63.04

18.45

94.15

6.35

74.99

21.23

16

76.34

20.84

83.28

34.45

32

67.22

7.93

52.94

25.61

64

75.40

11.17

66.83

13.01

128

68.40

9.20

Table 34: Effect of 1037 on E. coli biofilms. The mean is given as the percentage of untreated samples with standard deviation (SD)
and number of replicates (N).

Concentration
(g/ml)

E. coli 536

E. coli CFT073

Mean

SD

Mean

SD

100.00

9.35

16

100.00

17.10

15

73.06

16.00

101.40

28.15

79.00

19.66

85.48

15.95

79.79

16.10

111.70

29.07

16

70.58

13.93

80.53

13.61

87

S-ar putea să vă placă și