243
Discovery of antimicrobial peptides active against antibiotic resistant bacterial pathogens A thesis submitted to the University of Manchester for the degree of Doctor of Philosophy in the Faculty of Medical and Human Sciences 2015 Arif Felek School of Medicine

Discovery of antimicrobial peptides active against

  • Upload
    others

  • View
    8

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Discovery of antimicrobial peptides active against

Discovery of antimicrobial peptides active against antibiotic resistant bacterial pathogens

A thesis submitted to the University of Manchester for the degree of

Doctor of Philosophy in the Faculty of Medical and Human Sciences

2015

Arif Felek

School of Medicine

Page 2: Discovery of antimicrobial peptides active against

List of contents

2 | P a g e

LIST OF CONTENTS

List of contents ....................................................................................................... 2

List of figures ......................................................................................................... 8

List of tables ......................................................................................................... 18

List of abbreviations ............................................................................................. 20

Declaration ........................................................................................................... 23

Copyright statement ............................................................................................. 24

Dedication ............................................................................................................ 25

Acknowledgements .............................................................................................. 26

Publications from this work ................................................................................. 27

1 Chapter 1 General Introduction .................................................................... 28

1.1 Downfall of novel antimicrobial discovery ........................................... 32

1.1.1 Movement of “big pharma” away from antimicrobial drug

discovery ....................................................................................................... 32

1.1.2 Overview of target based genetic approaches and their failure to

discover novel antibiotics ............................................................................. 34

1.2 Revitalising natural product screening as a source of novel antibiotics 36

1.3 Bacterially derived natural products of interest..................................... 39

1.4 Bacteriocins ........................................................................................... 41

1.5 Bacteriocins of Gram positive bacteria ................................................. 43

1.5.1 Class I - Lanthionine containing bacteriocins (Lantibiotics) ......... 43

Page 3: Discovery of antimicrobial peptides active against

List of contents

3 | P a g e

1.5.2 Class II non-lanthionine containing bacteriocins ........................... 48

1.5.3 Class III large proteins: .................................................................. 50

1.5.4 Class IV Cyclic peptides ................................................................ 51

1.6 Bacteriocins of Gram-negative bacteria ................................................ 52

1.6.1 Colicins .......................................................................................... 52

1.6.2 Microcins ....................................................................................... 54

1.7 Miscellaneous RiPP bacteriocins .......................................................... 58

1.7.1 Sactibiotics (Sulphur to α-carbon antibiotics) or sactipeptides ...... 58

1.7.2 Thiopeptides ................................................................................... 58

1.7.3 Bottromycins .................................................................................. 59

1.7.4 Glycocins ....................................................................................... 59

1.7.5 Linear azol(in)e-containing peptides (LAPs) and lasso peptides ... 60

1.8 Polyketides ............................................................................................ 61

1.9 Non-ribosomal proteins ......................................................................... 62

1.10 Lysins: ................................................................................................... 64

1.11 Advances in natural product screening.................................................. 65

1.11.1 Purification and identification of the active compound ................. 65

1.11.2 Identification of producer strains ................................................... 67

1.11.3 Genome mining and scanning for natural products ....................... 67

1.12 Aims and objectives .............................................................................. 72

2 Chapter 2 Methods ........................................................................................ 74

Page 4: Discovery of antimicrobial peptides active against

List of contents

4 | P a g e

2.1 In vitro antimicrobial peptide screening and characterisation ............... 75

2.1.1 Bacterial isolates ............................................................................ 75

2.1.2 Identification of the producer strains using 16S rRNA gene

sequencing .................................................................................................... 76

2.1.3 Agarose gel electrophoresis ........................................................... 78

2.1.4 Assays used for the assessment of antimicrobial activity .............. 78

2.1.5 Optimisation of antimicrobial peptide production ......................... 83

2.2 Proteomic analysis of the identified antimicrobial peptides.................. 84

2.2.1 Purification of antimicrobial peptides from the culture media ...... 84

2.2.2 SDS-PAGE analysis and gel diffusion agar overlay assay ............ 88

2.2.3 MALDI-ToF mass determination and ESI-MS scan of the RP-

HPLC Purified Active Fractions ................................................................... 89

2.2.4 De novo peptide sequencing (MS/MS) .......................................... 90

2.2.5 Genome sequencing and annotation of the draft genome of

producer strains ............................................................................................. 93

2.3 Characterization of the purified antimicrobial peptides ........................ 94

UV absorbing material leakage assay ........................................................... 95

2.3.1 Physicochemical analysis of purified antimicrobial peptides (heat

and enzyme stability tests) ............................................................................ 95

2.3.2 Determination of minimum inhibitory concentration .................... 96

2.3.3 Haemolysis assay ........................................................................... 96

2.3.4 Eukaryotic cell toxicity studies ...................................................... 97

Page 5: Discovery of antimicrobial peptides active against

List of contents

5 | P a g e

2.3.5 UV Absorbing material leakage assay ........................................... 99

2.4 In silico Genome mining and identification of putative bacteriocin leads

………………………………………………………………………...99

2.5 Cloning and expression methods ......................................................... 102

2.5.1 Insert construction and primer design for cell free protein

expression ................................................................................................... 102

2.5.2 In vitro cell-free expression ......................................................... 106

2.5.3 Proteomic analysis of peptide products generated following

expression studies ....................................................................................... 107

3 Chapter 3 Pumicin NI04, a novel antimicrobial peptide from Bacillus

pumilus, is homologous to the EsxA virulence determinant .............................. 108

3.1 Introduction ......................................................................................... 109

3.2 Results ................................................................................................. 113

3.2.1 Two antimicrobial agents are produced by B. pumilus J1 ........... 113

3.2.2 Bacteriocin production can be optimised by manipulation of

growth conditions ....................................................................................... 115

3.2.3 De novo peptide sequence determination of peptide NI04 using

mass spectrometry analysis and genome interrogation facilitates

identification of the pumicin NI04 locus .................................................... 117

3.2.4 Pumicin NI04 is stable to heat treatments and protease degradation

confirms a proteinaceous nature ................................................................. 125

3.2.5 Pumicin NI04 has potent activity against resistant Gram positive

pathogens .................................................................................................... 126

Page 6: Discovery of antimicrobial peptides active against

List of contents

6 | P a g e

3.2.6 UV absorbing material leakage assay .......................................... 127

3.2.7 Pumicin NI04 has low levels of in vitro toxicity ......................... 130

3.3 Discussion ........................................................................................... 132

4 Chapter 4 Discovery and Analysis Of Peptide NI05 produced by Klebsiella

pneumoniae strain A7 ........................................................................................ 141

4.1 Introduction ......................................................................................... 142

4.2 Results: ................................................................................................ 147

4.2.1 Simultaneous antagonism assays for preliminary analysis of the

antimicrobial activity spectrum of peptide NI05 ........................................ 147

4.2.2 Optimisation of antimicrobial peptide production ....................... 148

4.2.3 Purification ................................................................................... 149

4.2.4 Matrix Assisted Laser Desorption Ionisation – Time of Flight

(MALDI-ToF) mass determination ............................................................ 150

4.2.5 Genome sequence and annotation ................................................ 151

4.2.6 In silico analysis of the draft genome and discovery of the

MccE492 gene cluster:................................................................................ 152

4.3 Discussion ........................................................................................... 158

5 Chapter 5 Genome mining of anaerobic bacteria, known producer bacteria

and cell free expression of pumicin NI04 .......................................................... 166

5.1 Introduction ......................................................................................... 167

5.2 Results ................................................................................................. 170

5.3 Genome mining results ........................................................................ 170

Page 7: Discovery of antimicrobial peptides active against

List of contents

7 | P a g e

5.3.1 Lead 3 ........................................................................................... 176

5.3.2 Lead 12 ......................................................................................... 177

5.3.3 Lead 14 ......................................................................................... 178

5.3.4 Unique putative bacteriocin candidates ....................................... 178

5.4 Cloning Results ................................................................................... 181

5.4.1 Insert confirmation and validation following transformation ...... 182

5.4.2 In vitro expression of pumicin NI04 ............................................ 186

5.5 Discussion ........................................................................................... 188

Chapter 6 Concluding Remarks ......................................................................... 199

References .......................................................................................................... 205

Total words Count: 53,113

Page 8: Discovery of antimicrobial peptides active against

List of figures

8 | P a g e

LIST OF FIGURES

Figure 1.1 Timeline of antibiotic discovery versus outbreak of resistance among

bacteria [from (Clatworthy et al., 2007)] ............................................................. 29

Figure 1.2 The amount of antibiotics being introduced into the clinic has been

drastically decreasing in recent years [From (Boucher et al., 2013)] .................. 30

Figure 1.3 Graphical representation and comparison of the number of antibiotic

candidates going through clinical trials between 2011 and 2013 [from (Pucci and

Bush, 2013)]. ........................................................................................................ 31

Figure 1.4 The process of target based antimicrobial drug discovery [adapted

from (Mills, 2006; Pucci, 2007)]. ........................................................................ 34

Figure 1.5 Advantages of natural products (NP) as antibiotic drug leads ........... 36

Figure 1.6 Natural products of interest in the current project. The figure above

lists the natural products produced by bacteria that have the potential to be used

as antibiotics. These products are shortly reviewed in this study. ....................... 40

Figure 1.7 The "Universal” bacteriocin classification scheme. This figure by

Heng and Tagg, (2006) illustrates the two major classification schemes that are

used and how they propose to combine the respective advantages to construct a

universal scheme. ................................................................................................. 42

Figure 1.8 Bacteriocin mode of action. The class II bacteriocins such as sakacin

can insert into the cell membrane by interacting with membrane proteins and

cause pore formation and depolarisation of the cell membrane. The class I

bacteriocins on the other hand can have a single or multiple modes of action such

as nisin. Large protein bacteriocins such as lysostaphin can also act on the cell

Page 9: Discovery of antimicrobial peptides active against

List of figures

9 | P a g e

wall by hydrolysing the interpeptide bridges of the peptidogycan layer [adapted

from (Cotter et al. 2005)]. .................................................................................... 44

Figure 1.9 Four classes of the lanthipeptide family can be separated by the

modification enzymes involved in generation of their lanthionine bridges. Solid

lines indicate the conserved regions thought to play roles in the catalysis

including zinc ligands that are missing in class III [from (Knerr and van der

Donk, 2012)]. ....................................................................................................... 46

Figure 1.10 Class II-A bacteriocin action is dependent on the presence of IID

and IIC proteins, which are membrane proteins of the Man-PTS mechanism.

Class II a bacteriocin first binds to the extracellular loop and then to the

transmembrane helices of these peptides causing conformational changes that

leads to efflux and cell death [from (Kjos et al., 2011)] ...................................... 49

Figure 1.11 Mechanisms of cellular entry used by colicins. ExbB, ExbD

components in the TonB pathway are used by group B colicins and the Tol

pathway (TolA, TolB, TolQ and TolR) is employed by group A colicins [from

(Lloubès et al., 2012)]. ......................................................................................... 53

Figure 1.12 Class I microcin action. A) Microcin (Mcc) J25 gains access to the

cell through the outer membrane (OM) receptor FhuA and then utilises the

TonB/ExbB/ExbD complex and SbmA protein to translocate into the cytoplasm

where it inhibits transcription by attacking the RNA polymerase enzyme

(Mathavan et al., 2014). B) Mcc B17 targets the DNA gyrase enzyme that is

responsible for supercoiling of the DNA, it gains access to the inner membrane

(IM) using the OmpF porin on the OM and is translocated into the cell by

exploiting the SbmA protein. C) Although the entry mechanism of Mcc C7/C51

Page 10: Discovery of antimicrobial peptides active against

List of figures

10 | P a g e

is unknown, it acts by inhibiting aspartyl-tRNA synthetase [from (Duquesne et

al., 2007a)]. .......................................................................................................... 55

Figure 1.13 Class IIb microcin action MccE492 (A), MccM (B), MccH47(C)

and MccI47 (D)* all exploit the catechol-siderophore receptors and are

translocated into the periplasm in a TonB pathway dependent manner. Once

inside, MccE492, using the ManY and ManZ membrane proteins of the mannose

permease complex as docking stations, inserts itself into the inner membrane

where it disrupts the IM potential by allowing protons to leak. On the other hand,

MccH47 inhibits ATP synthase by interfering with the F0 component once it

gains access to the periplasm. The exact mechanism of action for the remaining

microcins of this group is not known [modified from (Duquesne et al., 2007a)] .

*MccI47 peptide has not been isolated, but on the basis of predicted amino acid

sequence it contains the serine rich C-terminal that is associated with siderophore

interaction. ............................................................................................................ 57

Figure 1.14 A diagram of Microcin J25 structure, demonstrating the lasso fold

[from (Maksimov et al., 2012)] ........................................................................... 60

Figure 1.15 Examples of NRP structures. Tyrocidine A is a good example for

macrocyclic NRPs, while surfactine is a good a good example of branched

macrocyclic NRPs [from (Schwarzer et al., 2003)] ............................................. 63

Figure 1.16 A hybrid NRP/PK. NRPS derived structure is coloured in blue while

PKS derived structures are red [from (Walsh and Fischbach, 2010)].................. 64

Figure 2.1 Simultaneous antagonism test, the antimicrobial activity of potential

inhibitor producers against two indicator E. coli species (414 at the top and

DH5α at bottom of the plate) was observed as clear zones of inhibition............. 79

Page 11: Discovery of antimicrobial peptides active against

List of figures

11 | P a g e

Figure 2.2 Deferred antagonism assay plate configuration and scoring system

[From (Tagg and Bannister, 1979)].* Producer is removed and plate is

chloroformed prior to inncoulation of the indicator strains. ................................ 81

Figure 2.3 Image above shows a well diffusion assay performed against

indicator organism M. luteus. ............................................................................... 82

Figure 2.4 Spot assay shown in the diagram illustrates antimicrobial activity of a

sample against M. luteus. ..................................................................................... 83

Figure 2.5 The three-step strategy used for peptide purification. A graphical

representation of the purification strategy employed in this project [from

(Amersham Biosciences)]. ................................................................................... 84

Figure 2.6 Overview of the proposed workflow described below for in silico

mining of genomic sequence data for novel AMPs ........................................... 101

Figure 2.7 Utilisation of NcoI cut site using the BsaI restriction site. .............. 103

Figure 3.1 Spot assay performed with Sep-Pak C18 purified fractions against S.

aureus. Peptide NI03 eluted at 40% acetonitrile concentration and peptide NI04

at 60 and 70% acetonitrile fractions. .................................................................. 114

Figure 3.2 Availability of peptide NI04 in tryptic soy broth and nutrient broth

culture media under different incubation temperatures over a 24 hour period.

Error bars represent the standard deviation in peptide NI04 activity between

replicates. ........................................................................................................... 116

Figure 3.3 RP-HPLC chromatogram of peptide NI04. Each peak on the graph

represents the molecules or molecule groups eluted during the RP-HPLC

procedure at different acetonitrile concentrations (green line indicates the

acetonitrile concentration). The readings are taken at 215nm UV intensity.

Active fractions are labelled............................................................................... 117

Page 12: Discovery of antimicrobial peptides active against

List of figures

12 | P a g e

Figure 3.4 ESI-MS scan of active HPLC fraction showing the detected mass of

peptide NI04 as 10722.993. Other peaks represent background noise and a

possible impurity with a mass of 10771.341 (not present in other spectrographs).

............................................................................................................................ 118

Figure 3.5 A graphical representation of the differences between the genomes of

the B. pumilus SAFR-32 strain (inner ring) and B. pumilus J1 (outer, purple

ring). Comparisons were made with BRIG software. ........................................ 119

Figure 3.6 Protein sequence alignment of pumicin NI04 and the 10 most

homologous peptides identified using the PSI-BLAST algorithm (Altschul et al.,

1997). Also included are the known EsxA peptide sequences (EsxA [M.

tuberculosis] and virulence factor EsxA [S.aureus]) that are experimentally

confirmed to play a role in the virulence of their respective host organisms

(boxed in Green) and the highly similar EsxA homolog YukE peptide of

unknown function from B. subtilis (boxed in orange). Secondary structure

prediction, performed using jnetpred, is illustrated at the bottom of the

alignments and confirms the helix turn helix structure of WXG-100 peptides is

preserved in pumicin NI04 together with the W-X-G motif. ............................. 122

Figure 3.7 Neighbour joining tree calculated using the % identity of aligned

WXG family peptides. Branches are labelled with the % difference in the identity

of the amino acid sequences to each other. ........................................................ 123

Figure 3.8 Organisation of the biosynthetic cluster predicted to be involved in

the production and secretion of pumicin NI04 encoded in the producer B.

pumillus J1 genome and organisation of other operons encoding homologous

EsxA peptides, M. tuberculosis ESX1 locus (Burts et al., 2005), B. subtilis Yuk

locus (Huppert et al., 2014) and S. aureus Ess locus (Kneuper et al., 2014). Esx

Page 13: Discovery of antimicrobial peptides active against

List of figures

13 | P a g e

substrates are highlighted in red and proteins that are known to be involved in

regulation and secretions of ESX substrates are coloured blue, proteins that are

predicted to be involved in transport or regulation of these factors are highlighted

in green, proteins that have unknown or unrelated function are coloured grey

(Burts et al., 2005; Kneuper et al., 2014; McLaughlin et al., 2007; Ramsdell et

al., 2015). PROKKA 03052 (70.8% identity with YukBA [UniProt ref: C0SPA7]

), YukD (73% identity with YukD [UniProt ref: P71071]), YukC (100% identity

[UniProt ref: P71070]), PROKKA 03051 (45.86% identity with YueB [UniProt

ref: O32101]), PROKKA 03050 (33.5% identity with YueC [UniProt ref:

O32100]). ........................................................................................................... 124

Figure 3.9 The SDS-PAGE gel analysis of the Sep-Pak C18 purified samples,

against SeeBlu 2 protein marker. Gel overlay assay performed following analysis

demonstrated the peptide band at the expected mass interval has inhibitory

activity. The band indicated with the orange arrow was used for MS/MS

fragmentation and de novo amino acid sequencing. .......................................... 125

Figure 3.10 Release of UV absorbing materials from A) M. luteus and B) S.

aureus cells over time following treatment with pumicin NI04 at 1xMIC of

the test organism compared with pumicin NI04 solution (1xMIC) in test buffer

without any microbial inoculum to account for the absorbance inferred by

pumicin NI04. .................................................................................................... 129

Figure 3.11 Comparison of neutral red uptake of Vero cells against negative

(growth media) and positive (2% triton X-100) controls, following incubation

with differing concentrations of pumicin NI04 of up to 18x the MIC recorded

against M. luteus................................................................................................. 131

Page 14: Discovery of antimicrobial peptides active against

List of figures

14 | P a g e

Figure 3.12 Cell survival data collected using the trypan blue exclusion assay,

following a 24 hour incubation of keratinocyte cells with pumicin NI04

supplemented medium against negative (growth media) and positive (20mg/ml

SDS) controls. .................................................................................................... 131

Figure 3.13 A representation of the multidomain ATPase involved in ESX

substrate secretion. It is believed that the first domain of YukBA generates the

energy required for the translocation (dark blue) and the remaining two interact

with the substrate (lighter shades of blue) such as pumicin NI04 [from (Ramsdell

et al., 2015)] . ..................................................................................................... 140

Figure 4.1 Activity spectrum of prominent carbapenemases found in

enterobacteriaceae species [adapted from (Nordmann, 2014)]. ......................... 143

Figure 4.2 Prevelance of NDM related cases per country (Dortet et al., 2014)

............................................................................................................................ 144

Figure 4.3 HPLC chromatogram of the active fractions I3 and I4 obtained during

purification of peptide NI05. .............................................................................. 149

Figure 4.4 Mass spectrogram of three separate active HPLC fractions from

individual purification runs. The spectrum encompassed molecules up to 10kDa

in size, however the only detectable peaks were perceived below 2500Da....... 150

Figure 4.5 Draft genome sequence of producer organism K. pneumoniae A7

compared with the reference K. pneumoniae strain XH209 and K. pneumoniae

strain RYC492 that is known to express MccE492, using BRIG software. The

identical regions between both genomes are represented in solid colours and

differences are represented with faded colours or gaps. .................................... 152

Figure 4.6 Alignment of sequences of the structural genes encoding microcins

MccM, MccH47, MccE492 and MccG492 reveals striking homology at the C-

Page 15: Discovery of antimicrobial peptides active against

List of figures

15 | P a g e

terminus. Red arrow indicates the residue (Serine 84) where the post translational

moiety DHBS is attached to the MccE492 and red line indicates that the region

around this location that is also conserved in other members of the group. The

green line indicates the conserved GG/GA motif where leader sequences are

cleaved from the active peptide.......................................................................... 155

Figure 4.7 MceK3 peptide sequence is homologous to the MccM N terminus

prior to the stop codon TAG (Green arrow) that is also observed in the sequence

derived for the MceK fusion peptide. However, the fragment following the stop

codon also resembles and aligns with the serine rich C terminal of the classIIb

bacteriocins including the serine 84 (Red arrow) region where the DHBS

molecule is docked (Mercado et al., 2008). ....................................................... 156

Figure 4.8 Genome maps of the biosynthetic gene clusters involved in

production, export and immunity of peptides, MccE492 (mceA), MccG492

(mceL), MccH47 (mchB), MccM (mcmA). Peptides that are of the same colour

are homologues to each other [adapted from (Vassiliadis et al., 2010)]............ 157

Figure 4.9 Post translational modification pathway of MccE492 with the DHBS

siderophore moiety. 1) gycolysation of enterobactin by MceC. 2) cleavage of

enterobactin to achieve linear Glc-DHBS3, 3) attachment of the linear Glc-

DHBS3 and derivatives to the MccE492 precursor peptide (MceA) [from

(Vassiliadis et al., 2007)]. .................................................................................. 161

Figure 5.1 Graphical result representation of BAGEL3 software analysis. The

gene cluster belongs to lead 4 (see below), a putative class II lantibiotic

identified during this study and is surrounded by important motifs. LanK and

LanR are known to be involved regulation of class II bacteriocins, the ABC

Page 16: Discovery of antimicrobial peptides active against

List of figures

16 | P a g e

transporter is responsible for secretion) and SacCD encodes the SAM enzyme

that introduces sulphur bridges. ......................................................................... 168

Figure 5.2 A) Amino acid sequence alignments of lead 3 with two component

members of the class II lanthipeptide family, including lichenicidin A2 (orf023).

B) The biosynthetic cluster of lead 3 as predicted by BAGEL 3 software. ....... 176

Figure 5.3 A) Lead 12 sequence alignment shows conserved regions with

ruminococcicin, lacticin 481 and salivaricin 9. B) The putative bacteriocin is

accompanied by LanM and LanC modification components, multiple ABC

transporters and two bacteriocin related genes GerE and HisKA 3 with unknown

functions, as predicated by BAGEL 3................................................................ 177

Figure 5.4 A) the amino acid sequence alignment of lead 14 with circular

bacteriocins Enterocin AS-48, Leucocyclicin Q and Circularin A. B) The

anticipated biosynthetic gene cluster consists of two modification genes, HttB

and HttC and an ABC transporter. ..................................................................... 179

Figure 5.5 Unique bacteriocin sequences predicted by BAGEL 3 software. A)

Lead 4 is a predicted sactipeptide, B) lead 7 is a predicted LAPs peptide, C) lead

19 is a putative bottromycin and D) lead 9 is a likely class II lanthipeptide family

bacteriocin. ......................................................................................................... 180

Figure 5.6 Digested NI04 structural gene insert (left) and pET28a vector DNA

products (right) prior to gel purification. ........................................................... 182

Figure 5.7 Gel image displaying the comparison of colony PCR amplified pmnA

bands from the clones (CL) to those amplified from the producer. Positive

control (+) was insert DNA amplified from producer B. pumilus J1 DNA and the

negative control (-) was amplification from pET28a plasmid DNA with no insert

Page 17: Discovery of antimicrobial peptides active against

List of figures

17 | P a g e

(the template vector is visible in the negative control lane). Expected product

length was 319bp. .............................................................................................. 183

Figure 5.8 Insert orientation was confirmed using pmnA gene forward primers

and the T7 promoter region reverse primers. The expected product length was

435bp. ................................................................................................................. 184

Figure 5.9 CL1 and CL2 insert region sequence alignments with the expected

pET28a plasmid T7 region containing the pumicin NI04 gene pmnA. The

alignments showed that while CL2 encoded the correct sequence, the CL1 insert

had a mutation in the 3’ end and possibly encodes a non-functional peptide .... 185

Figure 5.10 Reverse purification of the in vitro expressed pumicin NI04

indicated with red arrows and positive control DHFR indicated with black

arrows, (1) Ni-NTA purified DHFR, (2) Ni NTA purified pumicin NI04, (3)

DHFR 100kDa ultrafiltration flow through, (4) pumicin NI04 100kDa

ultrafiltration flow through, (5) DHFR whole reaction protein, (6) pumicin NI04

whole reaction protein. ...................................................................................... 187

Figure 5.11 Enterocin AS-48 and cicularin A gene cluster arrangements [from

(van Belkum et al., 2011)] ................................................................................. 194

Page 18: Discovery of antimicrobial peptides active against

List of tables

18 | P a g e

LIST OF TABLES

Table 1.1 A list of recently introduced systemic antimicrobials with novel modes

of action [adapted from (Butler et al., 2013; Leach et al., 2011)]. ...................... 31

Table 2.1 A summary of the characterisation assays performed on each agent

identified during proteomic analysis .................................................................... 95

Table 2.2 A list of primers employed during the study. Blue nucleotides indicate

the restriction enzyme cut site, while red nucleotides highlight the start and stop

codons. ............................................................................................................... 103

Table 2.3 The content of the PCR reaction mixture used for amplification of the

inserts for use in construction of plasmid vector. .............................................. 104

Table 3.1 Antimicrobial activity of peptides NI03 and NI04 against indicator

species. Results were obtained using the spot on lawn assay with Sep-Pak C18

fractions. Activity was recorded as (+) if a visible inhibition zone was present

and (-) if no zone of inhibition was present. ...................................................... 114

Table 3.2 Effect of various media and additives on the production and

availability of peptide NI04, following overnight incubation at 37oC. .............. 115

Table 3.3 The unique sequence tags for peptide NI04 obtained from the de novo

peptide sequencing efforts. Common modifications that can occur were also

accounted............................................................................................................ 118

Table 3.4 Effect of a wide array of hydrolytic enzymes on the antimicrobial

activity of pumicin NI04. ................................................................................... 126

Table 3.5 Minimum inhibitory concentration (MIC) of pumicin NI04 against a

wide selection of Gram positive bacteria. MIC was calculated in AU/ml and

µg/ml in a selection of indicators using highly purified peptide........................ 127

Page 19: Discovery of antimicrobial peptides active against

List of tables

19 | P a g e

Table 3.6 Percentage of haemolysis that occurred following the addition of

pumicin NI04 to mouse erythrocytes. Triton-X100 detergent was used to achieve

absolute haemolysis and was used as the positive control. ................................ 130

Table 4.1 Peptide NI05 spectrum of activity determined through simultaneous

antagonism assays .............................................................................................. 147

Table 4.2 Effect of various media and additives on the production and

availability of peptide NI05, following overnight incubation at 37oC. .............. 148

Table 5.1 The short list of promising bacteriocin leads identified in the genomes

of 35 species of anaerobic bacteria of the genera Clostridium and

Propionibacterium and the producer bacteria K. pneumoniae A7 and B. pumilus

J1. Unique leads are novel leads that do not share homology with known

bacteriocins but are located with bacteriocin related genes. .............................. 172

Page 20: Discovery of antimicrobial peptides active against

List of abbreviations

20 | P a g e

LIST OF ABBREVIATIONS

ACN - Acetonitrile

AMP - Antimicrobial peptide

AMR - Antimicrobial resistance

AU - Arbitrary unit

BHI - Brain heart infusion (broth)

CAB - Columbia agar base

CA-MHB - Cation adjusted Muller

Hinton broth

CBA - Columbia blood agar

CLSI - Clinical and Laboratory

Standards Institute

CV - Column volume

FT - Flow through

HTS - High throughput screening

HVS - High vaginal swab

ESBL - Extended spectrum β-

lactamase

ESI - Electrospray ionisation

HPLC - High pressure liquid

chromatography

KPC - Klebsiella pneumoniae

carbapenemase

LAB - Lactic acid bacteria

LAP - Linear azol(in)e-containing

peptide

LC - Liquid chromatography

MALDI-TOF - Matrix assisted laser

desorption/ionisation time of flight

Mcc - Microcin

MIC - Minimum inhibitory

concentration

MRSA - Meticillin resistant

Staphylococcus aureus

MS - Mass spectrometry

MS/MS - Tandem mass spectrometry

NB - Nutrient broth

NDM-1 - New Delhi Metalloprotease

1

NP - Natural product

NRP - Non ribosomal peptide

NRPS – Non ribosomal peptide

Synthase

NMR - Nuclear Magnetic Resonance

OM - Outer membrane

PBS - Phosphate buffered saline

PK - Polyketide

PKS - Polyketide synthase

Page 21: Discovery of antimicrobial peptides active against

List of abbreviations

21 | P a g e

PTM - Post translational modification

RiPP - Ribosomally synthesised and

post translationally modified peptide

RP-HPLC - Reverse phased high

pressure liquid chromatography

SDS-PAGE - Sodium dodecyl

sulphate polyacrylamide gel

electrophoresis

TFA - Trifluoroacetic acid

TSB - Tryptic soy broth

WT - Wash through

YE - Yeast extract

UV - Ultra violet

Page 22: Discovery of antimicrobial peptides active against

Abstract

22 | P a g e

Abstract The University of Manchester, 2015

For the degree of Doctor of Philosophy - Arif Felek Discovery of antimicrobial peptides active against antibiotic resistant bacterial

pathogens Rapid development of antimicrobial resistance (AMR) among bacteria,

combined with diminished new antibiotic discovery rates, is an increasing threat to human health. Bacterially derived antimicrobial peptides (AMP) hold excellent potential as potent novel therapeutics. This study embraces traditional natural AMP discovery methods and the newer in silico genome mining tool BAGEL 3 to facilitate identification of novel antimicrobial agents. The traditional screening efforts led to the discovery of two promising antimicrobial producer strains; Bacillus pumilus J1 producing two AMPs, peptides NI03 and NI04, and Klebsiella pneumoniae A7, which produced peptide NI05. In silico mining of the B. pumilus J1 and K. pneumoniae A7 genomes and those from under exploited anaerobic bacteria using BAGEL 3 yielded 18 putative bacteriocin structures that were associated with multiple known and relevant bacteriocin accessory genes and/or carried significant homologies to known bacteriocins. Peptide NI04 proved to be active against Gram positive species only, including meticillin resistant Staphylococcus aureus and vancomycin resistant enterococci and peptide NI03, in addition to these pathogens, showed activity against E. coli. Peptide NI05 was active against Gram-negative pathogens including extended spectrum β-lactamase producing E. coli. All isolated peptides were observed to be proteinaceous in nature and highly heat stable.

Peptides were purified or partially purified using solid phase extraction followed by RP-HPLC. The mass of the peptides was determined using ESI or MALDI-TOF mass spectrometry. Tandem MS fragmentation of peptide NI04 generated several sequence tags. Draft genome sequences of the B. pumilus J1 and K. pneumoniae A7 producer strains were obtained using the Illumina MiSeq platform. This allowed identification of the genes encoding peptide NI04, which was confirmed to be novel and was named pumicin NI04. Further characterisation of pumicin NI04 demonstrated it was non-toxic to keratinocytes, Vero cells and non-haemolytic up to at least 18x the minimum inhibitory concentration. The discovery revealed that pumicin NI04 belongs to the WXG-100 peptide superfamily, having homology with the mycobacterial and staphylococcal virulence factor EsxA. This represents the first report of antimicrobial activity in a WXG-100 peptide and has intriguing evolutionary implications.

Although it was not possible to fully characterise peptides NI03 and NI05, when BAGEL 3 was used to mine the B. pumilus J1 genome, a promising putative bacteriocin candidate was identified that was homologous to Enterocin AS-45, which also confers anti Gram-negative activity and may be related to the activity observed for NI03, however more evidence is required. Investigations of the K. pneumoniae A7 bacteriocin on the other hand helped establish that the K. pneumoniae microcin E492 pathway was present and highly conserved in strain A7, and is likely to be responsible for the activity observed indicating that NI05 was not a novel peptide.

Page 23: Discovery of antimicrobial peptides active against

Declaration

23 | P a g e

DECLARATION

No portion of the work referred to in the thesis has been submitted in support of an

application for another degree or qualification of this or any other university or other

institute of learning.

Arif Felek

Page 24: Discovery of antimicrobial peptides active against

Copyright statement

24 | P a g e

COPYRIGHT STATEMENT

1. The author of this thesis (including any appendices and/or schedules to this thesis) owns certain copyright or related rights in it (the “Copyright”) and s/he has given The University of Manchester certain rights to use such Copyright, including for administrative purposes.

2. Copies of this thesis, either in full or in extracts and whether in hard or electronic copy, may be made only in accordance with the Copyright, Designs and Patents Act 1988 (as amended) and regulations issued under it or, where appropriate, in accordance with licensing agreements which the University has from time to time. This page must form part of any such copies made.

3. The ownership of certain Copyright, patents, designs, trade marks and other intellectual property (the “Intellectual Property”) and any reproductions of copyright works in the thesis, for example graphs and tables (“Reproductions”), which may be described in this thesis, may not be owned by the author and may be owned by third parties. Such Intellectual Property and Reproductions cannot and must not be made available for use without the prior written permission of the owner(s) of the relevant Intellectual Property and/or Reproductions.

4. Further information on the conditions under which disclosure, publication and commercialisation of this thesis, the Copyright and any Intellectual Property and/or Reproductions described in it may take place is available in the University IP Policy (see http://www.campus.manchester.ac.uk/medialibrary/policies/intellectual-property.pdf), in any relevant Thesis restriction declarations deposited in the University Library, The University Library’s regulations (see http://www.manchester.ac.uk/library/aboutus/regulations) and in The University’s policy on presentation of Theses.

Page 25: Discovery of antimicrobial peptides active against

Dedication

25 | P a g e

DEDICATION

I dedicate this thesis to my family;

To my wife Emine, for always being there for me with her endless love and

understanding, she is the light in my life. Also to my mother Ayten, father Osman and

brother Ismail for all of their support, sacrifices and unconditional love. I am very

lucky to have them in my life and without them none of this would have been

possible.

Page 26: Discovery of antimicrobial peptides active against

Acknowledgements

26 | P a g e

ACKNOWLEDGEMENTS

First of all I would like to express my deepest thanks to my

supervisor Dr. Mathew Upton, for all of his support and guidance. He was always

there to listen with all his invaluable advice, I could not have asked for more. I am

also ever so grateful to John Moat and Audrey Coke for their exceptional technical

help. The support they gave me went far beyond the laboratory. I am also indebted

to Dr. Catherine O’Neil for taking me on following the departure of Dr. Mathew

Upton from Manchester University. She was always very kind and understanding.

I would like to also thank Marjorie

Howard, Dr. Stacy Warwood and Dr. David Knight for training me on HPLC, Mass

spectrometry and de-novo peptide sequencing techniques and also express my

gratitude to Dr. Vikram Sharma for allowing me to use the proteomics facility at the

Plymouth University and Mathew Emery for his tireless support during my time in

Plymouth University.

I am also grateful to Majed Al-Ghoribi for his friendship and all of his

support during our PhD journey together. Finally I would like to thank everybody in

the Microbiology and Virology unit at Manchester University and the department of

Biomedical Sciences at Plymouth University. It has been an honour to be a part of

these teams and I look forward to continue working with everybody mentioned here

in the future.

Page 27: Discovery of antimicrobial peptides active against

Publications

27 | P a g e

PUBLICATIONS FROM THIS WORK

This study has produced the following published work, which includes

conference presentations (oral and poster) and manuscripts to be submitted

for publication.

Manuscript Felek A, Upton M. (2015) Pumicin NI04, a novel antimicrobial peptide from Bacillus pumilus, is homologous to the EsxA virulence determinant (in preparation) Oral and Poster Presentations Antibiotic alternatives for the new millennium, Eroscicon (November 2014) London UK, Peptide antibiotics for nasal decolonisation of MRSA carriage (Invited oral presentation). Society for General Microbiology Annual Conference (April 2014) Liverpool UK, Discovery of antimicrobial peptides active against antibiotic resistant bacterial pathogens (poster presentation) North West Microbiology Group Meeting (September 2013) Liverpool UK Developing Inhibitors of Antimicrobial Resistant Gram Negative Bacteria. (Poster presentation) Society for General Microbiology Spring Conference (March 2013) Manchester UK Discovery of Antimicrobial peptides active against Gram Negative Pathogens. (Poster presentation) Society for General Microbiology Spring Conference (March 2012) Dublin- Ireland. Developing Inhibitors of Antimicrobial Resistant Gram Negative Bacteria. (Poster presentation)

Page 28: Discovery of antimicrobial peptides active against

Chapter 1

28 | P a g e

Introduction

1 CHAPTER 1

GENERAL INTRODUCTION

Page 29: Discovery of antimicrobial peptides active against

Chapter 1

29 | P a g e

Introduction

Antibiotics are undoubtedly one of the most essential building blocks of

modern health care. As well as treatment of infected patients, they also play a key

role in preventing post-operative infections in surgery patients, which in turn

decreases the due morbidity and mortality rates considerably (Rice, 2008). However,

since the introduction of the first antimicrobial compounds into medicine, bacterial

resistance has been an ever increasing problem in antibiotic therapy, developing and

spreading rapidly amongst pathogens shortly after introduction of virtually any

antibiotic (Figure 1.1) (Clatworthy et al., 2007; Davies and Davies, 2010).

Nevertheless, in the early days resistance was not deemed very important, due

largely to the fact that before 1987 and especially during the 1940s-1960s (the

‘Golden Era’ of antibiotic discovery) humanity lived through a period of

unprecedented antibiotics wealth with high discovery rates (Silver, 2011).

Figure 1.1 Timeline of antibiotic discovery versus outbreak of resistance among bacteria [from (Clatworthy et al., 2007)]

Together, when combined with the idea that infectious diseases could be

wiped out, this wealth resulted in the abuse of antibiotics through unnecessary usage,

only to fuel the increasing development of antibiotic resistance within the bacterial

community and to this day antibiotic overuse and misuse continues both in clinics

Page 30: Discovery of antimicrobial peptides active against

Chapter 1

30 | P a g e

Introduction

and in agriculture (Aarestrup, 2005; Besser, 2003; Laxminarayan et al., 2013; Lee et

al., 2014; Shallcross and Davies, 2014; Tripathi et al., 2012).

Although appropriate measures are currently being implemented in the form

of improved antibiotic stewardship, better infection control measures and rising

public awareness to prevent the spread of antibiotic resistance, the slow introduction

of novel antimicrobials into the clinic (Figure 1.2 and Table 1.1) makes us face a

dramatically decreasing number of treatment options against multidrug resistant

(MDR) pathogens (Boucher et al., 2013; Livermore, 2011; White, 2011). The

Infectious Disease Society of America have listed six of the most problematic MDR

bacteria, namely the ESKAPE group of pathogens which are Enterococcus faecium,

Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii,

Pseudomonas aeruginosa and Enterobacter species (Rice, 2008). Among these there

are pan-antibiotic resistant species that cannot be treated with any conventional

antibiotics (Telang et al., 2011; Walsh and Toleman, 2012). In addition to its clinical

impact, antimicrobial resistance also has a severe effect on the economy with a

worldwide estimated cost of 100 trillion US dollars by 2050 (O’Neill, 2014).

Figure 1.2 The amount of antibiotics being introduced into the clinic has been drastically decreasing in recent years [From (Boucher et al., 2013)]

Page 31: Discovery of antimicrobial peptides active against

Chapter 1

31 | P a g e

Introduction

Table 1.1 A list of recently introduced systemic antimicrobials with novel modes of action [adapted from (Butler et al., 2013; Leach et al., 2011)].

Class Example agent Date approved

Oxazolidinones Linezolid 2000

Lipopeptides Daptomycin 2003

Lipoglycopeptides Telavancin 2009

Bedaquiline Diarylquinoline 2012

Nevertheless, the positive effect of increasing awareness, implication of

programs such as the IDSA’s “10 × ’20 initiative” that encourages the scientific

community to develop and obtain regulatory approval for 10 systematically

deliverable novel antibiotics by 2020 and proposals to change the policy on

antimicrobials has already started improving the future outlook positively (Boucher

et al., 2013; Infectious Diseases Society of America, 2010; Spellberg et al., 2011).

This can be observed in the rising number of antibiotics going through clinical trials

in 2013 compared with 2011 (Figure 1.3) (Pucci and Bush, 2013).

Figure 1.3 Graphical representation and comparison of the number of antibiotic candidates going through clinical trials between 2011 and 2013 [from (Pucci and Bush, 2013)].

Page 32: Discovery of antimicrobial peptides active against

Chapter 1

32 | P a g e

Introduction

However, it has been never more clear than today that this is an arms race that

will be fought for years to come and thus science must come up with new lucrative

novel antimicrobial discovery platforms and sources. Hence, it is important to

discuss the reasons behind the lack of investment in novel antimicrobial discovery

and to explore innovative approaches that may bring the breakthrough that is

urgently required.

1.1 DOWNFALL OF NOVEL ANTIMICROBIAL DISCOVERY

It is important to understand the causes of the problem before seeking

solutions. Literature agrees in two fundamental changes that are believed to have led

to the lack of antimicrobial discovery, the first one is the failure of adopted genetic

approaches that were meant to replace natural product screening for antimicrobial

discovery (Lewis, 2013). The second reason, which is also directly related to the

first, is the movement of “big pharma” away from antimicrobial drug discovery due

to the high cost and low success of development; as a result currently the main

drivers of antimicrobial discovery are smaller pharmaceutical and biotechnology

companies.

1.1.1 Movement of “big pharma” away from antimicrobial drug discovery

The retreat of big pharma from the field can be summarised under two

sections. The first reason is the fact that when the discovery and development costs

and the difficult route to receiving regulatory approval for antibiotics are taken into

account, these agents generate significantly lower revenues compared with

therapeutics for chronic diseases such as cholesterol reduction, diabetes and cancer

treatment (Livermore, 2011; Projan, 2003). This is mainly due to their short

treatment course and the fact that to prevent the rapid development of resistance, the

Page 33: Discovery of antimicrobial peptides active against

Chapter 1

33 | P a g e

Introduction

new antibiotics would be kept as a last resort, affecting initial revenues meaning

patents expire before costs are recovered (Bérdy, 2012; Katz et al., 2006; Livermore,

2011; Projan, 2003).

The second reason is the fact that value of the drug will drop rapidly once

resistant strains emerge and adapting to new resistance mechanisms is proving very

hard due to the unpredictability of mutations (Livermore, 2011; Livermore et al.,

2011). For example, the emergence of New Delhi metalloprotease-1(NDM-1) has

rendered both the β-lactams and β-lactamase inhibitors such as clavulanic acid less

valuable (Kumarasamy et al., 2010; Nordmann, 2014) and vancomycin resistance

had the same effect although less dramatic on vancomycin (De Vriese and

Vandecasteele, 2014; Harbarth et al., 2002; Livermore, 2007; Meziane-Cherif et al.,

2014).

However, investing only in countermeasures for NDM-1 isn’t enough. This

is due to the fact that resistance against β-lactams and β-lactamase inhibitors can

develop through more than one mechanism. One of the most important is the rapidly

spreading Klebsiella pneumoniae carbapenemase (KPC) (Nordmann, 2014; Robilotti

and Deresinski, 2014). KPC is also active against β-lactams as NDM-1, yet

structurally it is different, demanding individualised research (Nordmann, 2014).

When there are multiple targets, adaptation of counter measures against them is even

harder and more expensive (Livermore et al., 2011). However, steps are

implemented to improve the economic positon of pharmaceutical companies to

incentivise antibiotic development through legislation. GAIN (Generating Antibiotic

Incentives Now) act is one attempt approved by the US government, it increases

exclusivity of antibiotics and requires the food and drug administration to provide

Page 34: Discovery of antimicrobial peptides active against

Chapter 1

34 | P a g e

Introduction

updated clinical trial guidance and list pathogens that pose a relatively higher threat

to public health to provide a clearer path (US Congress, 2011).

1.1.2 Overview of target based genetic approaches and their failure to

discover novel antibiotics

In 1995, the genome sequence of Haemophilus influenzae was published and

this was quickly followed by many others; today there are 45,076 bacterial genomes

available in the joint genome institute (JGI) genomes online database (GOLD)

(https://gold.jgi-psf.org/, last access: 30/06/15). These advances were rapidly

followed by development of powerful bioinformatic tools that enable rapid

comparative analysis of gene libraries and soon the potential

of these resources, with regards to increasing antimicrobial discovery, was

realised (Bansal, 2005; Livermore, 2011). This has led to investigators abandoning

natural product screening methods that did not fit in with the prospect of rapid

antimicrobial discovery through utilisation of the rapid nature of new techniques to

establish target based high throughput screening (HTS) platforms (Lewis, 2013;

Livermore, 2011; Silver, 2011). The HTS methods, which use synthetic product

libraries, were quickly adopted by most drug companies. The approach is outlined in

Figure 1.4.

Figure 1.4 The process of target based antimicrobial drug discovery [adapted from (Mills, 2006; Pucci, 2007)].

Target identification

Target validation

Screening: isolated enzyme

Screening: whole cell

analysis

Hit & lead development

Clinical development

& Drug

Page 35: Discovery of antimicrobial peptides active against

Chapter 1

35 | P a g e

Introduction

Although powerful and much faster than natural product screening, the target

based genetic approaches have failed to produce any antimicrobials worthy of

introduction into clinical practice. The literature points to two main issues that limit

the target based methods. The first one is that many target based, high throughput

screening efforts are conducted in silico. This creates complications with

accessibility of the target site by selected compound, as this is not predictable

(Livermore, 2011). Although to overcome this issue, physical screening studies

targeting whole cell growth inhibition are performed, the exclusive nature of the

synthetic compound libraries used, which only or predominantly include compounds

that fit into Lipinski’s rule of five becomes a limiting factor (Lewis, 2013;

Livermore, 2011).A number of synthetic compounds that were discovered using the

whole cell screening approach also failed to make it through lead development due

to non-specific membrane activity that caused lysis of non-target cells including

erythrocytes during early analysis (Payne et al., 2007, 2004).

Lipinski’s rule of five is an algorithm that is used to identify possible drug

compounds according to their solubility and permeability within the body. In a

paper, Lipinski et al. (1997) state that the parameters which should be used to asses

suitability of a compound as a drug candidate is; molecular weight (less than 500Da),

Log P (not greater than 5), the number of H bond donors (not more than 5) and H

bond acceptors (not more than 10). However, as also noted by Lipinski et al. 2001,

some antibiotics lie outside the rule of five (Lipinski et al., 2001), some antibiotics

such as aminoglycosides and β-lactams are larger and more hydrophilic and thus

outside the parameters (Payne et al., 2007). This information alone identifies these

libraries as a limiting factor. The synthetic product libraries are designed for

physiological conditions in the human body, not for interacting with bacteria. As

Page 36: Discovery of antimicrobial peptides active against

Chapter 1

36 | P a g e

Introduction

bacteria come into contact and interact with small compounds constantly, many of

which could be harmful, they have developed coping mechanisms such as efflux

pumps to regulate entry of these compounds (Delcour, 2009; Wong et al., 2014;

Wright, 2014).

1.2 REVITALISING NATURAL PRODUCT SCREENING AS A SOURCE OF

NOVEL ANTIBIOTICS

Failure of the target based HTS approach has caused researchers to explore

other platforms of antibiotic discovery and the focus has been shifted back to the

natural product screening also named the “Waksman platform” (Lewis, 2013).

Although this shift is influenced by the advancements of methods employed for

natural product screening, it’s also due to the inherent advantages of the natural

products summarised in Figure 1.5 (Lewis, 2013).

NP

Effective killing action

Novel target discovery

Structural diversity

Rational modification

Untapped sources

Figure 1.5 Advantages of natural products (NP) as antibiotic drug leads

Page 37: Discovery of antimicrobial peptides active against

Chapter 1

37 | P a g e

Introduction

Natural antimicrobial products, unlike those found in synthetic product

libraries, have evolved with their host, among other reasons, for one particular job, to

infiltrate or interact with the target bacterial cell and kill it (Wright, 2014). Hence,

natural products suffer proportionally less from issues such as reduced cell

permeability, unlike many synthetic products, and in some cases readily contain the

properties required from an antibiotic, meaning that they will need less lead

optimisation (Lam, 2007). In addition, natural products have already been

successfully implemented in the clinic and the majority of antibiotics are derived

from these products (Wright, 2014). Thus, on many occasions in current literature

natural antimicrobials are referred to as “privileged compounds” with regards to their

potential as antibiotic candidates (Tan et al., 2015; Wright, 2014).

In addition, natural products cover a unique and large chemical space with

unmatched structural diversity. To give a few examples, these include stereogenic

centers such as chiral centers, heterocyclic substituents and polycyclic structures

(Dobson, 2004; Genilloud, 2014; Rosén et al., 2009; Szychowski et al., 2014). This

increases the value of natural product discovery two fold as these unique structures

can provide examples for combinatorial chemistry and create a vision for templates

to be used to modify and improve the activity of known and used compounds (e.g.

tigecycline from tetracycline) (Lam, 2007) or bring out the hidden potential in

previously dismissed compounds such as Linezolid (Newman, 2008), earning them

back for medicine.

Page 38: Discovery of antimicrobial peptides active against

Chapter 1

38 | P a g e

Introduction

The peptidomimetics approach is another creation of natural product inspired

synthetic chemistry. This involves replacement or rearrangement of the original

structure of an antimicrobial peptide (AMP) with a mimic to generate libraries with

increased activity or decreased toxicity (Morrison and Hergenrother, 2014). There

are many strategies to achieving mimics, one of which is replacing the α-amino acid

backbone with a β-amino acid containing mimic to make the antimicrobial more

stable against proteases while reducing the toxicity (Godballe et al., 2011; Johnson

and Gellman, 2013). Another strategy is to transfer the active region of the

antimicrobial to a backbone nitrogen on an N-substituted glycine (peptoid) molecule,

which in turn can also potentially decrease the toxicity and increase the proteolytic

stability (Godballe et al., 2011; Miller et al., 1994; Rotem and Mor, 2009;

Zuckermann et al., 1992). It is also possible to manipulate certain compounds to

mimic and improve upon the desirable properties of a known antimicrobial, such a

study conducted by Tan et al. (2015) which had successfully managed to create a

antimicrobially active library of pramanicin mimics derived from pyroglutamic acid

(Tan et al., 2015). It is also important to note here that, the value of natural products

is not only limited to their potential as drug leads but they may also play a crucial

part in discovery or understanding of novel drug targets, pathogenicity mechanisms

and other bacterial processes (Lam, 2007).

Nonetheless, natural product screening was largely abandoned. Some of the

disadvantages associated with this move can be attributed to the slow pace of

discovery associated with natural product research, concerns about large scale

production and patenting. However, rediscovery of existing compounds is often

accepted as the main reason (Harvey, 2008; Livermore, 2011; Newman, 2008).

Page 39: Discovery of antimicrobial peptides active against

Chapter 1

39 | P a g e

Introduction

However, recent estimates show that only 10% of the potential natural

products have been discovered from the known producer species and most bacterial

species are not yet cultivated in laboratories (Curtis et al., 2006; Walsh and

Fischbach, 2010). This means that there is still a plethora of possibilities waiting to

be discovered contrary to the prevailing belief in 1990s that natural products had

reached their limits, and the recent developments in natural product screening

methods are very promising. Advances in cultivation techniques such as the isolation

chip (iChip), a high throughput diffusion chamber system that allows culture of

many non-cultivable bacteria using traditional cultivation methods (Ling et al., 2015;

Nichols et al., 2010), and developments in analytical technologies (see Section 1.11)

have removed most of the difficulties associated with the older methods that caused

researchers to abandon natural product discovery and are encouraging the return to

natural products. Some significant antimicrobial peptide classes have been reviewed

in the following section.

1.3 BACTERIALLY DERIVED NATURAL PRODUCTS OF INTEREST

The antagonistic activity of some bacterial species against each other was

first recorded by Pasteur and Joubert in 1877 as they observed “common bacteria”

(likely E. coli) adversely effecting the growth of Bacillus anthracis. After this

discovery, colicin V one of the oldest known bacteriocins (currently classified as a

microcin, see below), was isolated from E. coli strain V and others (Cascales et al.,

2007; Davies et al., 1981). This was followed by many others and initially all were

named colicins but further research showed that many other bacteria were producing

colicin like substances and some of these substances were fundamentally very

different from each other, either in their production pathways or structurally (Jack et

Page 40: Discovery of antimicrobial peptides active against

Chapter 1

40 | P a g e

Introduction

al., 1995). Thus, many divisions had emerged for the classification of these

antimicrobially active natural products. Some of the most promising classes that this

project focuses on are listed in Figure 1.6 below.

Figure 1.6 Natural products of interest in the current project. The figure above lists the natural products produced by bacteria that have the potential to be used as antibiotics. These products are shortly reviewed in this study.

Antimicrobials of interest

Lysins Ribosomal peptides

Bacteriocins

Bacteriocins of Gram positive

bacteria

Class I Lantibiotics

Class II un-modified

Class III Large proteins

Clsass IV cyclic peptides

Bacteriocins of Gram negative

bacteria

Colicins

Microcins

Miscellaneous

Non-ribosomal peptides Polyketides

Modular polyketides

Aromatic polyketides

Page 41: Discovery of antimicrobial peptides active against

Chapter 1

41 | P a g e

Introduction

1.4 BACTERIOCINS

Bacteriocins are ribosomally produced AMPs, to which the producing

bacteria are immune (Cascales et al., 2007; Jack et al., 1995). It is suggested that

these peptides are produced to give an edge to the producing bacteria against other

competing bacterial species in the environment, as these agents generally target

closely related species (Cascales et al., 2007; Jack et al., 1995). However, their

activity ranges from broad spectrum agents that have interspecies activity, to very

narrow spectrum agents that are only active against other strains of the same species

[e.g. Cerein isolated from Bacillus cereus is observed to be only active against other

species of B. cereus] (Naclerio et al., 1993).

Although the term bacteriocin did not exist at the time of their discovery,

colicins are the first recorded bacteriocins to be discovered. At the time, a list of key

properties were created to identify colicins and a classification scheme was produced

according to the entry route of the agent into the cell (Cascales et al., 2007).

However, as research uncovered numerous new colicin like substances with

substantial structural differences from each other both from coliform and non-

coliform bacteria, the colicin based identification system was no longer applicable

and the need for a new classification scheme arose (Jack et al., 1995).

The first detailed classification system was proposed by Klaenhammer et al.

(1993) that separated bacteriocins into four groups according to their structural

components (Klaenhammer, 1993). This scheme is currently accepted and used by

many scientists, however, updated classification schemes by Kemperman et al.

(2003) and Cotter et al. (2005) have also been proposed and widely accepted

creating some confusion. In response, Heng and Tagg (2006) have produced a

Page 42: Discovery of antimicrobial peptides active against

Chapter 1

42 | P a g e

Introduction

universal scheme that combines the Klaenhammer et al (1993) and Cotter et al

(2005) classifications in an optimal scheme with some added updates to create a

proposal that encompassed the majority of the known bacteriocins at the time (Figure

1.7) (Heng and Tagg, 2006). It is important to note here that a rift exists between

classification of Gram positive and negative bacterial bacteriocins and the

aforementioned classifications schemes, with the exception of the Heng and Tagg

(2006) schema, are useful for differentiation of bacteriocins derived from Gram

positive bacteria. However, Kempermen et al. (2003) did use the Gram-negative

bacteriocin microcin J25 as an exemplar for their proposal.

Yet, with each new bacteriocin discovered the number of peptides with

additional unique structures continue to accumulate and as a result additional groups

that had not been covered by any of the previously described bacteriocin schemes

Figure 1.7 The "Universal” bacteriocin classification scheme. This figure by Heng and Tagg, (2006) illustrates the two major classification schemes that are used and how they propose to combine the respective advantages to construct a universal scheme.

Page 43: Discovery of antimicrobial peptides active against

Chapter 1

43 | P a g e

Introduction

have arisen, such as the sactibiotic and lasso peptide families (Arnison et al., 2013;

Mathur et al., 2014). Indeed, with new discoveries the existing classes such as

lantibiotics, within these schemes have also become even more varied and the

aforementioned bacteriocin classification schemes can’t justifiably describe the

groups (Arnison et al., 2013). What follows is an introduction of common

bacteriocin classes and the general titles of the universal Heng and Tagg scheme will

be retained. However, the groups that fall under the Ribosomally synthesised and

post-translationally-modified peptides (RiPPs) will be supplemented with the

recommendations form the universal nomenclature study for the greater class of

RiPPs by Arnison and colleagues (Arnison et al., 2013). Based on the same study, an

independent miscellaneous peptide family has been added to address the newly

arising RiPP bacteriocin groups that share conserved properties that differentiate

them from existing classes and justify their existence as a separate class; these

peptides may and do span Gram positive and negative producers.

1.5 BACTERIOCINS OF GRAM POSITIVE BACTERIA

1.5.1 Class I - Lanthionine containing bacteriocins (Lantibiotics)

These bacteriocins contain multiple ring structures that are linked to each

other by lanthionine (Lan) or 3-methyllanthionine (MeLan) bonds between

dehydrated residues. They widely contain dehydrated amino acids and undergo

extensive post-translational modifications (eg. Nisin) (Fujita et al., 2007; McAuliffe

et al., 2001). These antimicrobial peptides gather particular interest from researchers

for they are commonly produced by food grade lactic acid bacteria (LAB).

Moreover, some lantibiotics, like nisin have dual actions on the target cell, a feature

not seen in conventional antimicrobials that contributes to increased stability against

Page 44: Discovery of antimicrobial peptides active against

Chapter 1

44 | P a g e

Introduction

antimicrobial resistance (Figure 1.8) (Cotter et al., 2005; Mantovani and Russell,

2001). Nisin acts by either inhibiting cell wall synthesis through blocking the lipid II

transporter that is responsible for the transfer of peptidoglycan subunits from the

cytoplasm, or it exploits lipid II to insert itself into the cell membrane and form

pores that cause lysis leading to cell death (Cotter et al. 2005). Nisin, isolated in

1928 (Rogers, 1928), is one of the most well studied lantibiotics and also it is one of

the most widely used preservatives in the food industry. However, resistance to nisin

is extremely rare compared with conventional antibiotics (Mantovani and Russell,

2001).

Figure 1.8 Bacteriocin mode of action. The class II bacteriocins such as sakacin can insert into the cell membrane by interacting with membrane proteins and cause pore formation and depolarisation of the cell membrane. The class I bacteriocins on the other hand can have a single or multiple modes of action such as nisin. Large protein bacteriocins such as lysostaphin can also act on the cell wall by hydrolysing the interpeptide bridges of the peptidogycan layer [adapted from (Cotter et al. 2005)].

Nisin Sakacin LysostaphiLysostaphin Nisins Sakacin

Page 45: Discovery of antimicrobial peptides active against

Chapter 1

45 | P a g e

Introduction

Many bacteriocins like nisin act by forming pores in bacterial cell

membranes however it’s important to note that some bacteriocins may also act

intracellularly. Although these mechanisms are not as well defined some these

modes of action include, stabilisation of DNA gyrase (microcins B17) (Collin et al.,

2013; Thompson et al., 2014) and cleavage of RNA (carocin S2) (Chan et al.,

2011) and ribosomal RNA (colicin E2) (Cavera et al., 2015; Ng et al., 2010).

Lantibiotics in the Heng and Tagg (2006) scheme were divided into 3 sub-

groups, type A, B and C. Type A lantibiotics contain elongated structures and a

cationic charge. These peptides act by forming pores on the bacterial cell wall, while

type B lantibiotics are globular peptides and act by preventing the execution of vital

cell processes such as cell wall synthesis, through inhibition of key enzymes (Cotter

et al., 2005; McAuliffe et al., 2001; Nishie et al., 2012). Subtype C was added later

to account for the multi component lantibiotics (Heng and Tagg, 2006; Snyder and

Worobo, 2014).

However, the growing number of lantibiotic structures, biosynthetic

pathways and the discovery of non-antimicrobial lanthionin containing peptides, for

instance SapB (Kodani et al., 2004) and SapT (Kodani et al., 2005) [both involved in

aerial mycelium formation of streptomycetes], has strained the classification scheme

and the broader term “lanthipeptide” was adopted to encompass these non-

antimicrobials (Arnison et al., 2013; Goto et al., 2010). Lantibiotics became a large

subgroup of lanthipeptides of the RiPP family and a new classification scheme has

been created that divides the lanthipeptides into four groups according to

components involved in their maturation (Knerr and van der Donk, 2012) (Figure

Page 46: Discovery of antimicrobial peptides active against

Chapter 1

46 | P a g e

Introduction

1.9). It is the belief of this study that application of this classification to categorise

lantibiotics is the more efficient approach.

The classification of these bacteriocins that is proposed as part of the current

study, as also agreed by the nomenclature study performed by Arnison et al. (2013),

is as follows:

1.5.1.1 Lanthipeptide class I

This class is characterised by the presence of two individual enzymes LanB and

LanC that are involved in the production of its members. LanB is a dehydrogenase

enzyme that is responsible for dehydration of Ser and Thr residues (Karakas Sen et

al., 1999; Kluskens et al., 2005; Koponen et al., 2002) while LanC is a cyclase that

facilitates the circularisation of the dehydrated residues through intramolecular

cysteine bonds (Koponen et al., 2002; Li et al., 2006; Lubelski et al., 2008). This

class contains members of the previously named type A lantibiotics such as Nisin

(Rogers, 1928), subtilin (Banerjee and Hansen, 1988) and epidermin (Allgaier et al.,

1985).

Figure 1.9 Four classes of the lanthipeptide family can be separated by the modification enzymes involved in generation of their lanthionine bridges. Solid lines indicate the conserved regions thought to play roles in the catalysis including zinc ligands that are missing in class III [from (Knerr and van der Donk, 2012)].

Page 47: Discovery of antimicrobial peptides active against

Chapter 1

47 | P a g e

Introduction

1.5.1.2 Lanthipeptide class II

Class II contains the lanthipeptides that are dehydrated and cyclised by a

bifunctional LanM enzyme that consists of two domains; a dehydratase and a LanC-

like cyclase domain (Arnison et al., 2013; Begley et al., 2009; Knerr and van der

Donk, 2012; Singh and Sareen, 2014). Many peptides in this class were classified as

type B lantibiotics, actagardine (previously gardimycin) (Boakes et al., 2009) and

mersacidin (Bierbaum et al., 1995) are some examples. Class II also contains the two

component lantibiotics such as haloduracin (Lawton et al., 2007; McClerren et al.,

2006) and lacticin 3147 (Martin et al., 2004; McAuliffe et al., 1998; Ryan et al.,

1999).

1.5.1.3 Lanthipeptide class III

These lanthipeptides are modified with a tri-domain LanKC modification

enzyme that incorporates an N-terminal lyase domain, a central kinase domain to

dehydrate the Ser and Thr residues and a C-terminal putative cyclase that differ from

the cyclases of other classes as it lacks the three metal binding domains conserved in

the cyclases of these classes (Krawczyk et al., 2012; Müller et al., 2010). Some

examples of lantibiotics that belong to this class are erythreapeptin, avermipeptin,

griseopeptin (Völler et al., 2012) and curvopeptin (Krawczyk et al., 2012).

1.5.1.4 Lanthipeptide class IV

Similar to the class III lanthipeptides, class IV also employs a multicomponent

enzyme with a lyase and a kinase domain, however unlike Class III, LanL contains a

LanC like C-terminal domain (Arnison et al., 2013; Goto et al., 2011, 2010). To the

best of our knowledge there is only one putative lantibiotic, venezuelin, identified for

this class (Goto et al., 2010).

Page 48: Discovery of antimicrobial peptides active against

Chapter 1

48 | P a g e

Introduction

1.5.2 Class II non-lanthionine containing bacteriocins

These are small bacteriocins that differ from lantibiotics by the fact that they are

not subjected to post translational modifications, except disulfide bridges and they

exhibit different modes of action (Figure 1.8) (Cotter et al., 2005; Kjos et al., 2011;

Nissen-Meyer et al., 2009). Class II bacteriocins are divided further into groups A, B

and C (Heng and Tagg, 2006). Here a slight change to the scheme is suggested and

class II two peptide bacteriocins are placed in group-B and the miscellaneous class II

bacteriocins in group-C, as this is the accepted nomenclature in current literature:

1.5.2.1 Class II-A pediocin-like bacteriocins

These bacteriocins are renowned for their effectiveness against Listeria

species. They contain a highly conserved YGNG(V/L) (“pediocin box”) amino acid

string encoded on their N-terminus. This consensus can be further extended to

YGNG(V/L)xCxxxxCxVxWxxA (Eijsink et al., 1998; Héchard and Sahl, 2002; Kjos

et al., 2011). The first member of the family discovered was the pediocin PA-1

(Gonzalez and Kunka, 1987) and thus they are often defined as pediocin like

substances. A study performed on 50 representatives of this class has shown that

they could be further divided into eight subgroups according to their primary and 3D

structures and mode of action (Cui et al., 2012). Some prominent examples of this

group include Enterocin P (Cintas et al., 1997), Bavaricin MN (Kaiser and

Montville, 1996) and divercin V41(Bhugaloo-Vial et al., 1999).

A connection between the mode of action of Class II-A bacteriocins and the

mannose phosphotransferase system (man-PTS) was suspected (Gravesen et al.,

2002; Ramnath et al., 2000), however a study using the Lactococcin A has proven

these suspicions through deletion of man-PTS components which confirms that the

Page 49: Discovery of antimicrobial peptides active against

Chapter 1

49 | P a g e

Introduction

action of Lactococcin A, and likely other pediocin like peptides, is dependent on the

presence of IID and IIC transmembrane proteins of the Man-PTS (Figure 1.10) (Diep

et al., 2007; Kjos et al., 2011).

1.5.2.2 Class II-B two component bacteriocins

These bacteriocins are composed of two peptides that work together to

achieve lethal activity. Although occasionally some antimicrobial effect is observed

from one of the components only, the optimal activity is achieved when each peptide

is present and components from different pairs don’t complement each other

(Garneau et al., 2002; Oppegard et al., 2007). This was observed very clearly when

Anderssen et al. (1998) tested Plantaricins EF and JK and found that both pairs were

103 times more active than individual components and cross combinations of

subunits of each pair didn’t yield any complementary activity (Anderssen et al.,

1998; Oppegard et al., 2007). Genetic evidence also suggests that Class II-B

bacteriocins evolved to act as one unit as the genes encoding both subunits are

located close to each other and only one immunity gene is present for each pair, such

as is seen with lactocin 705 and Enterocin 1071 (Cuozzo et al., 2000; Franz et al.,

2002; Oppegard et al., 2007).

Class II a bacteriocin

Figure 1.10 Class II-A bacteriocin action is dependent on the presence of IID and IIC proteins, which are membrane proteins of the Man-PTS mechanism. Class II a bacteriocin first binds to the extracellular loop and then to the transmembrane helices of these peptides causing conformational changes that leads to efflux and cell death [from (Kjos et al., 2011)]

Page 50: Discovery of antimicrobial peptides active against

Chapter 1

50 | P a g e

Introduction

1.5.2.3 Class II-C Miscellaneous bacteriocins

This class is made up of linear class II bacteriocins that do not fit into any

other class. Epidermicin NI01 (Sandiford and Upton, 2012), LSEI 2163 (Kuo et al.,

2013), Lacticin Q (Fujita et al., 2007) and Lacticin Z (Iwatani et al., 2007) are some

of the best examples of this group. They contain no leader peptide or pediocin box

and don’t carry any post translational modifications (Cotter et al., 2005; Heng and

Tagg, 2006; Kjos et al., 2011; Nissen-Meyer et al., 2009).

1.5.3 Class III large proteins:

These bacteriocins are traditionally defined as structurally large and heat

labile molecules, however, Cotter et al. (2005) had proposed a modification to the

definition of Class III bacteriocins, defining them as non-bacteriocin lytic proteins

and suggested naming the class as bacteriolysins. However as Heng and Tagg (2006)

pointed out, many of these peptides including lysostaphin and zoocin A carry

bacteriocin like properties (Beatson et al., 1998; Thumm and Gotz, 1997). In

addition while some large bacteriocins (lysostaphin) kill through lysis (Figure 1.8,

page 44), some such as dysgalacticin don’t; this heat stable large peptide acts by

inhibiting glucose fermentation (Cotter et al., 2005; Joerger and Klaenhammer,

1986; Swe et al., 2009). Thus, instead of using the term bacteriolysins, Heng and

Tagg (2006) suggested the name “large proteins” and divide the group into two;

bacteriolytic and non-bacteriolytic large peptides.

Page 51: Discovery of antimicrobial peptides active against

Chapter 1

51 | P a g e

Introduction

1.5.4 Class IV Cyclic peptides

These peptides are head to tail cyclysed (cyclisation occurs through bonding

between C to N termini, not by intramolecular cyclisation) to assume their distinctive

form (Arnison et al., 2013; Gabrielsen et al., 2014). All 10 members of this group

are exclusively produced by Gram positive bacteria and recent observations have

shown that they share common properties and can be divided into two sub-groups

(Gabrielsen et al., 2014; Martin-Visscher et al., 2009). The first group (Class I)

contains the enterocin AS-48 like peptides that have a high isoelectric point and are

positively charged peptides while preserving the overall cationic charge (Arnison et

al., 2013; Gabrielsen et al., 2014). Enterocin AS-48 is a potent antimicrobial peptide

(José et al., 2014) and it’s the best known member of the cyclic peptides. Other

examples of this group include circularin A (Kemperman et al., 2003a, 2003b) and

lactocyclicin Q (Sawa et al., 2009).

The second sub-group (Class II) is much smaller with only two affiliates

gassericin A (Kawai et al., 1998) and butyrivibriocin AR10 (Kalmokoff et al., 2003).

These peptides are anionic and tend to be neutral or positive in charge when under

physiological conditions and have a low isoelectric point (Arnison et al., 2013;

Gabrielsen et al., 2014). Also, it must be mentioned that some circular bacteriocins

contain sulphur to α-carbon bonds and are currently classified under the sactibiotic

group such as the subtilosin A (Gabrielsen et al., 2014; Kawulka et al., 2003) (see

Section 1.7.1).

Page 52: Discovery of antimicrobial peptides active against

Chapter 1

52 | P a g e

Introduction

1.6 BACTERIOCINS OF GRAM-NEGATIVE BACTERIA

Although the term bacteriocin is generally associated with AMP produced by

Gram positive bacteria, the term was devised to include AMPs derived from both

Gram positive and negative bacteria (Jack et al., 1995). Gram-negative bacteriocins

in general can be divided into two main categories, the colicins and the microcins

depending on the size of the AMP, and very much like their Gram positive relatives,

they carry the bacteriocin traits described in Section 1.4.

1.6.1 Colicins

Colicins can be distinguished from microcins primarily with their larger size

(̴ 25-80kDa). There are four known modes of action for colicins; these are (1) pore

formation in the cytoplasmic membrane [e.g. Colicin A (Braun et al., 1994)], (2)

DNA degeneration [Colicin E2, E7, E9 (Mora and de Zamaroczy, 2014)] (3)

inhibition of protein synthesis [Colicin DF 13, E3 (Gillor et al., 2004; Mora and de

Zamaroczy, 2014)] and (4) inhibition of lipopolysaccharide and murein synthesis

[Colicin M (Gillor et al., 2004)]. However, the colicin must first gain access to the

target.

To gain access to the target site, each colicin binds to a specific receptor on

the target bacteria that is involved in the uptake of various nutrients and they are

translocated into the cell using either the Tol or the TonB pathways (Figure 1.11)

(Cascales et al., 2007; Gillor et al., 2004).

Page 53: Discovery of antimicrobial peptides active against

Chapter 1

53 | P a g e

Introduction

The class of a colicin, is also dictated by the pathway it exploits to gain

access to the target cell. The colicins that use the Tol pathway are classified as Group

A colicins and utilise TolA, TolB, TolQ and TolR (Figure 1.11) (Lloubès et al.,

2012). Group A colicins are encoded by small plasmids and are secreted into the

media. Although colicin producing strains code for immunity proteins that confer

resistance to the producing bacteria, the production mechanism of group A colicins

is dependent on the SOS pathway and is accompanied by the production of a lysis

protein lethal to the producing cell (Cascales et al., 2007; Gillor et al., 2004). Thus,

group A colicin production may be defined as an altruistic process as the production

is initiated under depleting nutrient availability and the release results in death of the

producer and in turn that of nearby sensitive bacteria providing advantage to the

remaining cells of the producer population in the competition for nutrients (Cascales

et al., 2007). The colicins that gain access to the target using the ExbB–ExbD–TonB

pathway (Figure 1.11) are classified as Group B colicins (Lloubès et al., 2012). In

Figure 1.11 Mechanisms of cellular entry used by colicins. ExbB, ExbD components in the TonB pathway are used by group B colicins and the Tol pathway (TolA, TolB, TolQ and TolR) is employed by group A colicins [from (Lloubès et al., 2012)].

Page 54: Discovery of antimicrobial peptides active against

Chapter 1

54 | P a g e

Introduction

contrast to Group A, the colicins in Group B are encoded by large plasmids and

unlike Group A they are not actively secreted into the media from the producer

(Cascales et al., 2007). However, it should be noted that colicins from both groups

can have shared properties (Cascales et al., 2007; Pilsl and Braun, 1995).

1.6.2 Microcins

The term microcin (Mcc) was first proposed in 1975 by Asensio and Perez-

Diaz, after they had discovered 10 antimicrobials from a culture of human faecal

bacteria, all of which had a molecular weight less than "1000" (Asensio and Pérez-

Díaz, 1976). Although microcins and colicins are mainly distinguished in the basis

of weight, there are other key differences between them, most importantly microcin

production does not depend on the SOS pathway and is non-lethal to the producing

bacteria (Duquesne et al., 2007a). In addition, colicins are found encoded on

plasmids, whereas operons encoding microcis can be found on both plasmids and the

bacterial chromosome.

Unlike other bacteriocins, very few microcins have been identified so far and

even less have been structurally characterised. With the recent discovery of microcin

S by Zschuttig et al. (2012) only 19 microcins have been identified to date (Zschüttig

et al., 2012). These figures are based on literature searches and examination of

entries in the BactiBase database (http://bactibase.pfba-lab-tun.org/main.php) carried

out before 1st July 2015. All known microcins, except microcin C, are translated as

propeptides that consists of a leader sequence that is cleaved later in the maturation

process and a core peptide (Duquesne et al., 2007a; Rebuffat, 2012). Both the

propeptide and the core peptide may go through post translational modifications. As

a result, within this small collection, there is vast structural diversity and varied

Page 55: Discovery of antimicrobial peptides active against

Chapter 1

55 | P a g e

Introduction

modes of action (Figure 1.12 and Figure 1.13).Hence, until recently microcins did

not have a well-defined classification scheme. In 2000 Gaillard-Gendron et al. had

suggested classifying microcins in two groups and Pons et al. (2002) later elaborated

on this attempt and proposed the first microcin classification scheme (Gaillard-

Gendron et al., 2000; Pons et al., 2002). According to this scheme, microcins were

divided into two groups depending on their action site, size and any post

translational modification. Class I mirocins are defined by their activity against

intracellular targets and have a mass of <5kDa. These microcins are post

translationally modified. In contrast, Class II microcins are defined as cell

membrane active agents with a mass of more than >5kDa and they contain no post

translational modifications (Pons et al., 2002).

Figure 1.12 Class I microcin action. A) Microcin (Mcc) J25 gains access to the cell through the outer membrane (OM) receptor FhuA and then utilises the TonB/ExbB/ExbD complex and SbmA protein to translocate into the cytoplasm where it inhibits transcription by attacking the RNA polymerase enzyme (Mathavan et al., 2014). B) Mcc B17 targets the DNA gyrase enzyme that is responsible for supercoiling of the DNA, it gains access to the inner membrane (IM) using the OmpF porin on the OM and is translocated into the cell by exploiting the SbmA protein. C) Although the entry mechanism of Mcc C7/C51 is unknown, it acts by inhibiting aspartyl-tRNA synthetase [from (Duquesne et al., 2007a)].

Page 56: Discovery of antimicrobial peptides active against

Chapter 1

56 | P a g e

Introduction

Following identification of the secreted post translationally modified form of

microcin E492 by Duquesne et al. (2006), which has a mass of 7906.02,

classification parameters suggested by Pons et al. (2002) were not sufficient and this

prompted Duquene et al. (2007) to propose a revised and a more detailed

classification method. According to this classification scheme, the classes are

assigned according to the “presence and type of the posttranslational modifications,

the gene cluster organisation and the leader peptide sequences” (Duquesne et al.,

2007a). Although the two schemes do share similarities, the classification scheme by

Duquesne et al. (2007) also divides classes initially according to their size, placing

low molecular weight microcins into Class I and microcins with a higher molecular

weight into Class II. Class I microcins are defined by extensive post translational

modifications on their peptide backbone and include microcin C (or C7–C51), the

only microcin that is post translationally modified with a nucleotide attachment, and

J25, a lassopeptide that contains a lasso fold (Arnison et al., 2013; Duquesne et al.,

2007a; Mathavan et al., 2014; Rebuffat, 2012).

Class II microcins, however, are further divided into an additional two

subgroups. The Class II microcins with no post translational modifications other than

disulphide bonds are placed under Class IIa (microcins L, V and N), while linear

microcins that may contain C-terminal post translational modifications are collected

under Class IIb (microcins E492, G492, M, H47, I47) (Duquesne et al., 2007a;

Rebuffat, 2012). It is also possible to refer to Class IIb microcins as the siderophore

microcins as all the current members of the group are modified with a siderophore

moiety attached to the C terminal.

Page 57: Discovery of antimicrobial peptides active against

Chapter 1

57 | P a g e

Introduction

Siderophores are secreted to bind and transport a single Fe3+ iron molecule

into the cells through the catechol-siderophore (FepA Cir, Fiu) receptors (Figure

1.13). Once attached to the microcin, it facilitates the trojan horse mode of action of

the microcin (Nolan and Walsh, 2008; Nolan et al., 2007; Rebuffat, 2012; Thomas et

al., 2004; Vassiliadis et al., 2010, 2007).

Figure 1.13 Class IIb microcin action MccE492 (A), MccM (B), MccH47(C) and MccI47 (D)* all exploit the catechol-siderophore receptors and are translocated into the periplasm in a TonB pathway dependent manner. Once inside, MccE492, using the ManY and ManZ membrane proteins of the mannose permease complex as docking stations, inserts itself into the inner membrane where it disrupts the IM potential by allowing protons to leak. On the other hand, MccH47 inhibits ATP synthase by interfering with the F0 component once it gains access to the periplasm. The exact mechanism of action for the remaining microcins of this group is not known [modified from (Duquesne et al., 2007a)] . *MccI47 peptide has not been isolated, but on the basis of predicted amino acid sequence it contains the serine rich C-terminal that is associated with siderophore interaction.

Page 58: Discovery of antimicrobial peptides active against

Chapter 1

58 | P a g e

Introduction

1.7 MISCELLANEOUS RIPP BACTERIOCINS

There have been a number of new RiPP bacteriocin families that have recently

emerged as sources of antimicrobial peptides that are too different to be classified

under the current bacteriocin classification schemes. In this section, some of the

more significant groups are described.

1.7.1 Sactibiotics (Sulphur to α-carbon antibiotics) or sactipeptides

Sactibiotic is the name given to the antimicrobially active members of the

sactipeptide family that are defined by the intramolecular bonds formed between

cysteine sulphur and α-carbon residues during the post translational modification of

these peptides (Arnison et al., 2013; Lohans and Vederas, 2013). Example

sactibiotics identified to date include thuricin CD (Rea et al., 2010), thurincin H (Sit

et al., 2011), subtilosin A (Kawulka et al., 2003) and propionicin F (Brede et al.,

2004).

1.7.2 Thiopeptides

Also known as thiazolyl or pyridinyl polythiazolyl peptides, are a large group

with over 100 members and they are best known for their antimicrobial activity

(Arnison et al., 2013; Bagley et al., 2005; Just-Baringo et al., 2014) but they also

exert anti-cancer (Pandit and Gartel, 2011) and anti-malarial traits (Aminake et al.,

2011). The complex structure of a thiopeptide is often composed with a macrocycle

supporting thiazole rings and the most distinguishing feature is the six membered

Page 59: Discovery of antimicrobial peptides active against

Chapter 1

59 | P a g e

Introduction

nitrogen containing central ring that can be located at three oxidation sites,

piperidine, dehydropiperidine, or pyridine (Arnison et al., 2013; Bagley et al., 2005;

Just-Baringo et al., 2014). One of the best known examples of thiopeptides is the

micrococcin (Su, 1948), and more recent discoveries include philipimycin (Zhang et

al., 2008a) and thiazomycin A (Zhang et al., 2008b).

1.7.3 Bottromycins

Bottromycins have established themselves as a unique class following the

solving of their structure in 2009 by Shimamura and colleagues through their studies

on bottromycin A2 and the ribosomal nature of bottromycin production was revealed

following subsequent studies on the biosynthesis pathway of bottromycin derivatives

including A2 and D (Crone et al., 2012; Gomez-Escribano et al., 2012; Hou et al.,

2012; Huo et al., 2012; Shimamura et al., 2009). Distinctive properties of

bottromycin structure include macrocyclic amidine and a decarboxylated C-terminal

thiazole ring (Arnison et al., 2013). Bottromycins are effective inhibitors of MRSA

and VRE isolates (Kobayashi et al., 2010).

1.7.4 Glycocins

These are a group of glycolpeptide bacteriocins that are glycosylated on

cysteine residues. The group contains two representatives; sublancin 168 and

glycocin F produced by Bacillus subtilis 168 and Lactobacillus plantarum KW30,

respectively (Arnison et al., 2013; Garcia De Gonzalo et al., 2014; Oman et al.,

2011; Stepper et al., 2011; Venugopal et al., 2011) .

Page 60: Discovery of antimicrobial peptides active against

Chapter 1

60 | P a g e

Introduction

1.7.5 Linear azol(in)e-containing peptides (LAPs) and lasso peptides

These are the two RiPP groups that cover a rather grey area. Although there

have been only a few antimicrobially active representatives identified so far, the

group includes AMPs produced by both Gram-negative (MccB17 [Linear azol(in)e-

containing peptide - LAP] and MccJ25 [lasso peptide] from E. coli) (Arnison et al.,

2013; Lee et al., 2008; Li et al., 1996; Salomón and Farías, 1992) and Gram positive

(Plantazolicin [LAP] from Bacillus amyloliquefaciens FZB42 (Scholz et al., 2011)

and lariatins A and B [lasso peptide] from Rhodococcus jostii K01-B0171(Iwatsuki

et al., 2007)) species. Lasso peptides are set apart from the other RiPPs by the lasso

fold that that forms a knotted structure that gives these peptides high stability against

enzymes and denaturing agents (Figure 1.14) (Arnison et al., 2013; Maksimov et al.,

2012; Mathavan et al., 2014). This is a good way of demonstrating how much more

we can learn from bacteriocins and as we learn more how the classification schemes

may further blur and we must expect these schemes to morph to accommodate newer

classes soon.

Figure 1.14 A diagram of Microcin J25 structure, demonstrating the lasso fold [from (Maksimov et al., 2012)]

Page 61: Discovery of antimicrobial peptides active against

Chapter 1

61 | P a g e

Introduction

1.8 POLYKETIDES

Polyketides (PK) are natural products biosynthesised by protein structures

called polyketide synthases (PKS). Polyketides show notable biological activity and

some of the most important antibiotic classes such as the macrolides (erythromycin)

belong in this family as well as important immunosuppressants and anti-tumour

agents. Until recently, PKs produced by bacteria were divided into two main

categories according to the PKS that they are derived from. Modular polyketides

were derived from PKS type 1 (PKS 1) while aromatic polyketides were derived

from PKS type 2 (PKS 2) (McDaniel et al., 2005). However, PKS type 3 driven

aromatic polyketides previously only observed in fungi and plants have been

recently observed in bacteria as well (Gross et al., 2006; Saruwatari et al., 2011).

PKS 1 consists of large multifunctional proteins and the polyketides derived

from it are modular with complex structures. These compounds are generally hard to

modify using chemical methods; however, they are much more responsive in

comparison with PKS 2 derived compounds (Watanabe et al., 2007). Erythromycin

is a PKS 1 derived antibiotic (McDaniel et al., 2005). Although there is not much

reference to their antimicrobial potential, it is noteworthy to state that PKS 2 consists

of monofunctional proteins, while unlike any of its cousins, PKS 3 is a single protein

(Watanabe et al., 2007).

PKS 1, involved in the production erythromycin, also played a key role in the

process of understanding the ribosome independent assembly systems. In 1990,

Cortes et al. and in 1991 Donadio et al. cloned three genes that were involved in the

production of this PK, unlocking the mystery behind the production of these

ribosome independent assembly lines and making genetic manipulations possible

Page 62: Discovery of antimicrobial peptides active against

Chapter 1

62 | P a g e

Introduction

(Cortes et al., 1990; Donadio et al., 1991). Although chemically hard to modify,

these compounds are now observed to be very flexible in terms of genetic

modifications (McDaniel et al., 2005). The main limitation that polyketide discovery

was facing, is now known to be the fact that many of the PKS assembly lines are

encoded by silent genes (Gross et al., 2006).

1.9 NON-RIBOSOMAL PROTEINS

In 1963, Mach et al. performed an experiment using a series of antibiotics

that inhibited ribosomal protein synthesis in Bacillus brevis, yet the group observed

that the cyclic protein tyrocidine was produced in each case, presenting the science

community with the very first evidence of ribosome independent peptide synthesis

(Felnagle et al., 2008; Mach et al., 1963). Some of the most important and widely

used antibiotics, such as penicillin, vancomycin and some of the recently introduced

antibiotics such as the lipopeptide, daptomycin are non-ribosomal peptides (NRP)

(Felnagle et al., 2008). Indeed NRPs are still a lucrative source of antimicrobial

discovery (Cochrane and Vederas, 2014).

Like PKs, NRPs are also assembled by proteins called non-ribosomal protein

synthetases (NRPS), these synthetases are encoded ribosomally and then work

together to assemble a protein and thus these proteins are considered independent of

ribosomal protein synthesis. The assembly mechanism contains a minimum of three

indispensable components; (1) an adenylation domain, (2) a peptidyl carrier protein

domain and (3) a condensation domain (Marahiel et al., 1997; Schwarzer et al.,

2003; Tambadou et al., 2014). Ribosome independence provides the NRPs with a

Page 63: Discovery of antimicrobial peptides active against

Chapter 1

63 | P a g e

Introduction

number of unique structures that clearly distinguish them from their ribosomally

synthesised counterparts, adding to their value as chemical templates (e.g. more

variety for combinatorial chemistry). These proteins are generally found as

macrocyclic, branched macrocylic, dimers or as trimers (Figure 1.15) (Schwarzer et

al., 2003). Also, unlike their counterparts, these compounds can commonly contain

non-proteogenic amino acids (e.g. N-methylated residues and ornithine), heterocylic

rings (e.g. oxazoline), fatty acids and glycolysations (Schwarzer et al., 2003; Sieber

and Marahiel, 2003).

Figure 1.15 Examples of NRP structures. Tyrocidine A is a good example for macrocyclic NRPs, while surfactine is a good a good example of branched macrocyclic NRPs [from (Schwarzer et al., 2003)]

Although very different from each other in many ways, PKS and NRPS

systems share substantial similarities; both systems employ Ppan carriers to achieve

covalent bonding of monomers and elongation of the chain. In fact the similarities

between PKS and NRPS machineries go as deep as to allow hybrid assembly lines

and thus hybrid products to appear (Figure 1.16). Some important hybrids include

epothilone D and ramapycin (Stein, 2005; Walsh and Fischbach, 2010).

Tyrocidine A

Surfactine

Page 64: Discovery of antimicrobial peptides active against

Chapter 1

64 | P a g e

Introduction

1.10 LYSINS:

This review concentrates on bacterially derived antimicrobials and because

the lysins referred to here are endolysins (commonly called lysins in the literature)

and are encoded by double stranded DNA bacteriophages, they will not be reviewed

in detail. However, they are interesting natural products and it is worth mentioning

that the lysogenised DNA of bacterial cells, with assistance from bioinformatics, can

be a good source for identifying these peptides (Fischetti, 2010). Bacterial genetic

libraries that now contain a large number of gene sequences are expected to contain

lysogenised bacterial DNA sequences (Schmitz et al., 2011). Thus, they also contain

the lysin genetic code embedded inside their gene sequence, which is required for the

production of the lysin encoded by the phage; Schmitz et al. (2011) identified a lysin

PlyCM in such a way.

Figure 1.16 A hybrid NRP/PK. NRPS derived structure is coloured in blue while PKS derived structures are red [from (Walsh and Fischbach, 2010)]

Page 65: Discovery of antimicrobial peptides active against

Chapter 1

65 | P a g e

Introduction

1.11 ADVANCES IN NATURAL PRODUCT SCREENING

1.11.1 Purification and identification of the active compound

Rediscovery of known compounds beyond doubt is the biggest hurdle for

natural antimicrobial discovery. It has a high cost in terms of equipment, money and

most importantly time. However, some of the most significant developments in the

field have occurred in purification and de-replication methods. Especially, rapid

identification methods with easily accessible electronic libraries makes de-

replication (differentiation of novel and known products from each other) immensely

easier. Older methods that required milligrams of sample have been replaced with

advanced equipment that is sensitive enough to work with small quantities of

peptides that are likely to be achieved through initial purification efforts (Wright,

2014). Mass spectrophotometry (MS) is generally preferred for the task and with the

introduction of highly sensitive systems such as the orbirtrap, it has become possible

to rapidly fragment peptides using tandem mass spectrometry (MS/MS) and perform

de novo peptide sequencing to identify the amino acid sequences of unknown agents

(Bouslimani et al., 2014; Potterat and Hamburger, 2013).

Nuclear Magnetic Resonance (NMR) is also a useful tool that is becoming

widely used, as in addition to higher field strengths the improvements in the system

allows better room temperature data collection and for the equipment to be installed

in non-specialised enclosures (Breton and Reynolds, 2013; Leeds et al., 2006;

Wright, 2014). Here it should be also noted that UV based detection methods are

also available, but not commonly used (Leeds et al., 2006).

Page 66: Discovery of antimicrobial peptides active against

Chapter 1

66 | P a g e

Introduction

High pressure liquid chromatography (HPLC) is still the gold standard

accepted for purification and recovery of natural products prior to identification.

However, both HPLC and solid phase separation techniques that can be used in

conjunction with HPLC have become more convenient and efficient (Lam, 2007;

Leeds et al., 2006).

By physically combining purification and de-replication methods explained

above, streamlined system can be set-up to increase the speed and sensitivity of

purification and identification. Examples of these systems include preparative LC-

MS, LC-MS/MS, LC-NMR and many others (Harrigan and Goetz, 2005; Leeds et

al., 2006; Nielsen et al., 2011). The data obtained from all of the methods above can

be collected in respective or interchangeable data banks that can be cross checked in

the future, making these methods highly amenable to high throughput screening

(Ibrahim et al., 2012; Wright, 2014; Yang et al., 2013).

In addition, next generation gene sequencing platforms (Pac-Bio, Roche 454,

Illumina MiSeq, Ion Torrent) have significantly increased the overall capacity and

the speed and sensitivity of genome sequencing while driving the costs down; this

feeds into the refreshed natural product discovery process seamlessly (Quail et al.,

2012; Wright, 2014). Through these approaches the need for time consuming reverse

genetics is eliminated and the coding gene sequence of the antimicrobial of interest

can be identified rapidly using the de novo sequencing data obtained from the

MS/MS analysis to query the draft genome of the producer bacteria directly. In

addition, easier access to efficient genome sequencing tools are helping create an

immense library of bacterial genomes making genome mining for natural products a

very feasible and attractive option (Genilloud, 2014).

Page 67: Discovery of antimicrobial peptides active against

Chapter 1

67 | P a g e

Introduction

1.11.2 Identification of producer strains

Isolation and identification of strains producing putative novel natural

products is crucially important both for yield optimisation and further studies. Older

methods have relied on microscopic observation and biochemical tests, which have a

decreased chance of identifying closely related species from each other or present

unrelated species as similar. The development of PCR based identification methods

such as 16S rRNA gene sequence analysis has become an invaluable and reliable

way of rapid bacterial identification (Marchesi et al., 1998; Woo et al., 2008). The

analysis of the data achieved by this approach has helped show the limited

biodiversity between the previously isolated producer species (Davies, 2011).

1.11.3 Genome mining and scanning for natural products

Although target based approaches have failed, some of the reasons behind the

switch are still valid. Conventional methods of natural product research in the pre-

genomic era involved screening of hundreds of bacteria against indicator strains

using methods such as overlay and stab inoculation assays. Although still effective,

these methods are time consuming and have limitations as in order for any activity to

be observed, bacteria of interest should be producing sufficient amounts of the

antimicrobial compound that achieves inhibition of the indicator strain. However,

bacteria may not produce enough of the active agent if optimal conditions are not

achieved and in some cases a gene may be cryptic or can be switched off under lab

conditions (Davies, 2011; Kang and Brady, 2013; Scherlach and Hertweck, 2009).

Page 68: Discovery of antimicrobial peptides active against

Chapter 1

68 | P a g e

Introduction

Cryptic genes are being recognised as a promising untapped source of natural

products including antimicrobial compounds (Zhu et al., 2012). It is possible to

initiate expression of the genes through manipulation of culture conditions or to

improve the production of certain peptides (Rigali et al., 2008; Zhu et al., 2014). A

good example is the isolation of a polythioamide antibiotic from Clostridium

cellulolyticum through addition of aqueous soil extract to the growth media (Lincke

et al., 2010). However, it is also possible to use in silico genomic homology based

techniques that take advantage of rapidly advancing and more affordable whole and

metagenomic sequencing tools. Complimentary bioinformatics software can detect

shared homologies between bacteriocins and when combined with improved

recombinant cloning techniques that can deal better with toxic genes (Guan et al.,

2013; Guo and Jia, 2014) and novel expression procedures (eg: in vitro expression)

(Whittaker, 2013), cryptic genes could be expressed without media optimisation.

These homology approaches rely on motifs, conserved sequences shared

between the genes or gene clusters encoding natural antimicrobials, derived from

existing genetic data to identify in vitro or in silico other gene sequences that may

yield active natural products. Motifs can be whole structures involved in the

synthesis of the antimicrobial or after its modification. The 6-Deoxyerythronolide B

synthase (DEBS), for example, is an essential part of the PKS machinery involved in

erythromycin production and is known to be conserved among biosynthesis

pathways of multiple other polyketide antibiotics belonging to the 14-membered

macrolide family including megalomicin and oleandomycin PKS (McDaniel et al.,

2005). On the other hand LanB, LanM, LanKC and LanT are essential within the

lantibiotic gene loci and, as discussed in Section 1.5.1, contain highly conserved

Page 69: Discovery of antimicrobial peptides active against

Chapter 1

69 | P a g e

Introduction

domains to carry out their activity. These features can be employed to drive genome

mining studies (Arnison et al., 2013; Begley et al., 2009; Blin et al., 2014; McDaniel

et al., 2005; Singh and Sareen, 2014).

Motifs can also be short sequences preserved within the structure of certain

classes of antimicrobial peptides. A good example is the extended pediocin box

YGNG(V/L)xCxxxxCxVxWxxA motif conserved among the class IIa bacteriocins

(Eijsink et al., 1998; Héchard and Sahl, 2002; Kjos et al., 2011). Another sequence

conserved in Gram positive producers of the class II bacteriocins and some in class I,

is the double glycine motif “LSX2ELX2IXGG”, which has been discovered to also

be present in the leader peptide sequence of Gram-negative bacterial microcins,

including MccV (colicin V) and multiple class II microcins (Dirix et al., 2004;

Havarstein et al., 1994). Later, this motif was identified more widely in Gram-

negative bacteria and has been exclusively extended and refined to

“M(R/K)ELX3E(I/L)X2(I/V)XG(G/A)” for Gram-negative organisms and

successfully used to identify in silico other possible antimicrobial peptides (Dirix et

al., 2004; Wang et al., 2011).

Bioinformatic approaches to antimicrobial peptide discovery don’t only

revitalise identification of actively produced natural products, but they open up a

world of new opportunities, for they may efficiently allow the identification of the

cryptic gene clusters effectively (Zazopoulos et al., 2003). Many methods have been

developed to aid in this effort but the first example is a study performed by Challis

and Ravel (2000), who compared known NRPS motifs to the genome of

Streptomyces coelicolor group, a novel NRPS that is involved in the production of

Coelichelin was isolated. Coelichelin is a peptide with siderophore activity. This

Page 70: Discovery of antimicrobial peptides active against

Chapter 1

70 | P a g e

Introduction

study marks the discovery of the first novel secondary metabolite identified by

relying solely on genomic homology (Challis and Ravel, 2000; Taylor and Wright,

2008).

Another early genome scanning procedure developed by Zazopoulos et al.

(2003) is an approach that is more laboratory oriented than in silico. It involves using

DNA fragments generated by shotgun DNA sequencing of high molecular weight

genome fragments for a cloning based screening approach. The larger fragments

generated are inserted into cosmid and bacterial artificial chromosome (BAC)

vectors to produce a cluster identification library (CIL) that possibly contains whole

natural product biosynthetic pathways. The shorter fragments are used in the

generation of genome sequence tags (GST). By comparing the GSTs to a database

containing gene sequences involved in natural product biosynthesis pathways (E.g.

PKS or NRPS coded antibiotics), matching GSTs that are possibly involved in such

processes are flagged to be used as probes to scan the CIL to identify members that

contain matching genes, these plasmids are then processed and sequenced to yield a

complete natural product biosynthesis pathway (Zazopoulos et al., 2003).

Using this method, several active compounds were identified including

antimicrobial, antifungal and anti-cancer drugs (Banskota et al. 2006; Zazopoulos et

al. 2003). An example is the identification of a PKS from the bacterium

Amycolatopsis orientalis that is involved in the production of polyketide ECO-0501,

an agent with significant antimicrobial activity, among 10 other additional secondary

metabolites during the course of the study (Banskota et al. 2006). This demonstrates

the efficiency of in silico based approaches and the possibilities it can unlock.

Page 71: Discovery of antimicrobial peptides active against

Chapter 1

71 | P a g e

Introduction

Today there are web-based programs that utilise and streamline the above

principles to identify similar gene sequences to that of conserved sequences of

known antimicrobial compounds, in order to define new natural products that can

possess potential antimicrobial activity. The Bacteriocin Genome Location tool

(BAGEL3) (http://bagel.molgenrug.nl/index.php/bagel3, last access: 21/05/15) (de

Jong et al., 2010) and secondary metabolite identification tool antiSMASH are two

good examples (Blin et al., 2013). More background for BAGEL software is given

in Chapter 5 of this thesis.

Through all the above methods, structural predictions can be made from the

genetic information obtained, as the physicochemical properties of the active

compound produced are essential for devising assays to detect the compound and/or

to promote conditions that will invoke its production (Lautru et al., 2005) as well as

allowing its cloning and expression.

Page 72: Discovery of antimicrobial peptides active against

Chapter 1

72 | P a g e

Introduction

1.12 AIMS AND OBJECTIVES

The literature review above was written to highlight the increasing threat that

antimicrobial resistance poses and expose the significance of, and underlining

reasons for diminished antibiotic discovery rates. Also, the intention was to

emphasise the wealth of natural antimicrobial products that are present in the

microbial world and highlight their continuing potential for development as novel

antibiotic leads. In addition, the review introduced some of the new methods and

technologies that can remove the barriers that have been plaguing more recent efforts

in novel natural product based antibiotic discovery.

It is the aims of this study to:

1. Help address the problem of antibiotic resistance by developing novel

inhibitors of prominent pathogenic bacteria. This will be achieved by

embracing traditional natural product screening approaches and combining

them with cutting edge dereplication and protein identification techniques.

The following objectives were set to achieve the stated aim:

a. A screening program will be conducted using bacteria from

environmental and clinical samples, to identify potent inhibitors of

Gram positive and Gram-negative bacteria.

b. Purification and dereplication of identified agents is going to be

facilitated through a combination of chromatography, mass

spectrometry and next generation sequencing techniques.

Page 73: Discovery of antimicrobial peptides active against

Chapter 1

73 | P a g e

Introduction

c. Following preliminary characterisation of the inhibitors to confirm

the physicochemical properties and the spectrum of activity, the

toxicity of the identified agents against eukaryotic cells will be

assessed.

2. Explore in silico homology searching software as an alternative means of

natural product antibiotic discovery. In addition, to develop an in house cell

free expression model that will enable expression of these in silico identified

peptides or, if possible, more efficient production of agents discovered as part

of the traditional screening approach. The objectives set to achieve this are as

follows:

a. The in silico bacteriocin discovery software tool BAGEL 3 will be

employed to screen the genomes of a selection of publicly available

anaerobic bacteria. After manual inspection, promising putative

bacteriocin candidates will be short listed for possible expression.

b. An expression model will be built around the PURExpress cell free

expression system and a T7 promoter regulated expression plasmid.

Page 74: Discovery of antimicrobial peptides active against

Chapter 2

74 | P a g e

Methods

2 CHAPTER 2

METHODS

Page 75: Discovery of antimicrobial peptides active against

Chapter 2

75 | P a g e

Methods

2.1 IN VITRO ANTIMICROBIAL PEPTIDE SCREENING AND

CHARACTERISATION

2.1.1 Bacterial isolates

For the natural product discovery programme samples from two sources

(clinical and environmental) were collected between September 2011 and February

2012. Clinical samples were collected from plates inoculated with high vaginal

swabs (HVS) as part of the Central Manchester Foundation Trust routine

bacteriology laboratory activities between January and February in 2012. Plates that

were found to be negative for carriage of likely agents of sexually transmitted

diseases were kindly made available for use in this project. Isolated bacterial

colonies from these plates were screened for antimicrobial agent production by Mr

John Moat (University of Manchester) using the simultaneous antagonism assay (see

below), and the antimicrobial producers were made available for further

identification and characterisation.

The environmental samples used were collected from settle plates placed at

various locations around the Microbiology Department at the University of

Manchester, soil samples were obtained from Whithworth and Platt fields parks in

Manchester, UK and isolates were recovered from swabs of table and cupboard tops

within the Manchester Royal Infirmary. An attempt was made to increase the chance

of discovering bacteriocins that would be effective against clinically relevant

bacteria, by collecting samples from clinical settings and human body that possibly

co-reside with clinically relevant bacteria as bacteriocins are known to be effective

against closely related bacteria (Cascales et al., 2007; Jack et al., 1995). These

samples were inoculated onto Columbia agar base (CAB) (Oxoid Ltd., Basingstoke,

Page 76: Discovery of antimicrobial peptides active against

Chapter 2

76 | P a g e

Methods

England) and Columbia blood agar (CBA) plates (oxoid), which were incubated at

37oC for 24 hours.

Routine propagation of strains under investigation was carried out using

CAB. Liquid cultures were prepared by inoculation of tryptone soy broth (TSB) and

incubated at 37oC for 24 hours. Bacteria that were retained for future study, in this

and other projects, were stored at -80oC in microbank beads (Fisher scientific,

Loughborough, UK). Single colonies isolated from the above screening plates were

tested for antimicrobial activity using the simultaneous antagonism assay (see

Section 2.1.4.1). Bacterial strains displaying promising activity were identified using

16S rRNA gene sequence determination (see Section 2.1.2). The producer

organism/s that displayed strong antimicrobial activity against indicator species were

chosen for further assessment.

2.1.2 Identification of the producer strains using 16S rRNA gene sequencing

The 16S rRNA gene sequencing was performed to identify the organisms

producing inhibitors (producer organisms) of interest. It is one of the most accurate,

time efficient and reliable identification methods currently available (Marchesi et al.,

1998). Chromosomal DNA was extracted from 0.5 ml of a 10ml overnight tryptic

soy broth (TSB) culture of the target producer using the PrepManTM Ultra extraction

kit (Applied Biosystems, California, USA) according to the manufacturer’s

instructions.

Page 77: Discovery of antimicrobial peptides active against

Chapter 2

77 | P a g e

Methods

PCR amplification of the 16S rRNA gene was performed using a set of

universal PCR primers allowing analysis of almost the entire target. These primers

were 63f (5'-CAG GCC TAA CAC ATG CAA GTC-3') and 1387r (5'-GGG CGG

WGT GTA CAA GGC-3') (Eurofins, Germany) (Marchesi et al., 1998) (10pmol/µl

final concentration). The PCR reactions were carried out using 2x BioMix™ Red

(Bioline Reagents Ltd. London, UK), ready to use reaction mixture that included an

ultra-stable Taq DNA polymerase, with 1µl of template DNA extract. The reaction

was carried out on an Biorad T100 thermal cycler (Biorad) and parameters consisted

of an initiation cycle at 94ºC for 2 min, followed by 35 cycles of a denaturation step

at 94ºC for 1 min, primer-annealing step at 60 ºC for 1 min and an extension step at

72ºC for 1 min. The reaction was completed with an extension step at 72ºC for 5

min, and the product was kept at 4ºC prior to clean up for PCR. The product was

visualised using gel electrophoresis (see Section 2.1.3) and quality and quantity was

assessed using a Nanodrop 1000 spectrophotometer (Thermo Scientific, Hemel

Hempstead, UK).

Prior to sequencing, the PCR product was purified using ExoSAP-IT PCR

Clean-up Kit (Affymetrix, High Wycombe, UK). ExoSAP-IT is a combined enzyme

kit that removes any remaining dNTPs and master mix residues that may interfere

with the sequencing reaction. Reagent from the kit (2 µl) was directly added to the

PCR product (5 µl) in 500 µl Eppendorf tubes (Eppendorf, Stevenage, UK). The

mixture was brought to and kept at 37ºC for 15 min to activate the enzymes and

perform the clean-up. Then the enzymes were denatured at 80ºC for 15 minutes.

The DNA sequencing reaction was performed by the staff of Manchester

University sequencing services (Stopford Building) using BigDye® Terminator v3.1

Page 78: Discovery of antimicrobial peptides active against

Chapter 2

78 | P a g e

Methods

Cycle Sequencing Kit (Applied Biosystems) and an Applied Biosystems 3730 (48-

capillary) Genetic Analyzer. The identity of the organisms was obtained by

comparing the 16S rRNA gene sequence to RDP data base (http://rdp.cme.msu.edu/

index.jsp).

2.1.3 Agarose gel electrophoresis

The gels were prepared by dissolving low melting point agarose (Fisher)

between 0.8%-1.5% (W/V) depending on the size of the sample in 1xTAE buffer by

heating in a microwave. The dissolved agarose was cooled to 42oC in a water bath

and 0.01% SYBR safe DNA stain was added and gently mixed. The gel was poured

using an appropriate gel tray and sample comb. Hyper ladder 1Kb (Bioline) and

hyperladder V (Bioline) were used as molecular weight markers and the required

amount of 6x loading buffer (Bioline) was mixed with the desired amount of sample

(~8 µl) to obtain a 1x final concentration solution, which was loaded onto the gel.

Loaded samples were resolved by electrophoresis between 45-60 minutes at 60-90

volts depending on the size of the nucleic acid fragments being examined.

2.1.4 Assays used for the assessment of antimicrobial activity

2.1.4.1 Simultaneous antagonism

The simultaneous antagonism test (Sandiford and Upton, 2012; Tagg and

Bannister, 1979) was initially used to screen and identify potential producer species

and to determine the activity spectrum of the antimicrobials they elicited. The test

involved inoculating the whole surface of an agar plate with an indicator isolate.

Indicator bacteria were chosen for their high susceptibility towards antibacterial

agents to increase the sensitivity of screening. For initial detection, four indicators

were chosen; M. luteus and S. aureus 1195 were chosen to assess presence of anti-

Page 79: Discovery of antimicrobial peptides active against

Chapter 2

79 | P a g e

Methods

Gram positive activity and E. coli DH5α and 414 were used to detect anti-Gram-

negative activity of the tested organism.

Following the initial identification of potential inhibitor producers, various

other indicator strains were included to extend the spectrum of activity being

measured (see Section 2.1.4.2). The inoculua for Gram positive indicators were

adjusted to a suspension equivalent to 0.5 McFarland standard and diluted 1:10.

However, for Gram-negative indicators, the inoculua were diluted 1:100, as

indicated by The British Society for Antimicrobial Chemotherapy (BSAC), for

determination of susceptibility to conventional antibiotics (Andrews et al., 2001).

Afterwards, potential or known antimicrobial producers were stab inoculated onto

the agar using a straight wire.

The experiments were carried out on both CBA and CAB (Oxoid) as

production of certain antimicrobials may depend on availability of specific nutrients

(Patzer et al., 2003; Zhu et al., 2014). The plates were incubated overnight and

observed for inhibition zones (Figure 2.1). Promising producers were taken further

through optimisation and antimicrobials of interest was purified and characterised.

Figure 2.1 Simultaneous antagonism test, the antimicrobial activity of potential inhibitor producers against two indicator E. coli species (414 at the top and DH5α at bottom of the plate) was observed as clear zones of inhibition.

Page 80: Discovery of antimicrobial peptides active against

Chapter 2

80 | P a g e

Methods

2.1.4.2 Deferred antagonism

This experiment was performed to assess the preliminary spectrum of activity

demonstrated by producer strains and followed the guidelines set out by Tagg and

Bannister (1979). Prior to the experiment a heavy inoculum of the producer strain

was streaked diagonally across a CBA plate and incubated aerobically at 37oC

overnight to allow diffusion of the antimicrobial into the agar. In parallel indicator

strains (E. coli 414, vancomycin resistant Enterococcus, B. subtilis, M. luteus, S.

aureus ACTC 1195, S. aureus 308, Meticillin resistant S. aureus, C. difficile,

Klebsiella pneumoniae 169, Klebsiella aerogenes, C. diphtheriae, extended

spectrum beta lactamase (ESBL) producing E. coli M2, ESBL E. coli 169, N.

gonorrhoea, Salmonella Typhimirium and P. aeruginosa 257 (all strains provided by

Manchester University Medical Microbiology laboratory, Manchester, England)

were grown in TSB broth (Oxoid) aerobically at 37oC overnight prior to experiment,

except C. difficile that was tested separately and was grown under anaerobic

conditions using Oxoid AnaeroGen packets in sealed culture jars (Oxoid) and N.

gonorrhoea grown on chocolate agar (Oxoid) and brain heart infusion broth (BHI) in

5% CO2.

The following day, the producers’ growth was removed by using a glass slide and

then the plate was chloroformed. In a fume hood, 5ml of chloroform was applied on

to a tissue placed in the middle of a wide glass petri dish, the agar plate was placed

upside down on the tissue and the agar was subjected to chloroform vapour for 30

minutes to eliminate producer bacteria and possible contaminants. After 30 minutes

the agar was aerated under the fume hood for at least 15 minutes. Following the

completion of the chloroforming process the indicator strains were streaked across

the plate perpendicular to the producer organism’s previous growth location using a

Page 81: Discovery of antimicrobial peptides active against

Chapter 2

81 | P a g e

Methods

sterile cotton swab (Figure 2.2) and the plates were incubated according to the

requirements of the indicator organisms on the plate. On the following day, unless

dictated otherwise by the recommended growth conditions of the indicator bacteria,

the results were processed according to extent of inhibition of the indicators by the

antimicrobial using the scoring system demonstrated below (Figure 2.2).

Figure 2.2 Deferred antagonism assay plate configuration and scoring system [From (Tagg and Bannister, 1979)].* Producer is removed and plate is chloroformed prior to inncoulation of the indicator strains.

2.1.4.3 Well diffusion assay

The well diffusion assay was primarily employed to record results of the

experiments that required semi-quantitative analysis of the antibacterial activity. The

test was performed by piercing up to eight equidistant holes onto a CBA or a CAB

plate with a flame sterilised cork-borer (7mm diameter) and the agar disk was

removed (Figure 2.3). The bases of the wells were sealed with low melting point

agarose before introducing the sample, to prevent leakage. Once sealed, 50µl of a

test material was loaded into each well and allowed to diffuse for 30 minutes or until

the well was dry.

Page 82: Discovery of antimicrobial peptides active against

Chapter 2

82 | P a g e

Methods

After wells had dried, the agar surface was exposed to chloroform vapour as

described above and spread inoculated with the appropriate indicator organism and

incubated appropriately. The results were recorded in arbitrary units per millilitre

(AU/ml) using the following formula (Saeed et al., 2004):

𝑨𝑨𝑨𝑨/𝒎𝒎𝒎𝒎 =(𝐑𝐑𝐑𝐑𝐑𝐑𝐑𝐑𝐑𝐑𝐑𝐑𝐑𝐑𝐑𝐑𝐑𝐑𝐑𝐑 𝐑𝐑𝐨𝐨 𝐡𝐡𝐑𝐑𝐡𝐡𝐡𝐡𝐑𝐑𝐡𝐡𝐡𝐡 𝐑𝐑𝐑𝐑𝐡𝐡𝐑𝐑𝐚𝐚𝐑𝐑 𝐝𝐝𝐑𝐑𝐑𝐑𝐝𝐝𝐡𝐡𝐑𝐑𝐑𝐑𝐝𝐝) 𝐱𝐱 𝟏𝟏𝟏𝟏𝟏𝟏𝟏𝟏

𝑽𝑽𝑽𝑽𝒎𝒎𝑽𝑽𝒎𝒎𝑽𝑽 𝑽𝑽𝒐𝒐 𝒕𝒕𝒕𝒕𝑽𝑽 𝒔𝒔𝒔𝒔𝒎𝒎𝒔𝒔𝒎𝒎𝑽𝑽

2.1.4.4 Spot assay

The spot assay was performed as a time efficient qualitative assessment of

antibacterial activity on liquid samples. This test was conducted by placing 20 µl

equidistant spots of test material onto a CBA, CAB or ISO sensitivity plates (Oxoid)

(Figure 2.4). The spots were allowed to dry and then chloroformed before the

indicator organism was spread inoculated onto the agar base. The inoculated plate

was incubated appropriately depending on the test organism and assessed for

inhibition zones.

Figure 2.3 Image above shows a well diffusion assay performed against indicator organism M. luteus.

Page 83: Discovery of antimicrobial peptides active against

Chapter 2

83 | P a g e

Methods

2.1.5 Optimisation of antimicrobial peptide production

Media used for production and purification of antimicrobial peptides was

optimised, taking into account effects of incubation time and temperature on the final

yield measured in AU/ml. Culture broths assessed included nutrient broth (NB)

(Oxoid), brain heart infusion broth (BHI) (Oxoid), TSB and supplemented forms of

TSB media. A 10ml sample of each broth was inoculated with an overnight starter

culture of B. pumilus J1 or K. pneumoniae A7 and incubated with shaking in a 37oC

incubator at 120rpm overnight. The following day, cultures were centrifuged in a

desktop microfuge at 15000xg, pellets were discarded and supernatants were

individually harvested and serially diluted in sterile distilled water. Dilutions were

assessed for antimicrobial activity using the agar well diffusion assay (see Section

2.1.4.3).

Following the results of this assay the broths that produced the best yield were

examined under different incubation temperatures (370C and 300C) over a 24 hour

period. Every 2 hours, a 1 ml aliquot was removed and tested as described above to

determine when the peptide production was initiated and identify the optimal

fermentation time and temperature for highest possible harvest yield.

Figure 2.4 Spot assay shown in the diagram illustrates antimicrobial activity of a sample against M. luteus.

Page 84: Discovery of antimicrobial peptides active against

Chapter 2

84 | P a g e

Methods

2.2 PROTEOMIC ANALYSIS OF THE IDENTIFIED ANTIMICROBIAL

PEPTIDES

2.2.1 Purification of antimicrobial peptides from the culture media

Purification of antimicrobial agents from the culture media was achieved in 3

steps; (1) capture, (2) intermediate purification and (3) polishing (Figure 2.5).

Purification cultures were initiated by inoculating a 1.5L TSB broth using an

overnight starter culture of the producer organism. The inoculated broth was then

incubated for 18 hours for B. pumilus J1 and 24 hours for K. pneumoniae A7 in a

shaking incubator at 120rpm. Following this, the culture medium was centrifuged at

3600xg at 40C for 20 minutes to remove the cells from the media. An overview of

the purification process is given here and details are provided in sections below. The

initial capture of the peptide from the harvested medium was facilitated using the

polymeric strata-XL liquid chromatography columns that retains samples based on

three properties; pi-pi bonding, hydrogen bonding as well as hydrophobic interaction

(Phenomenex, Macclesfield, UK) (see Section 2.2.1.1).

Figure 2.5 The three-step strategy used for peptide purification. A graphical representation of the purification strategy employed in this project [from (Amersham Biosciences)].

Page 85: Discovery of antimicrobial peptides active against

Chapter 2

85 | P a g e

Methods

The active fractions were pooled and rotary evaporated under pressure using

a Heidolph Laborota 4000 rotary evaporator (Heidolph, Essex, UK) to remove the

residual methanol from strata XL purification. The concentrate was then applied to

Sep-Pak C-18 columns (Waters, Elstree, UK) for intermediated purification (see

Section 2.2.1.2). Sep-Pak C18 cartridges (Waters) are excellent for capturing small

peptides as C18 columns contain longer side chains compared the C4 and C8

columns. However, this also makes them highly hydrophobic. Elution buffer was

removed from the active fractions using an Eppendorf concentrator-plus

(Eppendorf). AMPs of interest were isolated from the resulting semi-pure peptide

mixture using RP-HPLC with a Proteo C-12 HPLC column (Phenomonex) attached

to an AKTA purifier HPLC system (Amersham/GE Healthcare, Little Chalfont, UK)

(see Section 2.2.1.3).

2.2.1.1 Capture using STRATA –XL column:

Elution buffers for use in the Strata XL column [column volume (CV) =

15ml] were prepared with increasing concentrations of acidified methanol [30% v/v,

50% v/v and 90% v/v in water + 0.01% trifluoroacetic acid (TFA)] to gradually strip

the bound peptides from the column. Fractions were collected using a 50ml syringe.

Before loading the sample, the column had to be conditioned to enhance the

binding of the peptides and to remove any impurities that may remain from the

manufacturing process. The conditioning involved washing the column with 60ml (4

CV) of 99.9% HPLC grade methanol (Fisher Scientific) to strip any protein residues

from the column. Then, 120 ml of milli-Q water at pH7 was passed through the

column to remove methanol and conclude the conditioning process.

Page 86: Discovery of antimicrobial peptides active against

Chapter 2

86 | P a g e

Methods

Prior to loading, the sample pH was adjusted to pH 5.8-6.2 to improve

binding of the peptide to the column. TFA (Sigma-Aldrich) was used to decrease the

sample pH and sodium hydroxide (NaOH) (Fisher scientific) to increase it. A 20µl

aliquot was used for spot assay on a CBA and/or CAB plate as a control to ensure

the sample was active. Once the pH was adjusted, the sample was loaded onto the

column and the resulting fractions were slowly collected and labelled as flow

through (FT). During the elution, at every 200 ml a 20µl spot was placed onto a

labelled agar plate to determine the saturation point of the column. The column

saturation was reached at around 500ml for all samples tested. After the desired

amount of sample was passed through, the column was washed with 60ml of ultra-

pure water at pH7 to remove any unbound sample. The resulting fraction was

labelled as wash through (WT).

Subsequently, 45mls of 50% methanol was introduced to the column to strip

molecules that were weakly bound. Finally, peptides attached to the column were

gradually eluted in four 15 ml fractions of 90% methanol (pH2). All fractions

collected for B. pumilus J1 were tested against M. luteus and for K. pneumoniae A7

against E. coli DH5α using the spot assay to confirm presence of the AMP of

interest.

2.2.1.2 Intermediate sample cleaning with Sep-Pak C18 columns

Two different buffers were required for the experiment, Buffer A [MiliQ

water + 0.01% TFA and 2 % Acetonitrile (ACN) (TFA, Sigma-Aldrich; Acetonitrile,

Fisher scientific)] and Buffer B [ACN + 0.01% TFA]. Prior to sample loading, Sep-

Pak C18 cartridges were also conditioned; the column was stripped of protein

residues by passing 1 CV (Sep-Pak C18 light cartridge CV= 1ml; Sep-Pak C18 plus

Page 87: Discovery of antimicrobial peptides active against

Chapter 2

87 | P a g e

Methods

cartridge CV= 5ml) of Buffer B through and equilibrating with 2 CV of Buffer A.

The 2% ACN in buffer A prevents highly hydrophobic side chains of Sep-Pak C18

cartridge from ‘matting down’ in an attempt to cluster away from the water by

slightly reducing the surface tension.

After column conditioning, 1ml of Strata-XL purified active sample adjusted

to ~pH 2 using TFA to achieve optimal sample binding was loaded on to the Sep-

Pak C18 columns. The sample was then eluted and the FT fraction was collected and

tested for activity in every 250µl to determine the column’s saturation point. One

column volume of sample was observed to be the limit for each agent examined.

Then the cartridge was washed with 2 CV of Buffer A to remove unbound

sample and the WT fraction was collected. Bound peptides were eluted using 1 CV

of increasing concentrations of ACN, in 10% increments (10-90% v/v in 0.01%

TFA) and the resulting fractions were collected in 1.5 ml eppendorf tubes

(Eppendorf).

2.2.1.3 Reverse Phase High Pressure Liquid Chromatography

HPLC allows very high-resolution separation of the peptides present within a

sample. For this project, reverse phased-HPLC (RP-HPLC) was performed on the

Sep-Pak C18 purified and concentrated active fractions, as the polishing step. The

Jupiter C18 250x10mm HPLC columns (CV= 10ml; Phenomonex) were used

together with an AKTA purifier system (GE Healthcare).

The buffers were the same as those used for Sep-Pak C18 cartridges,

however buffers A and B were degassed by vacuum filtration for the HPLC

procedure. This prevents pressure build up as liquids are non-compressible when

mixed with gases. Pressure increases during HPLC may cause compression of the

Page 88: Discovery of antimicrobial peptides active against

Chapter 2

88 | P a g e

Methods

column contents causing erratic baseline readings or rupture the column. Buffer A

and B were connected to the AKTA purifier via the designated tubing. The following

procedures were carried out using the software provided by the manufacturer.

The column was conditioned by running 2 CV of Buffer B, followed by 5 CV

of Buffer A. For the initial elution, a program with a complete continuous gradient

from 0% Buffer A and 100% Buffer B leading towards 100% Buffer A in 8 CV at a

flow rate of 1ml/min, was entered to the AKTA purifier management software.

During the elution, 0.5ml fractions were collected in individual eppendorf tubes and

labelled. Each fraction represented a 0.53% increase in ACN concentration. The

proteins eluted during the purification process were visualised and recorded using

215 nm and 280 nm UV absorbance. Once the initial elution was completed, all

fractions collected were tested manually using the spot assay. Once the elution

fractions for each AMP were determined, to cater for a more time efficient run the

gradient was adjusted to 30% ACN+0.01% TFA and rising to 90% ACN+0.01%

TFA in 4 CV at a flow rate of 1ml/min. Where possible, the procedure was repeated

multiple times until the highest possible purity was obtained.

2.2.2 SDS-PAGE analysis and gel diffusion agar overlay assay

Sodium dodecyl sulphate polyacrylamide gel electrophoresis (SDS-PAGE)

electrophoresis using 4-12% Nu PAGE Bis-Tris Bolt gels (Life Technologies) was

employed to visualise the gross content of Sep-Pak C18 and HPLC purified active

fractions and the products of in vitro expression studies (see Section 2.5.10 below).

Prior to electrophoresis, the samples were treated with lithium dodecyl sulphate

(LDS) sample buffer (Life Technologies) and reduction agent (Life Technologies) as

per manufacturer’s instructions to denature and reduce the sample. See Blu 2 (Life

Page 89: Discovery of antimicrobial peptides active against

Chapter 2

89 | P a g e

Methods

Technologies) pre-stained protein ladder was used to estimate the protein masses.

The gels were run on Bolt mini gel (Life Technologies) electrophoresis tanks with

MES (2-ethanesulfonic acid) buffer (Life Technologies). The protein bands were

stained and visualised using the Coomassie brilliant blue R-250 (Fisher Scientific).

Antimicrobial activities of the visualised bands were tested with a gel

diffusion agar overlay assay. Assay detected bands were excised after multiple wash

cycles with milli-Q water for at least 4 hours. Excised gel strips were then gently

placed on the surface of a CAB plate. Simultaneously, 10ml CAB aliquots were

melted and cooled to 450C and, while still aqueous, the agar was inoculated with a

0.5 McFarland standard suspension of M. luteus cells equivalent to a 1:10 dilution.

The inoculated agar was poured over the gel strips and allowed to solidify and the

plate incubated at 37oC overnight.

2.2.3 MALDI-ToF mass determination and ESI-MS scan of the RP-HPLC

Purified Active Fractions

The molecular masses of the peptide particles recovered within the active

fractions obtained from the RP-HPLC purification process were analysed for

dereplication purposes using the matrix assisted laser desorption/ionisation time of

flight (MALDI-TOF) technique. Prior to analysis, 1ml of matrix was prepared by

mixing 10 mg α-cyano-4-hydroxycinnamic acid (Sigma-Aldrich) with 0.5ml ACN

(50%) and 0.5ml ethanol (50%) (Fisher Scientific). Then, 1μl of the sample was

placed on a 96 well ground steel MALDI plate (Bruker, Coventry, UK.) followed by

1μl of the matrix and the mixture was left to air dry at room temperature. Once

dried, the samples were processed using the Bruker Daltonics Ultraflex II TOF/TOF

Page 90: Discovery of antimicrobial peptides active against

Chapter 2

90 | P a g e

Methods

(Bruker), as per manufacturer’s instructions. On average, five shots were fired at

randomly picked locations on the sample well and added to the generated spectra.

The spectra were analysed using the predefined peak selection method designed by

the Biomolecular Analysis Facility, Michael Smith building, The University of

Manchester, using the on-board software. Samples with prominent peaks were also

analysed using electro spray ionisation mass spectrometry (ESI-MS) using an

Orbitrap mass spectrometer (Thermo Scientific) for further confirmation. This was

performed as a service by staff at the Biomolecular Analysis Facility.

The mass assigned to the active peptide was checked against the online

bacteriocin database, Bactibase (Hammami et al., 2010) and the published literature

to identify any other previously identified bacteriocins that contain an identical mass.

2.2.4 De novo peptide sequencing (MS/MS)

De novo peptide sequencing is a powerful tool that uses tandem mass

spectrophotometry to fragment and analyse peptides to determine the sequence of

constituent amino acids. The analysis of the subsequent data yields strings of amino

acid sequences that are called peptide tags. These tags can be compared with draft

genome sequence data (see Section 2.2.5) to search for a unique match. This

approach was used to search for the genes encoding AMPs in the B. pumilus J1 and

K. pneumoniae A7 strains.

The procedure is initiated with digestion of the query peptide with modified

sequencing grade trypsin (Promega, Southampton, UK). This procedure could be

performed both in gel and in solution; both versions were performed in this project.

This was done to enable the comparison of the results for greater confidence and also

by running the SDS gel overlay assay samples simultaneously in the same gel (see

Page 91: Discovery of antimicrobial peptides active against

Chapter 2

91 | P a g e

Methods

Section 2.2.2), it was possible to link the physically observable antimicrobial protein

band to the peptide tags obtained to antimicrobial activity.

2.2.4.1 In solution digestion

Dried peptide fraction was suspended in 100μl of 6 M Urea in Tris buffer

(pH 7.8). To ensure specific cleavage by trypsin, the peptide was reduced by adding

5μl of 10 mM dithiothreitol (DTT) (Sigma-Aldrich) to the suspension, vortexed and

allowed to react for 30- 60 minutes at room temperature. Reduced peptide was then

alkylated using 20μl of 50 mM 2-chloroacetamide (Sigma-Aldrich), the reaction

mixture was incubated in a dark place at room temperature for 30-60 minutes. This

was followed with re-treatment with 20μl DTT for a further 30-60 minutes. Prior to

addition of trypsin, the Urea was diluted with 775μl MilliQ water. Sequencing grade

modified trypsin (Promega) was added to the solution at a 1:50 ratio and mixed.

Digestion was carried out overnight in a 37oC incubator.

The digested peptide was recovered using Stage Tips (Stop and Go

Extraction Tips) containing stanched monolithic C18 adsorbent (Waters). The Stage

tips were treated as miniature Sep-Pak C18 columns (see Section 2.2.1.2), suspended

through the lid of an Eppendorf tube. The buffers or the samples were loaded into the

tip and the apparatus was centrifuged on a desktop microfuge briefly to facilitate the

flow of the sample and the buffers. Sample was eluted from the column using 80%

ACN + 0.01%TFA in a single step. Eluted peptide was dried down using the speed

vac and re-suspended in solution A (see Section 2.2.1). MS/MS fragmentation of the

digested peptide was performed using the Orbitrap LC-MS (Thermo-scientific) by

the Biomolecular Analysis Facility, Michael Smith building, The University of

Manchester.

Page 92: Discovery of antimicrobial peptides active against

Chapter 2

92 | P a g e

Methods

2.2.4.2 In gel digestion

In gel digestion was performed on excised gel fragments obtained following

the SDS PAGE gel analysis (see Section 2.2.2) of the Sep-Pak C18 purified active

fractions. The mass of the excised bands chosen for this experiment corresponded to

the HPLC purified peptide and to the correct molecular weight interval determined

by the See Blu 2 protein weight marker. Replicates of the gel fragments were tested

for antimicrobial activity using the agar gel overlay assay (see Section2.2.2). The

excised fragments were cut into smaller cubes prior to digestion and washed by

suspension in 100% ACN for 5 minutes and then in MilliQ water to remove the

coomassie blue stain. This step was repeated 3 times. After this point, the digestion

procedure was applied as explained in Section 2.2.4.1. However, prior to Stage Tip

recovery, the digested peptide was removed from the gel pieces by vigorous

vortexing and the supernatant was applied to the Stage Tips for peptide recovery.

The recovered peptide was subjected to the same analysis as the in solution digest.

2.2.4.3 Bioinformatics analysis of the generated data

De novo peptide sequencing was conducted both manually and with analysis

software packages PEAKS studio7 (Zhang et al., 2012) (Bioinformatics Solutions,

Waterloo, Canada) and MASCOT (Perkins et al., 1999) (Matrix Science, London,

UK) on the raw data resultant from the MS/MS fragmentation. Generated peptide

tags were compared with the draft genome of the producer organism (see Section

2.2.5) using the BLAST facility within the CLC main workbench 7 (Qiagen). Where

required, multiple alignments of the identified amino acid sequences were performed

in Jalview (Waterhouse et al., 2009) using the ClustalW alignment facility (Larkin et

al., 2007). Neighbour joining trees were also created in Jalview. Secondary structure

Page 93: Discovery of antimicrobial peptides active against

Chapter 2

93 | P a g e

Methods

prediction was carried out using Jnet secondary structure prediction service (Cole et

al., 2008; Cuff and Barton, 2000).

2.2.5 Genome sequencing and annotation of the draft genome of producer

strains

DNA for genome sequence determination was extracted from cells contained

in 0.5ml of an overnight culture of B. pumilus J1 or K. pneumoniae A7, which had

previously been inoculated with a single colony from CBA plates and incubated at

37oC. DNA extractions were carried out in house using the Blood and Tissue kit

(Qiagen, Manchester, UK) as per manufacturer’s instructions. Integrity of the DNA

was observed by loading an 8μl sample onto a 1% agarose gel in TAE, followed by

electrophoresis, and the concentration was measured using a Qbit device (Life

Technologies), as per the manufacturer’s instructions. DNA was stored at -20oC until

required for analysis.

The draft genome sequence was determined using the services of the Centre

for Genome Research, University of Liverpool. A shotgun library of DNA fragments

was sequenced using an Illumina HiSeq device with 150bp paired end reads. The

raw data was presented trimmed of Illumina adapter sequence using Cutadapt.

Sickle, a window based quality trimming software which uses a sliding window to

monitor the read quality and length, was set to trim the sequences when the read

quality score within the window was below 20 and remove the sequence if total read

length was below 10bp (Del Fabbro et al., 2013; Joshi and Fass, 2011). Raw data

were assembled using the CLC Main Workbench 7 de novo assembly tool. The

assembled sequences were annotated using the PROKKA prokaryotic genome

annotation software tool (Seemann, 2014) and by using xBASE genome annotation

Page 94: Discovery of antimicrobial peptides active against

Chapter 2

94 | P a g e

Methods

pipeline (Chaudhuri et al., 2008), for B. pumilus J1 using B. pumilus SAFR-32

(Accession number NC_009848) as a reference and using Klebsiella pneumoniae

XH209 (Accession number NZ_CP009461) as the reference for K. pneumoniae A7.

The annotated genome was visualised using Artemis software (Rutherford et al.,

2000) and the assembled draft sequences were compared with reference genomes

using the BLAST ring image generator (BRIG) (Alikhan et al., 2011) software to

generate a graphical representation of the differences between B. pumilus J1 and K.

pneumoniae A7 with their respective reference organisms. Where required, genome

annotations were corrected manually by direct pairwise gene alignments via Jalview

(Waterhouse et al., 2009) and/or database search with the position-specific iterated

BLAST (PSI-BLAST) (Altschul et al., 1997) algorithm using predicted peptide

sequences generated by six frame translation of the genome.

2.3 CHARACTERIZATION OF THE PURIFIED ANTIMICROBIAL PEPTIDES

Three promising antimicrobial peptides were isolated following the application

of above methodology. From B. pumilus J1, characteristics of peptides NI03 and

NI04 are described in Chapter 3 and from K. pneumoniae A7, the peptide NI05 is

addressed in Chapter 4. Due to reasons discussed in respective chapters concerning

the individual peptides, not all characterisation steps were performed on each

peptide, and the Table below (Table 2.1) summarises the tests performed for each

peptide.

Page 95: Discovery of antimicrobial peptides active against

Chapter 2

95 | P a g e

Methods

Table 2.1 A summary of the characterisation assays performed on each agent identified during proteomic analysis

Test Peptide NI03 Peptide NI04 Peptide NI05

Deferred Antagonism* Physicochemical analysis Minimum inhibitory concentration determination

Haemolysis assay

Eukaryotic Cell toxicity studies

UV absorbing material leakage assay

*Deferred antagonism assay was performed to characterise the spectrum of activity of the producer strains being tested. The activity observed represents the combined activity of any antimicrobial agents that might be expressed by the organism of interest.

2.3.1 Physicochemical analysis of purified antimicrobial peptides (heat and

enzyme stability tests)

Preliminary investigation of the physicochemical properties of the detected

peptides was carried out using heat and enzyme stability assays. For the heat stability

assay, a known concentration of peptide NI04 was incubated at 800C in a heating

block for up to 4 hours. For enzyme stability assay aliquots of known concentrations

of NI04 were treated with the hydrolytic enzymes α-amylase, α-chymotrypsin,

lipase, protease and trypsin (Sigma-Aldrich) at 1 mg/ml, while proteinase K (Sigma-

Aldrich) was used at 10mg/ml. NI04 was incubated with each of these enzymes for 4

hours.

The effect of both heat and enzyme stability assays were measured by

making doubling dilutions of the treated peptide aliquots and comparing the

consequent reduction in antimicrobial activity to that of the untreated peptide

sample, using the well inhibition assay described above.

Page 96: Discovery of antimicrobial peptides active against

Chapter 2

96 | P a g e

Methods

2.3.2 Determination of minimum inhibitory concentration

The minimum inhibitory concentration (MIC) of AMPs were determined

using broth micro-dilution of the Sep-Pak C18 cleaned peptide fractions following

CLSI guidelines (CLSI, 2012). The total protein content of pooled dried and

suspended fractions was calculated using the bicinchoninic acid assay (BCA assay)

(Smith et al., 1985) with the Pierce BCA Protein Assay Kit (Thermo Scientific) as

per manufacturer’s instructions. Two fold dilutions of the peptide were prepared in

cation adjusted Mueller-Hinton broth (CA-MHB) (Oxoid) and dispensed into the

wells in 0.1ml fractions. A bacterial suspension was prepared using colonies directly

from agar surface in CA-MHB and adjusted to 0.5 McFarland standard (1 × 108

colony forming units (CFU)/mL). Within 15 minutes of preparation the suspension

was diluted 1:20 in CA-MHB and the AMP containing broths were inoculated with

10ul (0.01ml) of this inoculum to yield approximately 5 × 104 CFU/well as

suggested by the CLSI. The prepared test plate was incubated in a 37°C incubator for

20 hours in an aerobic atmosphere.

2.3.3 Haemolysis assay

Haemolysis assays were performed to assess the toxicity of the

antimicrobial peptide of interest towards mice red blood cells (RBC) (kindly donated

by Dr Peter Warn, University of Manchester). The test was conducted using mice

RBCs, which were harvested by centrifugation at 400xg, washed with phosphate

buffered saline (PBS) and suspended at a concentration of 5% in PBS containing the

AMP of interest at doubling dilutions from 1:10 to 1:160. Haemolytic activity was

measured using a spectrophotometer at 490 nm with 1% Triton-X 100 as control.

The haemolytic activity was calculated using the following formula (Vaucher et al.,

2010):

Page 97: Discovery of antimicrobial peptides active against

Chapter 2

97 | P a g e

Methods

%Haemolysis =𝐴𝐴𝐴𝐴𝐴𝐴 𝑜𝑜𝑜𝑜 𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝 𝑝𝑝𝑡𝑡𝑡𝑡𝑝𝑝𝑝𝑝𝑝𝑝 𝐴𝐴𝑏𝑏𝑜𝑜𝑜𝑜𝑝𝑝 × 𝐴𝐴𝐴𝐴𝐴𝐴 𝑜𝑜𝑜𝑜 𝑢𝑢𝑢𝑢𝑝𝑝𝑡𝑡𝑝𝑝𝑡𝑡𝑝𝑝𝑝𝑝𝑝𝑝 𝐴𝐴𝑏𝑏𝑜𝑜𝑜𝑜𝑝𝑝

𝐴𝐴𝐴𝐴𝐴𝐴 𝑜𝑜𝑜𝑜 𝑇𝑇𝑡𝑡𝑝𝑝𝑝𝑝𝑜𝑜𝑢𝑢𝑇𝑇100 𝑝𝑝𝑡𝑡𝑝𝑝𝑡𝑡𝑝𝑝𝑝𝑝𝑝𝑝 𝐴𝐴𝑏𝑏𝑜𝑜𝑜𝑜𝑝𝑝 × 𝐴𝐴𝐴𝐴𝐴𝐴 𝑜𝑜𝑜𝑜 𝑢𝑢𝑢𝑢𝑝𝑝𝑡𝑡𝑝𝑝𝑡𝑡𝑝𝑝𝑝𝑝𝑝𝑝 𝐴𝐴𝑏𝑏𝑜𝑜𝑜𝑜𝑝𝑝× 100

2.3.4 Eukaryotic cell toxicity studies

Toxicity of peptide NI04 against eukaryotic cells was measured using the

trypan blue exclusion assay on keratinocyte cell lines and neutral red uptake assays

on Vero cells. The experiments were conducted using semi-purified stock peptide

fractions at up to 18x MIC concentrations for M. luteus, diluted in cell culture

growth media. SDS (20mg/ml) (Fisher Scientific) was used as a positive control and

neat maintenance medium as the negative control. Cells were incubated in 5% CO2

at 37oC overnight.

2.3.4.1 Trypan blue exclusion assay with keratinocytes

For the trypan blue exclusion assay (Louis and Siegel, 2011), old culture

media was removed from 100% confluent keratinocyte tissue culture monolayers

(kindly donated by Dr Catherine O’Neil, University of Manchester) and replaced

with keratinocyte growth medium 2 (PromoCell, Heidelberg, Germany) containing

doubling dilutions of stock peptide from a starting concentration of 18X MIC (M.

luteus). Cells were incubated in 5% CO2 at 37oC overnight. The following day, the

keratinocytes were observed under the microscope for possible cytopathic effects

and following microscopic evaluation, medium was removed and cells were

harvested using trypsinisation. Harvested cells were then diluted 1:1 with 0.4%

trypan blue (Fisher Scientific) and cell viability was assessed visually under an

inverted microscope using a haemocytometer. Both viable cells (non-stained) and

dead (stained) ones were counted.

Page 98: Discovery of antimicrobial peptides active against

Chapter 2

98 | P a g e

Methods

2.3.4.2 Neutral red uptake assay

The neutral red assay is based on the ability of live cells to incorporate the

neutral red dye in their lysosomes. The variation of this quantitative viability test,

performed on Monkey Vero cell lines in this project, has been based on the method

by Repetto and colleagues (Repetto et al., 2008). A 40 mg/ml neutral red solution

(Fisher Scientific) was prepared by dissolving the powdered dye in distilled water.

The cells were grown in 96 well tissue culture plates (Corning, Amsterdam,

Netherlands) with antibiotic free DMEM growth media (Dulbecco's Modified Eagle

Medium, DMEM+10% foetal bovine serum) (Life Technologies) to half confluence

overnight at 370C in 5% CO2. Half confluent cell monolayers were then treated with

growth medium (without phenol red) containing doubling dilutions of peptide NI04

up to 18x MIC, with negative (DMEM growth media) and positive [2% Triton X-

100 (Fisher Scientific)] controls set up in triplicate. Treated cells and the neutral red

solution were incubated separately overnight in a 370C, 5%CO2 incubator. Once

incubation had concluded, cells were inspected under a microscope for

morphological abnormalities.

Once the observations were noted, the treatment media was removed from

the wells and replaced with 100μl neutral red solution and incubated for 2 hours at

5% CO2 at 37oC. Following the exposure, neutral red solution was removed and cells

were washed with 150μl of PBS. Then 150μl of neutral red de-stain solution was

added into the wells, which were shaken for 10 minutes until a homogenous solution

was obtained. The OD of the resulting solution was measured in a plate reader at

540nm.

Page 99: Discovery of antimicrobial peptides active against

Chapter 2

99 | P a g e

Methods

2.3.5 UV Absorbing material leakage assay

This assay was performed to gather preliminary data concerning the mode of

activity of peptide NI04. Overnight broths of M. luteus and S. aureus 1195 cells were

prepared in TSB broth and diluted to OD 1.0 (at 600nm) using a spectrophotometer.

Adjusted cell suspensions were transferred into an Eppendorf tube and the bacterial

cells were harvested by centrifugation at 3600xg for 10 minutes. Supernatant was

discarded and the cell pellet was washed twice with PBS and then suspended in PBS

solution containing 1xMIC concentration of Sep-Pak purified antimicrobial peptide

(S. aureus 20μg/ml and M. luteus 10μg/ml) and incubated for four hours at 370C

aerobically. Every hour, a 150μl sample was taken and centrifuged at 15000xg for 10

minutes in a desktop centrifuge to remove bacterial cells. Then, 100μl of the cell free

supernatant was transferred into a 96 well UVstar (Grenier) UV transparent plate and

UV absorbance was measured at 260nm using a FLUOstar Omega (BMG-Labtech,

Ortenberg, Germany) UV/VIS spectrophotometer.

2.4 IN SILICO GENOME MINING AND IDENTIFICATION OF PUTATIVE

BACTERIOCIN LEADS Draft bacterial genomes of the producers B. pumilus J1 and K. pneumoniae A7

(as described in Chapters 3 and 4) were subjected to bioinformatic analyses, together

with genomes of 35 anaerobic bacteria from Clostridium and Proprionibacterium

genera (Clostridium acetobutylicum ATCC 824, Clostridium acidurici 9a,

Clostridium acetobutylicum 824, Clostridium beijerinckii NCIMB 8052, Clostridium

botulinum A ATCC 19397, Clostridium botulinum BKT015925, Clostridium

botulinum A2 strain Kyoto, Clostridium botulinum A3 strain Loch Maree,

Clostridium botulinum B strain Eklund 17B, Clostridium botulinum B1 strain Okra,

Clostridium botulinum Ba4 strain 657, Clostridium botulinum E3 strain Alaska E43,

Page 100: Discovery of antimicrobial peptides active against

Chapter 2

100 | P a g e

Methods

Clostridium botulinum H04402 065, Clostridium cellulolyticum H10, Clostridium

cellulovorans 743B, Clostridium_kluyveri DSM 555, Clostridium lentocellum DSM

5427, Clostridium ljungdahlii DSM 13528, Clostridium novyi NT, Clostridium

perfringens SM101, Clostridium perfringens strain 13, Clostridium phytofermentans

ISDg, Clostridium saccharolyticum WM1, Clostridium saccharoperbutylacetonicum

N1 4HMT, Clostridium species BNL1100, Clostridium species SY8519, Clostridium

stercorarium DSM 8532, Clostridium sporogenes strain DSM 795, Clostridium

tetani E88, Clostridium tetani 12124569, Clostridium thermocellum ATCC 27405,

Propionibacterium acidipropionici ATCC 4875, Propionibacterium acnes SK137,

Propionibacterium acnes TypeIA2 Pacn17, Propionibacterium propionicum

F0230a),using the bacteriocin research tool BAGEL3

(http://bagel.molgenrug.nl/index.php/bagel3) (van Heel et al., 2013) (Figure 2.6).

Annotated putative bacteriocin candidates/clusters returned by BAGEL software

were then short-listed. Those putative bacteriocins that had significant homology to

known bacteriocins with an E value less than 0.001, as identified by BAGEL or

through use of PSI BLAST and the Bactibase (Hammami et al., 2010)internal

BLAST homology tool, were included in this short-list.

Page 101: Discovery of antimicrobial peptides active against

Chapter 2

101 | P a g e

Methods

Figure 2.6 Overview of the proposed workflow described below for in silico mining of genomic sequence data for novel AMPs.

Page 102: Discovery of antimicrobial peptides active against

Chapter 2

102 | P a g e

Methods

2.5 CLONING AND EXPRESSION METHODS

2.5.1 Insert construction and primer design for cell free protein expression

To validate the concept that functional antimicrobial peptides can be obtained

from mined putative bacteriocin sequences using commercially available cell free

expression systems, assuming they conferred antimicrobial activity, attempts were

made to clone and express the pumicin NI04 structural gene, pmnA using a cell free

expression system. What follows is an overview of the approaches used and a

detailed description of each stage in the process.

Once the target gene (insert) was selected, using sticky ends generated by

restriction enzymes, it was inserted into the chosen expression vector, pET28a,

where expression is tightly controlled by the T7 promoter region. Prior to insertion,

the target gene was PCR amplified using manually designed PCR primers flanked by

restriction enzyme sites, Nco1 and BamH1 to exclude the polyhistidine-tagged (His-

tagged) region of the pET28a expression vector. This was done to avoid any possible

interference to antimicrobial activity that may occur due to inclusion of a His-tag.

Restriction sites were followed by 4 base long random overhangs to improve

restriction enzyme binding (Green and Sambrook, 2001) (Table 2.2). The target

binding portion of the primers were designed to contain between the first 18-22 base

pairs of the 3’ and 5’ ends of the target gene, depending on suitability of the

nucleotide content.

Page 103: Discovery of antimicrobial peptides active against

Chapter 2

103 | P a g e

Methods

Table 2.2 A list of primers employed during the study. Blue nucleotides indicate the restriction enzyme cut site, while red nucleotides highlight the start and stop codons.

However, a modification had to be made for the pmnA gene to utilise the

NcoI cut site as it contains the ATG start codon and dictates the first nucleotide of

the inserted sequence as Guanine. To overcome this problem pmnA gene primers

were designed with an alternative, BsaI, cut site. BsaI cuts downstream of its binding

site and creates an NcoI compatible overhang (Figure 2.7).

The pET28a plasmids were extracted and purified from transformed E. coli

XL-1 blue cells (Agilent Technologies, Stockport, UK) from the laboratory stock

using a QIAprep spin miniprep kit (Qiagen) as per manufacturer’s instructions.

Bacterial DNA extraction was performed using the Qiagen blood and tissue kit

(Qiagen). Purified DNA was quantified using a nanodrop spectrophotometer

(Thermoscientific) and was visualised by agarose gel electrophoresis (see Section

2.1.3). In addition, 0.8% agarose gels were used to check integrity of extracted

genomic DNA and plasmids. All samples were stored at -200C.

Target Forward primer (5’end of target) Reverse primer (3’end of target)

pmnA gene 5’-GCCG(GGTCTC)TCATG TCAGGAATTATTCGCGTAACC

-3’

5’-GCC(GGATCC)TTA GCCGCGGATTTGGCTAGC-

3’

T7 sequencing

primers TAATACGACTCACTATAGGG GCTAGTTATTGCTCAGCGG

Figure 2.7 Utilisation of NcoI cut site using the BsaI restriction site.

Page 104: Discovery of antimicrobial peptides active against

Chapter 2

104 | P a g e

Methods

The target gene sequence (insert) was amplified using Phusion® high-fidelity

PCR master mix with HF buffer (Thermoscientific) for the cloning procedure and

Biomix Red (Bioline) for routine diagnostic PCR checks. Reaction mixture and

conditions are recorded in Table 2.3 below. The PCR reactions were carried out in a

Bio-Rad thermal cyclers. PCR products were purified by Qiagen PCR clean up kit

(Qiagen) and eluted from columns in 50µl for storage at -20oC.

Table 2.3 The content of the PCR reaction mixture used for amplification of the inserts for use in construction of plasmid vector.

Component 50µl Reaction

2x Phusion® high-fidelity PCR master mix with HF buffer

25µl

Forward Primer 0.5µM

Reverse Primer 0.5µM

Template DNA 50-250ng

DMSO (100%) 1.5µl

dH2O Make up to 50µl final volume

Amplified inserts were analysed by agarose gel electrophoresis (see Section

2.1.3; 1.5 % gels) and a Nanodrop Spectrophotometer (Thermoscientific) was

employed to estimate product yield and purity.

Double restriction digests were then set up for both the plasmid and the

insert, using restriction enzymes BsaI-HF (New England Biosciences, Herts, UK

[NEB]) and BamHI-HF (NEB). The reactions (50µl) were prepared according to the

manufacturer’s instructions and conducted for at least 1 hour. The products were

purified using gel electrophoresis (see Section 2.1.3) followed by gel extraction of

the desired bands using the QIAquick gel extraction kit (Qiagen), as per

manufacturers instructions.

Page 105: Discovery of antimicrobial peptides active against

Chapter 2

105 | P a g e

Methods

In the following step, the digested and gel-purified pET28a plasmid and

inserts were ligated using the T4 DNA ligase (NEB). The reaction was performed

with 3:1 vector to insert ratio and carried out according to manufacturer’s

instructions in a 50µl reaction volume at room temperature. The following formula

was used to calculate the appropriate amount of insert to be used in relation to the

employed vector:

The ligase was heat inactivated at 65°C for 10 minutes and the ligated

product was used in transformation of competent cells. The completed plasmid

construct was transformed into the chemically competent E. coli XL1 Blue cells

were used as host for the transformation. Three transformation reactions were set up;

a) Positive control with pUC18 control plasmid (Agilent Technologies), b) Negative

control with cut but un-ligated vector and without addition of insert and c) Ligated

experimental construct.

The transformation procedure was also carried out according to the

manufacturer’s instructions and the transformation mixture was plated onto a Luria

agar base containing 50µg/ml kanamycin (LB/Kan) and incubated overnight at 370C

aerobically. Colonies that developed on the antibiotic selection plates were tested for

presence of the insert with colony PCR. Reactions were carried out using the insert

specific primers. Colonies were picked off the agar surface using a pipette tip and

suspended in 50μl of dH2O (DNA and RNA free) (Fisher). DNA was then extracted

by heating the samples to 95°C for 5 minutes in a heating block to lyse the cell and

deactivate enzymes. Following heat treatment, samples were spun briefly in a

Page 106: Discovery of antimicrobial peptides active against

Chapter 2

106 | P a g e

Methods

microfuge to collect any condensation. Collected samples were used as template for

the PCR reaction (see Section 2.5.3).

The size of the insert was confirmed using pmnA gene cloning primers

against pmnA amplified from the native B. pumilus J1 genome and the insert

orientation was confirmed using pmnA forward primer and T7 reverse primer.

Results were visualised by gel electrophoresis (see Section 2.1.3).

The DNA sequence of the recombinant pmnA gene from selected constructs

was confirmed by sequencing from T7/pET sequencing primer sites (Table 2.2)

following PCR using Biomix Red mastermix (Bioline). The amplicons were

sequenced by the staff of Manchester University Sequencing Services (Stopford

building) (see Section 2.1.2).

2.5.2 In vitro cell-free expression

In vitro expression was carried out using the PURExpress® In Vitro Protein

Synthesis Kit (NEB). The experiment was carried out using the manufacturer

provided method to express from the pET28a-pmnA plasmid. The assay contains all

necessary reagents and substrates required for the synthesis of the desired peptide

encoded within a T7 promoter regulated operon and omits the use of bacterial cells,

which eliminates most of the complications associated with recombinant expression.

Manufacturer provided DHFR plasmid was used as a control; this expresses

dihydrofolate reductase enzyme, which can be easily monitored using SDS-PAGE

gel electrophoresis to indicate successful performance of the protocol. The assay

produces analytical amounts of peptide that were easily purified using HisPur nickel-

nitrilotriacetic acid (Ni-NTA) agarose beads (Thermoscientific) as all components

Page 107: Discovery of antimicrobial peptides active against

Chapter 2

107 | P a g e

Methods

involved in the synthesis process are supplied with an attached histidine tag (His-

Tag).

2.5.3 Proteomic analysis of peptide products generated following expression

studies

Peptide expression was tested using the well diffusion assay and SDS-PAGE

gel electrophoresis (see Sections 2.1.4.3 and 2.2.2). Prior to analysis, to reduce the

total peptide content and increase the resolution, the cell free expression products

were subjected to 100kDa centrifugal filtration using Amicon ultra 0.5 ml centrifugal

filters (Merck Millipore, Hertfordshire, UK). Analysis of the filtered cell free

expression products were conducted using both un-purified products and Ni-NTA

agarose bead purified (as per manufacturer’s instructions) peptide.

Part of the peptide mixture obtained was used for well diffusion assays (see

Section 2.1.4.3; 50μl/well) to check for antimicrobial activity. The remaining lysate

was used for SDS-PAGE analysis (see Section 2.2.2) to investigate the recombinant

peptide yield and size.

Page 108: Discovery of antimicrobial peptides active against

Chapter 3

108 | P a g e

Introduction

3 CHAPTER 3

PUMICIN NI04, A NOVEL ANTIMICROBIAL PEPTIDE FROM

BACILLUS PUMILUS, IS HOMOLOGOUS TO THE ESXA VIRULENCE

DETERMINANT

Page 109: Discovery of antimicrobial peptides active against

Chapter 3

109 | P a g e

Introduction

3.1 INTRODUCTION

It only took two years for the first meticillin resistant strain of Staphylococcus

aureus to emerge after introduction of penicillinase stable β-lactams in 1951 (Barber,

1961; Chambers and Deleo, 2009). Although improved infection control and

hygiene interventions have led to a significant decrease in infections caused by

meticillin resistant S. aureus (MRSA), it is still one of the most prominent antibiotic

resistant pathogens encountered in hospital and community infections and is a

member of the ESKAPE group (Boucher et al., 2009). In fact, antibiotic resistance

has become one of the most pressing challenges that modern medicine faces, with

resistance traits accumulating and spreading rapidly among bacterial pathogens

(Barriere, 2015; Cosgrove, 2006; Jean and Hsueh, 2011; Livermore, 2007; Martínez

and Baquero, 2014). This challenge must be met urgently since introduction of

antibiotics has not only decreased human morbidity and mortality resulting from

bacterial infections, but as prophylactics their use allows many modern medical

procedures to be performed; the reality of a post antibiotic era cannot be easily

imagined (Rice, 2008). Rapid resistance acquisition or development is fuelled in part

by misuse of antibiotics and especially high selective pressure is applied in hospitals.

This is further complicated by rapid dissemination of antibiotic resistant pathogens

through advancing global transport (Hede, 2014; Hsueh, 2012; W. Li et al., 2014;

Zhang et al., 2006). In tandem, the antibiotic discovery pipeline is poorly populated,

while the time taken by bacteria to develop resistance towards new antibiotics

decreases (W. Li et al., 2014; Livermore, 2011).

Page 110: Discovery of antimicrobial peptides active against

Chapter 3

110 | P a g e

Introduction

Although the outlook is considerably less bleak for infections caused by the

Gram positive bacteria with the introduction of five new in classes of antibiotics

since 2000, therapy of infectious diseases still relies heavily on use of derivatives of

existing antibiotics, and has done so since the 1970s (Butler et al., 2013). However,

the number of derivatives that can be developed from existing classes is finite and

further development grows more expensive (Coates et al., 2011). Worryingly,

resistance is already emerging towards some recently introduced drugs, including

daptomycin and linezolid (Cavalcante et al., 2014; Fessler et al., 2014; McElvania

TeKippe et al., 2014; Montero et al., 2008).

A supply of antibiotics from new classes, with novel modes of action is

required more consistently. Bacteriocins, bacterially derived antimicrobial peptides,

are important candidates, not only for their potential medical applications but the

wealth of knowledge that we could learn via studying these antimicrobials developed

through the sieve of evolution (Walsh and Fischbach, 2010).

Bacteriocins are generally active against species closely related to the

producing bacteria (the producers) and have unique structures, varied modes of

action mediating rapid killing and the fact that they are ribosomally encoded allows

the native peptides to be rationally modified to improve efficacy and stability or

reduce toxicity (Abriouel et al., 2011; Cotter et al., 2005; Godballe et al., 2011).

Bacteriocins are also observed to be affected less by cross resistance with other

antibiotics, although limited evidence is emerging to support the suggested

occurrence of cross resistance, even if it is indirectly through promotion of bacterial

competition (Koch et al., 2014).

Page 111: Discovery of antimicrobial peptides active against

Chapter 3

111 | P a g e

Introduction

Another factor compounding the deficiency in antibiotic discovery is the lack

of success with the genetic approaches that were adopted to replace time consuming

and laborious natural product screening (Livermore, 2011; Newman, 2008).

However, recent improvements in screening and discovery technologies, especially

the advancement of rapid dereplication methods that are able to work with small

sample volumes, such as high resolution mass spectrometry (eg. Orbitrap) has

reignited interest in both bacteriocin research and in wider aspects of natural product

discovery. Most recently, innovative techniques have been developed that allow

culture and screening of previously uncultivable bacteria and introduction of the

iChip device led to the discovery of the non-ribosomal peptide teixobactin, a potent

first in class antibiotic (Ling et al., 2015). Such innovations will help promote the

revolution in natural product discovery.

In the present study, by taking advantage of these new technologies, we have

characterised a potent inhibitor of Gram positive bacteria, peptide NI04, which is one

of two bacteriocins recovered from an environmental isolate, Bacillus pumilus strain

J1. This unusual and effective antimicrobial peptide has shown particular similarities

to the Esx-A protein, which is described as a significant virulence factor in other

bacteria including Mycobacterium tuberculosis and some species of S. aureus (Burts

et al., 2005; De Jonge et al., 2007; Sundaramoorthy et al., 2008). However, B.

pumilus is not known for its pathogenicity towards humans and, in our toxicity tests,

no adverse effects were observed. Our addition to the understanding of EsxA-like

peptides indicates that this family of peptides may have evolved in a divergent

manner to exploit membrane-damaging activity for different means. The second

active agent that was isolated from B. pumillus strain J1 was peptide NI03, which is

active against both Gram positive and negative species. However, due to suboptimal

Page 112: Discovery of antimicrobial peptides active against

Chapter 3

112 | P a g e

Introduction

production and stability concerns, detailed study of this peptide was not conducted,

though some promising data relating to NI03 were recovered following in silico

studies conducted during analysis of the producer’s genome (See Chapter 5).

Page 113: Discovery of antimicrobial peptides active against

Chapter 3

113 | P a g e

Results

3.2 RESULTS

3.2.1 Two antimicrobial agents are produced by B. pumilus J1

From the environmental samples collected for analysis in this study, sixty-

eight individual colonies were isolated and of these, ten showed strong antimicrobial

activity. Among these, strain J1, isolated from a settle plate, showed striking

antimicrobial activity towards Gram positive bacteria and also, to a lesser degree,

towards Gram-negative species and thus strain J1 was chosen for detailed

examination. The 16S rRNA gene sequence data revealed that this bacteria was a

member of B. pumilus species and it was designated as B. pumilus strain J1.

Following identification of this strain as a producer of antimicrobial activity, and

preliminary investigation of the inhibitory spectrum of the strain, efforts were made

to purify the active agent(s) from liquid culture.

Following Strata-XL and then Sep-Pak C18 fractionation of supernatants

from TSB broth cultures, two antimicrobial agents with different activity spectra

were isolated from the producer organism B. pumilus J1. The first inhibitor, which

eluted in 40% ACN+0.01TFA fractions (Figure 3.1) was named peptide NI03 and

was active against Gram positive and negative indicator strains.

Page 114: Discovery of antimicrobial peptides active against

Chapter 3

114 | P a g e

Results

Figure 3.1 Spot assay performed with Sep-Pak C18 purified fractions against S. aureus. Peptide NI03 eluted at 40% acetonitrile concentration and peptide NI04 at 60 and 70% acetonitrile fractions.

Peptide NI04 was active against only Gram positive bacteria, except the

Gram negative bacterium N. gonorrhoea (Table 3.1). This peptide eluted in 60 and

70% ACN+0.01TFA fractions (Figure 3.1). However, peptide NI03 was not

subsequently present in all purification batches and did not respond to optimisation

studies. Thus, efforts were concentrated on the identification of peptide NI04 and

NI03 was not progressed.

Table 3.1 Antimicrobial activity of peptides NI03 and NI04 against indicator species. Results were obtained using the spot on lawn assay with Sep-Pak C18 fractions. Activity was recorded as (+) if a visible inhibition zone was present and (-) if no zone of inhibition was present.

Gram negative indicator strains NI03 NI04

Gram positive indicator strains NI03 NI04

E. coli 414 + - B. subtilis - - E. coli DH5a + - C. difficile + + ESBL E. coli M2 + - C. diphtheriae + + ESBL E. coli 169 + - M. luteus + + K. aerogenes - - S. aureus 308 + + K. pneumoniae 169 - - S. aureus 1195 + + N. gonorrhoea - + MRSA + + P. aeruginosa 257 - - VR Enterococci + + Salmonella Typhimirium

+ -

Page 115: Discovery of antimicrobial peptides active against

Chapter 3

115 | P a g e

Results

3.2.2 Bacteriocin production can be optimised by manipulation of growth

conditions

A variation in the level of recovery of peptide NI04 was observed early on in

the project. To address these variations, an optimisation study was performed using

various media and additives and two incubation temperatures, over a 24 hour period.

Peptide concentration was measured in AU/ml with the well diffusion assay using

the indictor organism M. luteus. Observations revealed that TSB and NB produced

the highest yields of peptide NI04, while the use of TSB supplemented with YE (1%

w/w), lactose (1% w/w) and maltose (1% w/w) had a negative effect on the yield

(Table 3.2). Sucrose supplementation (1% w/w) of TSB YE, while having no

negative effects did not enhance the production compared with basic TSB. BHI

medium was also revealed to be suboptimal for peptide NI04 production (Table 3.2).

Table 3.2 Effect of various media and additives on the production and availability of peptide NI04, following overnight incubation at 37oC.

Broth Additives Peptide NI04

(AU/ml) TSB* - 640 TSB YE* (1%) 320 TSB YE (1%), Lactose (1%) 320 TSB YE (1%), Sucrose (1%) 640 TSB YE (1%), Maltose (1%) 320 TSB Sucrose (1%) 640 BHI* - 320 NB* - 640

* TSB-(Tryptic Soy Broth), YE-(Yeast extract), BHI-(Brain heart infusion broth), NB-(Nutrient broth)

Page 116: Discovery of antimicrobial peptides active against

Chapter 3

116 | P a g e

Results

Following on from these results, a time-controlled experiment was performed

assessing the effect of incubation temperature and time on the quantity of peptide

available. Overall, TSB proved to be the most consistent medium, reaching the

highest observed yield of 2560AU/ml at both 30oC and 37oC (Figure 3.2). However,

peptide NI04 availability was also influenced by incubation time, with the amount of

recoverable peptide declining significantly after 18 hours when B. pumilus J1 was

incubated at 370C (Figure 3.2). Optimal production was seen between 12 and 18

hours under all test conditions.

Figure 3.2 Availability of peptide NI04 in tryptic soy broth and nutrient broth culture media under different incubation temperatures over a 24 hour period. Error bars represent the standard deviation in peptide NI04 activity between replicates.

Page 117: Discovery of antimicrobial peptides active against

Chapter 3

117 | P a g e

Results

3.2.3 De novo peptide sequence determination of peptide NI04 using mass

spectrometry analysis and genome interrogation facilitates identification

of the pumicin NI04 locus

Active Sep-Pak C18 fractions that eluted at 60% and 70% ACN+0.01%TFA

were pooled and applied to RP-HPLC column. Active fractions obtained from the

RP-HPLC eluted at ~ 45% acetonitrile concentration (Figure 3.3) and were subjected

to ESI-MS analysis. Mass spectrometric evaluation of the RP-HPLC purified peptide

NI04 suggested a mass of 10722.993Da (Figure 3.4). The detected mass was queried

against online bacteriocin data-bases and published literature for de-replication

purposes. The query search resulted in no matches, indicating that peptide NI04 had

a novel mass.

Figure 3.3 RP-HPLC chromatogram of peptide NI04. Each peak on the graph represents the molecules or molecule groups eluted during the RP-HPLC procedure at different acetonitrile concentrations (green line indicates the acetonitrile concentration). The readings are taken at 215nm UV intensity. Active fractions are labelled.

Active fraction H13

Page 118: Discovery of antimicrobial peptides active against

Chapter 3

118 | P a g e

Results

Figure 3.4 ESI-MS scan of active HPLC fraction showing the detected mass of peptide NI04 as 10722.993. Other peaks represent background noise and a possible impurity with a mass of 10771.341 (not present in other spectrographs).

MS analysis was followed by de novo peptide sequencing using MS/MS

fragmentation. The unique sequence tags generated from the raw data using the

MASCOT software and manual investigations are listed in Table 3.3. These peptide

tags can be used to interrogate the bacterial genome to facilitate the discovery of the

gene(s) responsible for the production of the AMP. Thus, the Draft genome of B.

pumilus J1 was generated using the illumina MiSeq platform.

Table 3.3 The unique sequence tags for peptide NI04 obtained from the de novo peptide sequencing efforts. Common modifications that can occur were also accounted.

Sequence Length Modifications QYGVESQEVLNQVDR 15 Unmodified MSDLLTDVSNQLDQTANTLESTDQDIASQIR 31 Unmodified GMWEGASSEAFADQYEQLKPSFIK 24 Oxidation (M) MISDLK 6 Unmodified MSGIIR 6 Acetyl (Protein

N-term)

Page 119: Discovery of antimicrobial peptides active against

Chapter 3

119 | P a g e

Results

The draft genome sequence was assembled from a total of 2,998,202 raw

sequence reads having an average length of 62 bp. Assembly resulted in 1490

contigs comprising a draft genome of 4,421,862 bp [N50 = 9581 (50% of the genome

was contained in contigs of greater than 9,581 bp)]. From the draft data, it was

estimated that average genome coverage was 193 fold. The genome was annotated

using the PROKKA genome annotation software. The B. pumilus SAFR-32 stain

was used as the reference during the assembly and annotation. The differences

between the genomes of the B. pumilus J1 and the reference strain are graphically

represented in the diagram generated using the BRIG software (Figure 3.5).

Figure 3.5 A graphical representation of the differences between the genomes of the B. pumilus SAFR-32 strain (inner ring) and B. pumilus J1 (outer, purple ring). Comparisons were made with BRIG software.

Page 120: Discovery of antimicrobial peptides active against

Chapter 3

120 | P a g e

Results

Sequence tags generated from the MS/MS fragmentation data were mapped

back to the draft bacterial genome, both manually using the BLAST facility in CLC

Main Workbench and using the built in search features of PEAKS studio 7 and

MASCOT software. This allowed identification of the peptide NI04 amino acid

sequence as

-MSGIIRVTPEELRATAKQYGVESQEVLNQVDRLNRMISDLKGMWEGASSE

AFADQYEQLKPSFIKMSDLLTDVSNQLDQTANTLESTDQDIASQIRG-

Sequence tags generated that helped to identify the sequence are highlighted

in different colours on the peptide sequence with a predicted mass of 10854.207.

This differed from the observed mass by 131.214Da, which implies cleavage of the

N-methionine from the amino acid sequence (Frottin et al., 2006).

The encoding gene sequence in the B. pumilus J1 genome was annotated as

esxA using the PROKKA annotation tool. The annotation was based on an observed

30.7% identity with a virulence factor from S. aureus, EsxA, which is secreted

through a specialised T7SS pathway called the ESX (Ess/Yuk) pathway. Another

EsxA virulence factor that shared homology with peptide NI04 was Mycobacterium

tuberculosis EsxA, which is secreted by the ESX-1 pathway. However, further

analysis using the PSI-BLAST algorithm revealed that peptide NI04 shared most

identity with another EsxA homolog observed in B. subtilis, namely YukE (91.7%)

(See, Figure 3.6 and Figure 3.7). The function of YukE is unknown, although its

secretion mechanism has been studied (Baptista et al., 2013; Huppert et al., 2014;

Sysoeva et al., 2014).

Page 121: Discovery of antimicrobial peptides active against

Chapter 3

121 | P a g e

Results

EsxA peptides are classified in the WXG protein superfamily (Pfam:

PF06013), which encompasses proteins of approximately 100 amino acids with

various functions including virulence, cell wall maintenance and sporulation (Akpe

San Roman et al., 2010; Ates et al., 2015; Fyans et al., 2013; Garces et al., 2010;

Huppert et al., 2014; Sundaramoorthy et al., 2008). We propose that this novel

antimicrobial peptide be named pumicin NI04 and that it is encoded by pmnA.

Further analysis of the gene locus around pmnA also revealed the presence of other

T7SS pathway components such as a predicted ATPase (PROKKA 03052) sharing

70.8% identity with YukBA (UniProt ref: C0SPA7) that is essential for substrate

secretion via the B. subtilis ESX (Yuk) pathway (Huppert et al. 2014). In addition,

homologs of other Yuk loci transport and accessory proteins were also uncovered in

the pumicin NI04 biosynthesis cluster; YukD (UniProt ref: P71071) (73% identity

with YukD), YukC (UniProt ref: P71070) (100% identity), PROKKA 03051 (45.9%

identity with YueB (UniProt ref: O32101), PROKKA 03050 (35% identity with

YueC (UniProt ref: O32100) (Figure 3.8) (Huppert et al., 2014).

Page 122: Discovery of antimicrobial peptides active against

Chapter 3

Results

122 | Pa

ge

Figure 3.6 Protein sequence alignment of pumicin NI04 and the 10 most homologous peptides identified using the PSI-BLAST algorithm (Altschul et al., 1997). Also included are the known EsxA peptide sequences (EsxA [M. tuberculosis] and virulence factor EsxA [S.aureus]) that are experimentally confirmed to play a role in the virulence of their respective host organisms (boxed in Green) and the highly similar EsxA homolog YukE peptide of unknown function from B. subtilis (boxed in orange). Secondary structure prediction, performed using jnetpred, is illustrated at the bottom of the alignments and confirms the helix turn helix structure of WXG-100 peptides is preserved in pumicin NI04 together with the W-X-G motif.

WXG motif

Page 123: Discovery of antimicrobial peptides active against

Chapter 3

Results

123 | Pa

ge

Figure 3.7 Neighbour joining tree calculated using the % identity of aligned WXG family peptides. Branches are labelled with the % difference in the identity of the amino acid sequences to each other.

Page 124: Discovery of antimicrobial peptides active against

Chapter 3

Results

124 | Pa

ge

Figure 3.8 Organisation of the biosynthetic cluster predicted to be involved in the production and secretion of pumicin NI04 encoded in the producer B. pumillus J1 genome and organisation of other operons encoding homologous EsxA peptides, M. tuberculosis ESX1 locus (Burts et al., 2005), B. subtilis Yuk locus (Huppert et al., 2014) and S. aureus Ess locus (Kneuper et al., 2014). Esx substrates are highlighted in red and proteins that are known to be involved in regulation and secretions of ESX substrates are coloured blue, proteins that are predicted to be involved in transport or regulation of these factors are highlighted in green, proteins that have unknown or unrelated function are coloured grey (Burts et al., 2005; Kneuper et al., 2014; McLaughlin et al., 2007; Ramsdell et al., 2015). PROKKA 03052 (70.8% identity with YukBA [UniProt ref: C0SPA7] ), YukD (73% identity with YukD [UniProt ref: P71071]), YukC (100% identity [UniProt ref: P71070]), PROKKA 03051 (45.86% identity with YueB [UniProt ref: O32101]), PROKKA 03050 (33.5% identity with YueC [UniProt ref: O32100]).

Page 125: Discovery of antimicrobial peptides active against

Chapter 3

125 | P a g e

Results

Further confirmation was achieved by demonstrating that the identified

peptide had antimicrobial activity and the expected mass using SDS-PAGE gel

electrophoresis with an agar overlay (Figure 3.9). A clear zone of inhibition was

present around the band of interest. Analysis of the purified peptide was performed

in triplicate and revealed a band visible at the expected interval between the 6kDa

and 14kDa protein standards used.

Figure 3.9 The SDS-PAGE gel analysis of the Sep-Pak C18 purified samples, against SeeBlu 2 protein marker. Gel overlay assay performed following analysis demonstrated the peptide band at the expected mass interval has inhibitory activity. The band indicated with the orange arrow was used for MS/MS fragmentation and de novo amino acid sequencing.

3.2.4 Pumicin NI04 is stable to heat treatments and protease degradation

confirms a proteinaceous nature

The proteinaceous structure of pumicin NI04 was confirmed using enzyme

stability assays (Table 3.4). As expected, proteolytic enzymes such as protease,

pepsin and proteinase K abolished the antimicrobial activity, while interestingly

trypsin had reduced but not neutralised the activity and α-chymotrypsin had no

effect. Lipid active lipase and carbohydrate active amylase enzymes had no visible

effect either. Pumicin NI04 was also observed to be highly heat stable; exposure to

800C for up to 4 hours had no noticeable effect on the antimicrobial activity (data not

shown).

Page 126: Discovery of antimicrobial peptides active against

Chapter 3

126 | P a g e

Results Table 3.4 Effect of a wide array of hydrolytic enzymes on the antimicrobial activity of pumicin NI04.

Enzyme Retained antimicrobial activity

Protease 0%

Pepsin 0%

Proteinase K 0%

Trypsin 25%

α-Chymotrypsin 100%

Lipase 100%

α-Amylase 100%

3.2.5 Pumicin NI04 has potent activity against resistant Gram positive

pathogens

Pumicin NI04 was shown to have activity against a broad spectrum of Gram

positive bacteria (Table 3.5). Two sets of results are reported in the table below, in

one column the AU/ml value is reported where a higher value represents a higher

level of inhibitory activity and in the second column the actual MIC in µg/ml is

recorded. Low MIC/high activity values were recorded for multiple indicator

organisms including Streptococcus pneumoniae, S. aureus and antimicrobial

resistant species such as MRSA, but VRE species, especially strain 20, were

observed to be surprisingly more susceptible to pumicin NI04, compared with other

Gram positive bacteria tested.

Page 127: Discovery of antimicrobial peptides active against

Chapter 3

127 | P a g e

Results Table 3.5 Minimum inhibitory concentration (MIC) of pumicin NI04 against a wide selection of Gram positive bacteria. MIC was calculated in AU/ml and µg/ml in a selection of indicators using highly purified peptide.

Indicator organism* MIC AU/ml MIC µg/ml

M. luteus 2560 10

S. aureus 308 320 20

S. aureus 1195 320 20 MSSA 27 107 ND MRSA 6 160 40 MRSA 13 213 ND MRSA 37 133 ND MRSA 226 107 ND VRE 16 213 ND VRE 17 160 ND VRE 18 320 20 VRE 19 160 ND VRE 20 2560 ND VRE 21 160 ND Strep. pneumoniae 1 427 20 Strep. pneumoniae 2 640 ND Strep pneumoniae 4 160 ND Strep. pneumoniae 6 640 ND Strep. pneumoniae 7 107 ND Strep. pneumoniae 11 320 ND

3.2.6 UV absorbing material leakage assay

The UV absorbing material leakage experiments were performed to gain

insight into the mode of action of pumicin NI04. The results revealed that soon after

the addition of pumicin NI04 to M. luteus cells, there was a steady release of UV

absorbing materials from the M. luteus cells; at the end of three hours there was a

0.047OD change in the measurable UV absorbing materials (Figure 3.10).

*-Meticillin sensitive S. aureus (MSSA), Methicillin resistant S. aureus (MRSA) Vancomycin resistant Enterococcus faecium (VRE), ND: Not determined due to low quantity of purified peptide

Page 128: Discovery of antimicrobial peptides active against

Chapter 3

128 | P a g e

Results

Observations obtained from S. aureus cells were slightly different; instead of

a swift and constant release of UV absorbing particles, a delayed but rapid release

was observed. Also, the overall recorded difference in the absorbance was higher for

S. aureus, with a reading of 0.099 OD obtained at 4 hours (Figure 3.10).

Page 129: Discovery of antimicrobial peptides active against

Chapter 3

Results

129 | Pa

ge

Figure 3.10 Release of UV absorbing materials from A) M. luteus and B) S. aureus cells over time following treatment with pumicin NI04 at 1xMIC of the test organism compared with pumicin NI04 solution (1xMIC) in test buffer without any microbial inoculum to account for the absorbance inferred by pumicin NI04.

A

B

Page 130: Discovery of antimicrobial peptides active against

Chapter 3

130 | P a g e

Results

3.2.7 Pumicin NI04 has low levels of in vitro toxicity

In haemolysis assays, no haemolytic activity was recorded, irrespective of the

concentration of pumicin NI04 tested (Table 3.6). Dilutions were made from peptide

concentrations of 18x the MIC for M. luteus.

Table 3.6 Percentage of haemolysis that occurred following the addition of pumicin NI04 to mouse erythrocytes. Triton-X100 detergent was used to achieve absolute haemolysis and was used as the positive control.

Pumicin NI04 concentration (AU/ml) 256

128

64

32

16 Triton-X 100 (2%)

% Haemolysis 1.63 0.90 1.08 1.08 0.90 100

% STDEV 1.73 0.36 0.55 0.85 0.54 -

Promising results were also obtained in toxicity studies performed against

human keratinocytes and monkey Vero cells using peptide concentrations of up to

18x the MIC for M. luteus. In the neutral red assay, no difference was perceived in

the ability of the Vero cells to incorporate neutral red into their lysosomes, regardless

of the concentration of pumicin present in the culture medium (Figure 3.11).

Assessment of keratinocyte cell viability data using the trypan blue exclusion

assay also indicated that the presence of pumicin NI04 up to 18x M. luteus MIC had

no toxic effect (Figure 3.12). In addition, during the microscopic examination of

both cell lines following treatment, no visible abnormalities were detected between

the treated and untreated cultures. Total cell destruction occurred in the wells

containing either SDS or Triton X-100.

Page 131: Discovery of antimicrobial peptides active against

Chapter 3

131 | Pa

ge

Results

Figure 3.11 Comparison of neutral red uptake of Vero cells against negative (growth media) and positive (2% triton X-100) controls, following incubation with differing concentrations of pumicin NI04 of up to 18x the MIC recorded against M. luteus.

Figure 3.12 Cell survival data collected using the trypan blue exclusion assay, following a 24 hour incubation of keratinocyte cells with pumicin NI04 supplemented medium against negative (growth media) and positive (20mg/ml SDS) controls.

Page 132: Discovery of antimicrobial peptides active against

Chapter 3

132 | P a g e

Discussion

3.3 DISCUSSION

Using a natural product screening effort, B. pumilus strain J1 was isolated from

settle plates placed around the Manchester Royal Infirmary Clinical Sciences

Building. The strain produced two antimicrobial peptides with distinct spectra of

activity, which eluted at two different solvent concentrations from Sep-Pak C18

columns. Peptide NI03 demonstrated antagonism towards both Gram positive and, to

a lesser degree, towards Gram-negative species. However, this peptide was only

expressed in adequate quantities in a small number of fermentation batches. It is well

documented that expression of some bacteriocins is highly dependent on culture

conditions, however, optimisation studies performed failed to improve the

consistency and the yield and it is likely a more tailored approach specific to peptide

NI03 is required, such as removal of key media ingredients or addition of

components from the natural niche of B. pumilus in an effort to ‘elicit’ production of

the peptide (Lincke et al., 2010; Rigali et al., 2008; Zhu et al., 2014). Peptide NI03

is discussed in more detail later in this thesis (see Chapter 5, Sections 5.3.3 and 5.5).

In contrast, peptide NI04 was active against a wide variety of Gram positive

bacteria and, although its expression was also erratic, it was recoverable in good

yields following optimisation of growth media and incubation time. Thus, until

more information can be gathered regarding peptide NI03 that may guide

optimisation, it is not an efficient use of resources to pursue this inhibitor. A decision

was made to focus efforts on recovery and characterisation of peptide NI04. As a

result, the novel antimicrobial peptide pumicin NI04 has been identified. The peptide

was observed to be proteinaceous in structure, heat stable and have a mass of

10722.993Da. De-novo sequencing of the active peptide using MS/MS helped to

Page 133: Discovery of antimicrobial peptides active against

Chapter 3

133 | P a g e

Discussion

trace the sequence back to a gene identified as an esxA homolog in the draft genome

of the producer B. pumilus strain J1. The esxA homolog identified in the genome

encoded a peptide with a predicted mass of ~10854.207, which differed from the

observed mass by 131.214Da. This mass difference suggests cleavage of the N-

terminal methionine from the peptide sequence by the bacteria that causes 131Da

mass shift (Demirev et al., 2001; Frottin et al., 2006). One of the peptide tags

generated by the MASCOT driven de novo peptide sequencing, also suggested that

pumicin NI04 contains an N-terminal acylation which may occur prior to or after the

cleavage of N-terminal methionine but the expected 42Da mass shift wasn’t

observed on the MS analysis of the whole peptide (Arnesen, 2011; Helbig et al.,

2010; Parker et al., 2010). Nevertheless, it is important to note that the EsxA

peptides of mycobacteria are N-α-terminally acetylated (Mba Medie et al., 2014;

Okkels et al., 2004).

EsxA peptides belong to the WXG 100 superfamily of peptides. These are

peptides released into the environment by Gram positive bacteria through the ESX

(ESAT-6/Ess/Yuk) secretion system, an ortholog of the T7SS dedicated to secretion

of ESX substrates, such as EsxA, EsxB, EsxC and YukE (Burts et al., 2005; Huppert

et al., 2014; Kneuper et al., 2014). The WXG 100 superfamily is constituted of

proteins with approximately 100 amino acids and are identified by their helix-turn-

helix structure and conserved W-X-G motif (Pallen, 2002). These properties are

found in pumicin NI04 (Figure 3.6). EsxA peptides are generally associated with

virulence in organisms such as Mycobacterium tuberculosis, S. aureus and B. subtilis

(Burts et al., 2005; De Jonge et al., 2007; Sundaramoorthy et al., 2008). However,

there are exceptions, as demonstrated by the plant pathogen Streptomyces scabies.

Fyans and colleagues witnessed that abrogating EsxA production in this organism

Page 134: Discovery of antimicrobial peptides active against

Chapter 3

134 | P a g e

Discussion

didn’t affect the pathogenicity, but led to deficits in the development of the bacterial

cell (Fyans et al., 2013). Furthermore, presence of ESX genes in non-virulent

bacterial strains, such as Streptomyces coelicolor provides proof that these peptides

are not obligatory virulence factors (Akpe San Roman et al., 2010; Converse and

Cox, 2005; Huppert et al., 2014). In addition, B. pumilus, although a pathogen of

fish, like many other bacillus species rarely causes infections in humans (Bentur et

al., 2007; Tena et al., 2007).

Possible functional divergences amongst EsxA homologs can be explained by

the fact that, although strong similarities in predicted structures are present

between EsxA peptides, there are also considerable variations. In an iterative study,

Poulsen and co-workers managed to group a selection of WXG 100 proteins into

three categories, namely CFP-10 (10kDa culture filtrate protein) often referred to as

EsxB, ESAT-6 (6kDa early secreted antigenic target) and sagEsxA like. The latter

two categories are often referred to as EsxA (Poulsen et al., 2014).

A homolog of the identified antimicrobial peptide pumicin NI04 is present

within the previously described peptides; BPUM2860 (UniProt ref: A8FGZ8) groups

together with the sagEsxA like category (Poulsen et al., 2014). These WXG 100

proteins are encoded by mono-cystronic genes and diverge from the CFP-10 and

ESAT-6 like proteins that are encoded by bi-cistronic operons (Poulsen et al., 2014).

CFP-10 and ESAT-6 are mutually dependent and form heterodimers, while sagEsxA

like peptides form homo-dimers explaining the absence of EsxB (CFP-10) from the

pumicn NI04 operon (Poulsen et al., 2014). The homo-dimeric nature of these

peptides is also supported by the findings of Sysoeva and colleagues (2014), where

they suggested recognition of the YukE homo-dimer facilitates the secretion of this

Page 135: Discovery of antimicrobial peptides active against

Chapter 3

135 | P a g e

Discussion

peptide. It is also suggested that these dimers are secreted as a single unit (Baptista et

al., 2013; Sysoeva et al., 2014). However, in the current study there was no evidence

from the MS studies or SDS-PAGE analysis that would imply presence of dimers of

pumicin NI04 in the purified fractions, although they were not actively sought. In

addition, within each category including the sagEsxA group there are considerable

variations in the amino acid sequences between members, as noted by Poulsen et al.

(2014). This can also be observed in Figure 3.6. Without experimental evidence, it is

not possible to comment on the specific effect of the observed sequence variations

amongst the EsxA peptide family, regarding their general function or between B.

subtilus YukE and pumicin NI04. However, the deduction that can be made from the

aforementioned results of this study, combined with the functional diversity

observed between EsxA like peptides in the literature and the presence of these

proteins in non-virulent bacteria (Akpe San Roman et al., 2010; Ates et al., 2015;

Fyans et al., 2013; Garces et al., 2010; Huppert et al., 2014; Sundaramoorthy et al.,

2008), indicate that these proteins may have evolved divergently to fulfil different

functions within the particular bacterial niche being inhabited. There has been no

experimental evidence published showing antimicrobial activity of EsxA like

peptides or other peptides belonging to the WXG 100 superfamily, making pumicin

NI04 unique in this respect. It would be logical to assess the antimicrobial activity of

pumicin NI04 homologs from other Bacillus species including YukE. Also, in the

absence of reports where such assays have been performed on members of

Mycobacterium or Staphylococcus, it would also be interesting to investigate the

wider phenotype of EsxA proteins from these genera.

Page 136: Discovery of antimicrobial peptides active against

Chapter 3

136 | P a g e

Discussion

It is important to note that membrane lysis, which may form the basis of

antibacterial activity in antimicrobial peptides (Destoumieux-Garzón et al., 2003;

Diep et al., 2007; Hobbs et al., 2008; Ooi et al., 2009) (See Sections 1.5.2 and 1.6.2),

has been demonstrated as a mode of virulence employed by M. tuberculosis and S.

aureus and is facilitated through homologs of EsxA peptides, specifically ESAT-6

(EsxA) and CFP-10 (EsxB) (De Jonge et al., 2007; Gao et al., 2004; Guinn et al.,

2004; Hsu et al., 2003; Mba Medie et al., 2014). In contrast, while pumicin NI04

was toxic to bacterial cells, during the in vitro toxicity studies pumicin NI04 was

observed to be non-toxic towards human keratinocyte and monkey Vero cell lines at

up to 18x the MIC values. It is also worth investigating the particular effects that

pumicin NI04 may have on macrophages, considering the lytic mode of action

suggested for EsxA peptides is often described to be directed towards macrophages

(Gao et al., 2004; Mba Medie et al., 2014; Simeone et al., 2012).

The UV absorbing material leakage assay performed in this study is used for

assessing pore formation on the bacterial cell membrane and the results obtained

established the leakage of UV absorbing materials from S. aureus and M. luteus cells

in response to pumicin NI04 (Devi et al., 2010; Eumkeb et al., 2012; Su et al., 2012;

Yildirim et al., 2004; Zhou et al., 2008). These results are in line with the

observation made concerning the membrane lysis action of EsxA like peptides

during virulence. Although the above data suggest that pumicin NI04 is interfering

with the integrity of the bacterial cell membrane, it is not sufficient by itself to draw

final conclusions regarding the mode of action of pumicin NI04 and further studies

must be conducted.

Page 137: Discovery of antimicrobial peptides active against

Chapter 3

137 | P a g e

Discussion

In addition, pumicin NI04 shares properties associated with bacteriocins in

many respects, such as activity against closely related bacterial species (Table 3.5),

heat stability and self-immunity (Cascales et al., 2007; Cotter et al., 2005). It also

proved to be an efficient antimicrobial against antibiotic resistant species such as

MRSA and VRE. However, given the combination of structural and

physicochemical characteristics exhibited by pumicin NI04, it is not possible to

classify it using the commonly accepted bacteriocin classification schema, as NI04

has a mass greater than 10kDa, yet it is heat stable.

In the most recently suggested bacteriocin classification scheme, Heng and

Tagg (2006) combined the widely accepted classification schema from Cotter and

colleagues (2005) and Klaenhammer (1993) into one, dividing the known peptides

into four categories: lantibiotics; unmodified peptides; large proteins; and cyclic

peptides. Granted there is a group allocated to large bacteriocins by the scheme,

but it is defined by heat liability (Heng & Tagg 2006; Klaenhammer 1993) and

includes heat labile large bacteriocins such as Dysgalacticin and Linocin M18.

Therefore, pumicin NI04 cannot be classified to this group (Swe et al., 2009;

Valdés-Stauber and Scherer, 1994), though pumicin NI04 is not the only large heat-

stable antimicrobial peptide; Propionicin SM1 (Miescher et al., 2000), isolated from

Propionibacterium jensenii DF1 is also heat stable, yet it should be noted that this

protein is much larger than NI04 having a mass of 19,942Da.

The only antimicrobial peptide previously identified from members of B.

pumilus is the heat stable bacteriocin, pumilicin 4 from the WAPB4 strain (Aunpad

and Na-Bangchang, 2007). However, due to a lack of structural data, it is not

possible to make a direct comparisons of these AMP’s, though pumicin NI04 is

Page 138: Discovery of antimicrobial peptides active against

Chapter 3

138 | P a g e

Discussion

significantly larger than pumilicin 4, which has a reported mass of 1994.62 Da

(Aunpad and Na-Bangchang, 2007).

Analysis of the gene loci around pmnA suggests that, like other WXG 100

peptides, pumicin NI04 is also secreted by an ESX secretion pathway; though it does

have some sequence variations, it is mainly orthologous to the Yuk pathway of B.

subtilis. In the absence of EsxB, it contains a homolog of the ubiquitin like yukD

gene similar to the B. subtilis Yuk pathway (Figure 3.8). YukD like protein family

members are conserved with T7SS and associated secretion systems and are shown

to have ubiquitin like structures but with a distinctly shorter C-terminal sequence

lacking the GG motif required for covalent bonding with other peptides (van den Ent

and Löwe, 2005). Ubiquitins are mainly eukaryote associated proteins that are

involved in a process called ubiquitination; once an ubiquitin bonds to a peptide it

signals proteolytic degradation or post translational modification of the bonded

peptide (Pickart, 2001; van den Ent and Löwe, 2005). However, as YukD lacks the

GG motif, the exact function of this protein is unknown (Hershko and Ciechanover,

1998; Pickart, 2001; van den Ent and Löwe, 2005). Nevertheless, instead of

virulence like EsxB, YukD was observed to be somehow involved in the export of

YukE from Bacillus subtilis (Burroughs et al., 2011; Burts et al., 2008; Huppert et

al., 2014; Iyer et al., 2006). Nevertheless, the YukD homolog EsaB in the S. aureus

Newman strain Ess pathway, where it’s found together with EsxA, B and C, was

observed to be unnecessary for the export of EsxA and B peptides but it acted post

translationally as a negative regulator of EsxC (Burts et al., 2008; Kneuper et al.,

2014). Thus, due to a lack of consensus function, Kneuper and colleagues (2014)

offered that this peptide is a regulatory peptide that may fulfil different roles

demanded by different encoding strains. Indeed, if it does interact with the pumicin

Page 139: Discovery of antimicrobial peptides active against

Chapter 3

139 | P a g e

Discussion

NI04 in a manner similar to that observed in eukaryotic ubiquitins or as a negative

regulator, it may fulfil the role of an immunity protein in the pumicin NI04 cluster

and thus it would be interesting to observe the effect of its deletion from the pumicin

NI04 cluster in B. pumilus strain J1.

The pumicin NI04 biosynthesis cluster also contains the predicted ATPase,

PROKKA 03052, that shares 70.8% identity with yukBA. ESX secretion pathways

employ members of the FtsK-SpoIIIE ATPase family that contain three ATPase

domains to facilitate secretion of their substrates (Ramsdell et al., 2015), except in

the Ess pathway where EssC mediated secretion requires only one ATPase domain

(Burts et al., 2005; Ramsdell et al., 2015). YukBA is a member of this family that is

translated as a single protein containing all three ATPase domains necessary for

substrate translocation (Huppert et al., 2014; Ramsdell et al., 2015; Sysoeva et al.,

2014). These proteins are also found in two component configurations such as in the

eccCa (1domain) and eccCb (2 domains) encoded ATPase from the Mycobacteria

ESX-1 pathway. The EccCb component in this pathway is essential for interaction

with the EsxB substrate (Champion et al., 2006; McLaughlin et al., 2007; Ramsdell

et al., 2015). Pumicin NI04 ATPase is predicted, by comparison with NCBI's

conserved domain database, to encode three ATPase domains, like YukBA (Figure

3.13) (Marchler-Bauer et al., 2014; Ramsdell et al., 2015). The structure of ESX

ATPases suggest that they are membrane bound and the yueB gene observed in Ess

and Yuk pathways, which is suggested to encode a phage receptor, may be the

membrane associated part of the complex (Huppert et al., 2014; Kneuper et al.,

2014; São-José et al., 2006, 2004). YukC and YueC are also shown to be involved in

secretion of YukE, yet their exact function in the Yuk pathway is still unknown

(Baptista et al., 2013; Burts et al., 2005; Huppert et al., 2014).

Page 140: Discovery of antimicrobial peptides active against

Chapter 3

140 | P a g e

Discussion

Figure 3.13 A representation of the multidomain ATPase involved in ESX substrate secretion. It is believed that the first domain of YukBA generates the energy required for the translocation (dark blue) and the remaining two interact with the substrate (lighter shades of blue) such as pumicin NI04 [from (Ramsdell et al., 2015)] .

In conclusion, pumicin NI04 has proved to be a rather intriguing

antimicrobial peptide with potent activity and high thermal stability, with the closest

homologs having unknown functions or being previously reported to be virulence

factors. This is the first evidence of a peptide belonging to the WXG 100 protein

family displaying antimicrobial activity, suggesting that it is exported from the

producing cell by the ESX secretion system. Although many questions still remain,

one of the most important is whether this peptide is an exception or if there are other

EsxA or WXG 100 peptides that have evolved to have antimicrobial properties.

Pumicin NI04 may have proven to be a potent and non-toxic antimicrobial agent, but

esxA encoded peptides are reported to be highly immunogenic, which may suggest

the possibility of allergic complications if pumicin NI04 were to be applied in vivo,

though the immunogenic properties of pumicin were not investigated. Overall,

pumicin NI04 may not be the immediate answer we were looking for to overcome

antibiotic resistance (in Gram positive bacteria), but it has revealed an interesting

perspective on the evolution and properties of the WXG 100 protein family.

Page 141: Discovery of antimicrobial peptides active against

Chapter 4

141 | P a g e

Introduction

4 CHAPTER 4

DISCOVERY AND ANALYSIS OF PEPTIDE NI05 PRODUCED BY

KLEBSIELLA PNEUMONIAE STRAIN A7

Page 142: Discovery of antimicrobial peptides active against

Chapter 4

142 | P a g e

Introduction

4.1 INTRODUCTION

Gram-negative bacteria have become especially problematic amongst

multidrug resistant pathogens; the fact that four of the six ESKAPE pathogens are

Gram-negative bacteria is just one indication (Rice, 2008). The main reason is that

while there are a number of Gram positive active antibiotics making their way into

the clinic, development is proving more challenging for agents active against Gram-

negative bacteria (Boucher et al., 2009; Cornaglia, 2009; Page and Heim, 2009;

Pucci and Bush, 2013).

All the while, Gram-negative bacteria are getting rapidly more resistant and

treatment options are becoming limited. The high levels of resistance in Gram-

negative bacteria can be mainly linked to the emergence and rapid spread of

extended spectrum β-lactamase enzymes (ESBLs), carbapenemases and other

substrate specific β-lactamases (Cantón et al., 2012; Nordmann, 2014). This rapid

spread can be attributed to the fact that these genes are carried by plasmids that are

efficiently transferable (Cantón et al., 2012; Händel et al., 2015; Kumarasamy et al.,

2010; Rasheed et al., 2013). It is also disturbing that these resistance traits were

observed to spread and be retained amongst normal flora, creating a readily

accessible resistance reservoir to pathogenic species (Kumarasamy et al., 2010; J. W.

Lee et al., 2014; Walsh and Toleman, 2012).

Carbapenemase mediated resistance is one particularly concerning resistance

trait, as these enzymes are active against all β-lactam antibiotics including

carbapenems, against which ESBL enzymes are inactive (Nordmann, 2014). There

are different classes of carbapenemases that confer different levels of resistance to a

broad spectrum of β-lactams, the classification scheme used is called the Ambler

Page 143: Discovery of antimicrobial peptides active against

Chapter 4

143 | P a g e

Introduction

classification (Nordmann, 2014). Particularly significant carbapenemases include the

penicillinase KPC and oxacillinase OXA-48, but the metallo enzyme New Delhi

Metallo-β-lactamase (NDM-1) that confers resistance to almost all known β-lactam

antibiotics including carbapenems is especially concerning (Figure 4.1) (Dortet et

al., 2014; Nordmann, 2014; Robilotti and Deresinski, 2014).

Figure 4.1 Activity spectrum of prominent carbapenemases found in enterobacteriaceae species [adapted from (Nordmann, 2014)].

This is important as imipenem was long considered the gold standard for

treating infections caused by Gram-negative bacteria (Bush, 2010; López-Rojas et

al., 2011). NDM-1 enzymes have been described in all of the four Gram-negative

ESKAPE species, in both individual cases and outbreaks affecting a growing number

of countries including the UK, France, Spain, Italy, India and the USA (Figure 4.2)

(Berrazeg et al., 2014; Bush, 2010; Dortet et al., 2014; López-Rojas et al., 2011;

Nordmann, 2014; Perry et al., 2011; Robilotti and Deresinski, 2014).

Page 144: Discovery of antimicrobial peptides active against

Chapter 4

144 | P a g e

Introduction

Although NDM-1 does not hydrolyse the β-lactam aztreonam, it is commonly

observed that the organisms carrying NDM-1 encode multiple β-lactamases (in some

cases up to eight), where at least one is active against aztreonam (Barguigua et al.,

2013; Bush, 2010; Mitchell et al., 2015). What makes the situation even more

concerning is the co-existence of non-β-lactam antibiotic resistance mechanisms

within most β-lactam resistant Gram-negative pathogens. These are also found in

combinations that confer resistance to most or all non-β-lactam antibiotics (Liu et

al., 2006), giving rise to ‘pan-resistant’ pathogens.

Flouroquinolone and aminoglycoside resistance mechanisms are commonly

found in co-existence with β-lactamase resistance mechanisms (Doumith et al.,

2012; Matsumura et al., 2013). Flouroquinolone resistance is mostly mediated

through the mutation of target enzymes DNA gyrase and DNA topoisomearase IV as

well as various efflux mechanisms (Jacoby, 2005; Ribera et al., 2004).

Figure 4.2 Prevelance of NDM related cases per country (Dortet et al., 2014)

Page 145: Discovery of antimicrobial peptides active against

Chapter 4

145 | P a g e

Introduction

Aminoglycoside resistance is achieved by production of modifying enzymes that

deactivate the aminoglycosides (Bou et al., 2000; Ribera et al., 2004; Van Looveren

and Goossens, 2004). The combination of these mechanisms leave tigecyline (Lam,

2007) and colistin, notable for its toxic nature, as the only treatment options in most

cases (Boucher et al., 2009; Rodríguez-Hernández et al., 2006). However, colistin

resistance is also spreading in strains of A. baumannii and clinical failure of colistin

therapy has been reported by a number of studies, further underlining the urgent need

for novel antibiotic discovery and development (Rodríguez-Hernández et al., 2006;

Telang et al., 2011).

While bacteriocins of Gram positive bacteria present a significant potential

resource to be developed as antibiotics against related species or genera, they lack

the same efficacy against Gram-negative bacteria. Instead, colicins and microcins

present the same opportunity for Gram-negative bacteria (Cotter et al., 2013).

Although potency of microcins as chemotherapeutics is far less studied compared

with their Gram positive cousins, many microcins have been shown to have

significant activity against clinically important pathogens. This includes both narrow

spectrum agents such as MccPDI active against Shigella species and E. coli

(including enterohemorrhagic E. coli O157:H7 and O26 strains) (Eberhart et al.,

2012) and broader spectrum agents such as MccB17 that demonstrates activity

against a collection of Citrobacter, Klebsiella, Salmonella, Shigella and

Pseudomonas species (Cotter et al., 2013). Analysis of microcin C and its analogues

for potential chemotherapeutics produced some promising results and are reviewed

by Vondenhoff and Van Aerschot (2011).

Page 146: Discovery of antimicrobial peptides active against

Chapter 4

146 | P a g e

Introduction

Thus, a screening effort was initiated that targeted inhibitors of Gram-negative

bacteria and yielded a promising candidate isolated from an HVS sample. The

isolated K. pneumoniae strain A7 has been identified to contain a gene cluster

responsible for production of microcin MccE492/G492.

Page 147: Discovery of antimicrobial peptides active against

Chapter 4

147 | P a g e

Results

4.2 RESULTS:

4.2.1 Simultaneous antagonism assays for preliminary analysis of the

antimicrobial activity spectrum of peptide NI05

Peptide NI05 was partially purified from a non-capsulated Klebsiella

pneumoniae strain, which was identified using 16S rRNA gene sequence analysis.

This bacterium was identified from a collection of 96 isolates recovered from

screening plates, which had been inoculated with bacteria grown from HVS samples.

The producer K. pneumoniae strain and the partially purified antimicrobial peptide,

NI05, was significantly active against K. pneumoniae species but was also

particularly efficacious against E. coli strains, including ESBL producers (Table 4.1).

Table 4.1 Peptide NI05 spectrum of activity determined through simultaneous antagonism assays

Gram-negative indicator strains NI05 Gram positive

indicator strains NI05

E. coli 414 +3 B. subtilis -

E. coli DH5a +4 C. difficile -

ESBL E. coli M2 +3 C. diphtheriae -

ESBL E. coli 169 +1 M. luteus -

K. aerogenes - S. aureus 308 -

K. pneumoniae 169 +3 S. aureus 1195 -

N. gonorrhoea - MRSA -

P. aeruginosa 257 +1 VR Enterococci -

Salmonella Typhimirium

+1

*Extended spectrum β-lactamase producing E. Coli (ESBL E. coli), Methicillin resistant S. aureus (MRSA), Vancomycin resistant Enterococcus faecium (VRE)

Page 148: Discovery of antimicrobial peptides active against

Chapter 4

148 | P a g e

Results

No activity was observed against Gram positive species, which is an

interesting contrast to pumicin NI04, a Gram positive active agent that was active

against a Gram-negative bacterium, N. gonorrhoea.

4.2.2 Optimisation of antimicrobial peptide production

Although clear activity was observed during simultaneous antagonism assays

on solid media, production of peptide NI05 in liquid culture was very poor with no

detectable antimicrobial activity following incubation in TSB. The quantities that

achieved inhibition of the E. coli strain DH5α were only attainable through 1% yeast

extract supplementation (Table 4.2). Other additives included were observed to have

no effect. In addition, changing the incubation temperature from 370C to 300C and

increasing incubation time up to 48 hours failed to increase the amount of peptide

that was available. During the time controlled experiment, peptide NI05 reached

quantities detectable by well diffusion assay at 14 hours (data not shown).

Table 4.2 Effect of various media and additives on the production and availability of peptide NI05, following overnight incubation at 37oC.

Broth Additives Peptide NI04 (AU/ml)

TSB* - 0

TSB YE* (1%) 80

TSB YE (1%), Lactose (1%) 80

TSB YE (1%), Sucrose (1%) 80

TSB YE (1%), Maltose (1%) 80

TSB Sucrose (1%) 0

BHI* - 0

NB* - 0

* TSB-(Tryptic Soy Broth), YE-(Yeast extract), BHI-(Brain heart infusion broth), NB-(Nutrient broth)

Page 149: Discovery of antimicrobial peptides active against

Chapter 4

149 | P a g e

Results

4.2.3 Purification

The low abundance of the peptide NI05 in culture media made purification of

this agent more challenging. It prevented vigorous purification as following multiple

chromatography steps, the peptide likely became too dilute and the anticipated active

fractions lost their activity, thus only a single pass of RP-HPLC purification

produced a fraction with detectable antimicrobial activity (20 AU/ml). Better activity

was observed following Strata and Sep-Pak C18 purification (80 AU/ml). The

peptide eluted from the Sep-Pak C18 column at 50% ACN+0.01TFA concentration.

During these studies, an interesting observation was made. When Strata-XL

fractions were not subjected to chloroform vapour, the detectable activity rose from

80 AU/ml to 160 AU/ml during the well diffusion assays. It was also observed that,

both RP-HPLC and Sep-Pak C18 purified fractions rapidly lost their activity,

especially after day two. During the polishing stage, with some possible impurities,

active fractions were obtained from HPLC fractions I3 and I4 (Figure 4.3). These

fractions were analysed using MALDI ToF spectrometry.

Figure 4.3 HPLC chromatogram of the active fractions I3 and I4 obtained during purification of peptide NI05.

Active fractions I3 and I4

Page 150: Discovery of antimicrobial peptides active against

Chapter 4

150 | P a g e

Results

4.2.4 Matrix Assisted Laser Desorption Ionisation – Time of Flight (MALDI-

ToF) mass determination

The active fractions obtained from the RP-HPLC were subjected to MALDI-

ToF analysis to determine the mass of the peptide responsible for the antimicrobial

activity. Analysis of the active fractions produced a number of peaks under 2000Da,

however all masses observed were also present in fractions lacking antimicrobial

activity (Figure 4.4). Thus, at the present concentrations we were unable to detect a

peptide peak that could consistently be associated with antimicrobial activity.

Figure 4.4 Mass spectrogram of three separate active HPLC fractions from individual purification runs. The spectrum encompassed molecules up to 10kDa in size, however the only detectable peaks were perceived below 2500Da.

I3

I3

I4

Inactive

Mass (M/Z)

Inte

nsity

(Au)

Page 151: Discovery of antimicrobial peptides active against

Chapter 4

151 | P a g e

Results

4.2.5 Genome sequence and annotation

Genome sequencing results were obtained as trimmed contiguous sequences

(contigs) from the University of Liverpool’s Centre for Genome Research. The

obtained data was assembled in house using the CLC Main Workbench software.

The draft genome sequence was assembled from a total of 3,455,707 raw sequence

reads having an average length of 142.9. Assembly resulted in 180 contigs

comprising a draft genome of 5,415,176 bp [N50 299,210 (50% of the genome was

contained in contigs of 299,210)]. From the draft data, it was estimated that average

genome coverage was 181 fold. The draft genome was designated as K. pneumoniae

A7 and annotated with the xBASE annotation service using the K. pneumoniae strain

XH209 as reference and also with PROKKA annotation software for confirmation. A

visual circular comparison of the K. pneumoniae strain A7 to strains XH209 and

RYC492 genomes was generated. This showed that compared with the RYC492 and

XH209, K. pneumoniae A7 contained a number of regions that were not as well

conserved in the genome of the reference Klebsiella strains (Figure 4.5).

Page 152: Discovery of antimicrobial peptides active against

Chapter 4

152 | P a g e

Results

Figure 4.5 Draft genome sequence of producer organism K. pneumoniae A7 compared with the reference K. pneumoniae strain XH209 and K. pneumoniae strain RYC492 that is known to express MccE492, using BRIG software. The identical regions between both genomes are represented in solid colours and differences are represented with faded colours or gaps.

4.2.6 In silico analysis of the draft genome and discovery of the MccE492 gene

cluster:

Although it was not possible to identify a specific mass through MALDI

analysis, peptide NI05 had some peculiar properties, such as low abundance/activity

in a spectrum of broths and susceptibility to chloroform. Based on these

observations, literature relating to a collection of similar antimicrobials was

examined leading to the following deductions (A) the only antimicrobial peptide

previously isolated from K. pneumoniae was MccE492 (Lorenzo, 1984) and there is

Page 153: Discovery of antimicrobial peptides active against

Chapter 4

153 | P a g e

Results

a newly discovered putative peptide MccG492 (Vassiliadis et al., 2010) from the

same cluster, though no physicochemical data are yet available on this antimicrobial,

(B) MccM, a microcin from E. coli also shows susceptibility to chloroform vapour

(Patzer et al., 2003; Vassiliadis et al., 2010) and (C) MccH45 is a microcin that

shares the same cluster as peptide MccM (Vassiliadis et al., 2010). Data relating to

these previously reported peptides were included in subsequent studies to drive an

homology based analysis.

A multiple alignment of the candidate amino acid sequences was performed

(Figure 4.6) and portions of the consensus sequence and the leader sequence of each

individual sequence were queried against the draft genome of the K. pneumoniae A7

genome. These enquiries yielded multiple matching regions and two of these regions

were identified as MccE492/MccG492 and MceS2. In addition, an MceK like region

that was also homologous to the MccM leader sequence was identified in strain A7.

Some of these peptides, such as MccE492, were not recognised by the annotation

software used (xBASE/PROKKA), thus the regions of interest were only discovered

following the above analysis.

This process revealed the presence of the complete MccE492/G492 locus

from K. pneumoniae RYC492 in the K. pneumoniae A7 genome indicating that

peptide NI05 is most probably not a novel antimicrobial, but is an additional

example of MccE492/G492. The only difference observed in strain A7 was the fact

that instead of the truncated MceK fusion protein, two ORF’s were observed in this

location; one that was homologous to McmA was named MceK2 and one that was

homologous to MccM was named MceK3, in line with the current consensus

nomenclature. MceK3 was also truncated at the same location as MceK (Figure 4.7).

The identified genetic locus is represented in Figure 4.8. It contains both putative

Page 154: Discovery of antimicrobial peptides active against

Chapter 4

154 | P a g e

Results

microcin candidates and a complete set of the genes orthologous to those that encode

MccE492, H47 and I47 transport and modification machinery in K. pneumoniae

RYC492, E. coli H47 and E. coli CA46.

Page 155: Discovery of antimicrobial peptides active against

Chapter 4

Results

155 | Pa

ge

Figure 4.6 Alignment of sequences of the structural genes encoding microcins MccM, MccH47, MccE492 and MccG492 reveals striking homology at the C-terminus. Red arrow indicates the residue (Serine 84) where the post translational moiety DHBS is attached to the MccE492 and red line indicates that the region around this location that is also conserved in other members of the group. The green line indicates the conserved GG/GA motif where leader sequences are cleaved from the active peptide.

Page 156: Discovery of antimicrobial peptides active against

Chapter 4

Results

156 | Pa

ge

Figure 4.7 MceK3 peptide sequence is homologous to the MccM N terminus prior to the stop codon TAG (Green arrow) that is also observed in the sequence derived for the MceK fusion peptide. However, the fragment following the stop codon also resembles and aligns with the serine rich C terminal of the classIIb bacteriocins including the serine 84 (Red arrow) region where the DHBS molecule is docked (Mercado et al., 2008).

Page 157: Discovery of antimicrobial peptides active against

Chapter 4

Results

157 | Pa

ge

Figure 4.8 Genome maps of the biosynthetic gene clusters involved in production, export and immunity of peptides, MccE492 (mceA), MccG492 (mceL), MccH47 (mchB), MccM (mcmA). Peptides that are of the same colour are homologues to each other [adapted from (Vassiliadis et al., 2010)].

Page 158: Discovery of antimicrobial peptides active against

Chapter 4

158 | P a g e

Discussion

4.3 DISCUSSION

Microcins of Gram-negative bacteria are far less well explored with only 19

representatives compared with 209 representatives from the Gram positive bacteria

recorded in the Bactibase database (http://bactibase.pfba-lab-tun.org/, date accessed:

09/05/2015). Most of these are derived from Escherichia species and there is only

one biochemically identified entry from K. pneumoniae. Nevertheless, their small

number contains a vast structural diversity making classification of these peptides

difficult (Duquesne et al., 2007b). Broadly speaking, bacteriocins of Gram-negative

bacteria are divided into two groups, microcins that are smaller than 10kDa and

larger colicins, which are over 10kDa in mass.

Microcins can be further sub divided into class I microcins that are 5kDa or

smaller, plasmid encoded, post translationally modified (PTM) peptides and class II

microcins that are larger than 5kDa. Class II microcins that contain PTMs are

categorised under Class IIb and those that don’t contain PTMs are collected under

Class IIa (see Section 1.6.2) (Rebuffat, 2012). During the current study, a K.

pneumoniae strain designated A7, identified from an HVS sample, presented potent

antimicrobial activity exclusively directed towards Gram-negative indicator strains,

including P. aeruginosa and Salmonella Typhimurium but was especially effective

against E. coli strains.

Although partial purification of the antimicrobial agent was successful, the

RP-HPLC fractions obtained were not sufficient to isolate a distinctive mass or for

further detailed evaluation, such as tandem mass spectrometry. Hence, an alternative

approach was chosen using the physicochemical properties of peptide NI05 to gather

similar AMPs for a comparative analysis. The amino acid sequences of these

Page 159: Discovery of antimicrobial peptides active against

Chapter 4

159 | P a g e

Discussion

antimicrobials were employed as a guide for an in silico bacteriocin search within

the K. pneumoniae A7 genome.

The only other bacteriocin previously isolated from K. pneumoniae is

MccE492, which belongs to the Class IIb microcin group. In addition, peptide NI05

was a heat stable molecule that was expressed poorly by the producer strain in

culture media and it was labile to chloroform vapour. A literature search revealed

that another microcin MccM produced by E. coli and classed in the IIb group

showed similar characteristics (Patzer et al., 2003). Chloroform susceptibility is

observed in other bacteriocins such as nisin Z and pediocin N5P from Gram positive

bacteria and is utilised as a characterisation trait (Banerjee et al., 2013; Pasteris et

al., 2014; Strasser de Saad and Manca de Nadra, 1993). Thus, the search criteria

were based around members of the Class IIb group of microcins.

Efforts culminated in the discovery of the MccE492/G492 gene cluster in

strain A7. Both microcins encoded in this cluster belong to the class IIb microcin

family and share a conserved serine rich C terminus, where a siderophore moiety is

attached that binds to a single iron molecule to facilitate their Trojan horse like

activity (Thomas et al., 2004). Hence, members of this group are also defined as

“siderophore microcins” (Rebuffat, 2012). Siderophores are chelating agents

secreted by the bacterial cell to bind and facilitate the transport of highly insoluble

iron (Fe+3) molecules into the bacterial cell (Saha et al., 2015). These complexes are

recognised by cell surface receptors and translocated into the cell. Among these

receptors are Cir, FepA and in particular Fiu that recognise the salmochelin

siderophore resultant from enterobactin (Arnison et al., 2013; Fischbach et al., 2006;

Strahsburger et al., 2005; Vassiliadis et al., 2007). Enterobactin itself is formed by

Page 160: Discovery of antimicrobial peptides active against

Chapter 4

160 | P a g e

Discussion

three N-(2,3-dihydroxybenzoyl)-L-serine (DHBS) molecules that are bound together

to form a cyclic trimer by NRP machinery making salmochelin bacteriocin hybrid

NRP/RiPP peptides (Fischbach et al., 2006; McIntosh et al., 2009). This machinery

is also encoded in the K. pneumoniae A7 genome. In addition to MccE492, MccM

and MccH47 (Vassiliadis et al., 2010) were also shown to exploit this pathway by

carrying a postranslationally attached DHBS molecule on their C terminus that is

recognised by the salmochelin receptors. The resulting translocation of the microcin

into the cell is driven by a TonB and mannose permease dependent pathway (Bieler

et al., 2005; Poey et al., 2006). Mechanisms of this modification are studied most

extensively in MccE492. It was observed to greatly enhance the antimicrobial

activity of unmodified MccE492 (Destoumieux-Garzón et al., 2006) and although a

modified version was initially referred to as MccE492m, it is now considered to be

the mature peptide (Duquesne et al., 2007a).

The modification pathway is most extensively described by Vassiliadis and

colleagues (2007) and involves the MceACDIJ gene cluster (Figure 4.9); the putative

glucosyl transferase MceC is involved in the glycosylation of the enterobactin (step 1

in Figure 4.9). Once enterobactin is glycosylated, it is cleaved by an enterobactin

esterase (MceD) to achieve linearized glc-DHBS3. If the glc-DHBS3 is not utilised

this cleavage process was observed to be continuous until a single glc-DHBS

molecule is achieved and thus gives rise to intermediate MccE492 products

MccE492-glc-DHBS2 and MccE492-glc-DHBS (Step 2). The linearized glc-DHBS3

is then attached to the C-terminal serine-84 of EccE492, through covalent bonding

aided by the MceI peptide homologous to the RTX toxin acyltransferase family

(Mercado et al., 2008), which are involved in fatty acid acylation and activation of

Page 161: Discovery of antimicrobial peptides active against

Chapter 4

161 | P a g e

Discussion

protoxins such as E. coli haemolysin (Destoumieux-Garzón et al., 2006; Duquesne et

al., 2007a; Trent et al., 1999) (Step 3).

Although its contribution to this mechanism is still unknown MceJ,

homologous to the MchC from the MccH47 gene cluster, is conserved in

biosynthetic gene clusters of E. coli CA46 and H47 strains and also conserved in the

K. pneumoniae A7 cluster (Duquesne et al., 2007a; Vassiliadis et al., 2010).

However, it is observed to be necessary for the production of a functional MccE492

peptide and is believed to have a role in the process where glc-DHBS3 is fused to the

precursor peptide using serine-84 for docking (Mercado et al., 2008)(Lagos et al.,

2001; Thomas et al., 2004; Vassiliadis et al., 2007). The genes involved in

biosynthesis of MccE492 were shown to be exchangeable with their counterparts

involved in MccH47 and MccM biosynthesis (Vassiliadis et al., 2010).

Figure 4.9 Post translational modification pathway of MccE492 with the DHBS siderophore moiety. 1) gycolysation of enterobactin by MceC. 2) cleavage of enterobactin to achieve linear Glc-DHBS3, 3) attachment of the linear Glc-DHBS3 and derivatives to the MccE492 precursor peptide (MceA) [from (Vassiliadis et al., 2007)].

Page 162: Discovery of antimicrobial peptides active against

Chapter 4

162 | P a g e

Discussion

Since the description of this mechanism, there have not been significant

updates, except the discovery that the adenylation domain of EntF, a protein

involved in enterobactin production, was also essential for Mcc492 maturation in a

manner unrelated to enterobactin production (Mercado et al., 2008). Also, since the

discovery of MccH47 and MccM in 2010 no new members of this group have been

identified. However, research into microcins that utilise similar Trojan horse like

activity such as MccJ25 that carries a lasso fold conferring additional stability to the

microcin (Arnison et al., 2013; Hammami et al., 2015; Mathavan et al., 2014),

MccB17 (Mathavan and Beis, 2012; Thompson et al., 2014) and nucleic acid

carrying MccC (Vondenhoff and Van Aerschot, 2011; Zukher et al., 2014) has

generated some more interest in this group of AMPs.

K. pneumoniae A7 also carried the MceG and MceH genes that together

encode a functional ATP-binding cassette (ABC) transporter unit that provides

energy for transport and release of the MccE492. These genes are commonly found

in bicistronic conformation, as in the K. pneumoniae A7 cluster (Vassiliadis et al.,

2010). The MceF protein is also thought to play a role in transport however its

function is still not clear (Duquesne et al., 2007a; Lagos et al., 2009, 2001;

Vassiliadis et al., 2010). These regions are also related to mchF and mchE genes that

encode an ABC transporter in E. coli CA46 and H47 (Azpiroz and Lavina, 2004;

Azpiroz et al., 2001). These observations demonstrate that the K. pneumoniae A7

operon encodes a complete set of putative genes required for export and modification

of siderophore microcins.

Page 163: Discovery of antimicrobial peptides active against

Chapter 4

163 | P a g e

Discussion

It was interesting to see that the MccE492/G492 cluster is conserved wholly

between the two Klebsiella species including the truncated MceK like peptides,

MceK2 that shares similarities with the mcmI gene product, which confers self-

immunity to MccM by encoding the immunity protein McmA (Braun et al., 2002;

Duquesne et al., 2007a; Vassiliadis et al., 2010). Also present is the truncated

MceK3 peptide that shares a high degree of homology with the MccM leader

sequence (Vassiliadis et al., 2010). MceK3 was highlighted as a putative bacteriocin

candidate by the BAGEL analysis of the K. pneumoniae A7 genome, which

identified the region as including a predicted class II bacteriocin gene (see, Chapter

5)

The gene encoding MceK3 was also truncated at the same location as MceK

by a premature TAG stop codon that occurs after the GG motif, where the leader

cleavage of the microcin prepetide is expected to occur before the serine rich C-

terminal (Figure 4.7). Intriguingly, a similarly truncated MccM sequence is also

observed in E. coli strain H47 (Vassiliadis et al., 2010). In addition to MceK, the

MhcS2 coding region also shares N terminal homologies with MccI47 but lacks the

serine rich C-terminal region (Vassiliadis et al., 2007).

It is known that MccE492 is still active without its posttranslational

modification, although less so (Duquesne et al., 2007b; Thomas et al., 2004), which

may explain the conservation of this region. However, an interesting attribute of

MccE492 is its ability to form amyloids (Arranz et al., 2012; Bieler et al., 2005). In

the case of MccE492 it is suggested that these amyloids may be involved in

bacteriocin regulation and act as microcin reservoirs or protect the producer (Arranz

et al., 2012) and it was also observed that successful post translational modification

Page 164: Discovery of antimicrobial peptides active against

Chapter 4

164 | P a g e

Discussion

of the bacteriocin impedes the amyloid formation (Marcoleta et al., 2013b). Thus, it

will be interesting, considering the close relationship of this family (Vassiliadis et

al., 2010), to investigate if the truncated peptide is capable of exerting bactericidal

activity and/or forming amyloids that may be involved directly or indirectly in

another desirable process and thus are conserved.

Amyloid formation is mainly associated with neurodegenerative conditions

such as Alzheimer’s disease in humans due to defective peptides such as N-

terminally truncated Aβ peptides (Dobson, 1999; Lashuel et al., 2002; Oberstein et

al., 2015). However, in bacteria amyloids were also described in constructive

processes such as hyphae formation in Streptomyces coelicolor (Claessen et al.,

2003) and the curli and Fap amyloids that are involved in biofilm formation in

Salmonella enteritidis and P. aeruginosa (Austin et al., 1998; Chapman et al., 2002;

Dueholm et al., 2013), respectively, which are only some examples.

In conclusion, during this study it has been demonstrated that the draft

genome of K. pneumoniae strain A7, which was isolated from an HVS sample,

encodes the MccE492/G492 gene cluster in its entirety, suggesting that this a well

conserved region and is possibly important in the ecology of K. pneumoniae strains.

Thus, closer inspection of the uncharacterised genes in this cluster will be important.

It was disappointing to see these well-studied regions were not annotated by the

software used and had to be manually uncovered. While isolation of peptide NI05

was not readily possible, in light of the current findings, it is most likely that the

antimicrobial activity observed from strain A7 was, at least in part if not wholly,

conferred by MccE492. It may be concluded that the peak representing the peptide

was masked during MALDI-ToF as result of its low abundance in the purified

Page 165: Discovery of antimicrobial peptides active against

Chapter 4

165 | P a g e

Discussion

fractions, due to presence of high intensity contaminants such as those at 1795Da

and 926Da, as all the necessary components for its production and secretion are

present and the physicochemical properties of NI05 and MccE492 are so similar.

Page 166: Discovery of antimicrobial peptides active against

Chapter 5

166 | P a g e

Introduction

5 CHAPTER 5

GENOME MINING OF ANAEROBIC BACTERIA, KNOWN PRODUCER

BACTERIA AND CELL FREE EXPRESSION OF PUMICIN NI04

Page 167: Discovery of antimicrobial peptides active against

Chapter 5

167 | P a g e

Introduction

5.1 INTRODUCTION

Ribosomal peptides such as bacteriocins that are the focus of this study are

hard to locate in genome sequence data as they can be encoded by poorly conserved

and small open reading frames (ORFs). These can be overlooked by programs such

as BLAST that relies on sequence to sequence homology (Lee et al., 2008; Schmidt

et al., 2005). There are some powerful search tools that can identify more distant

relationships, such as PSI BLAST, which profiles homology and HHpred that

generates Hidden Markov Model (HMM) to HMM profile comparisons (Altschul et

al., 1997; Söding et al., 2005). However, novel antimicrobial discovery using these

tools is labour intensive. A LanT peptide driven study performed by Singh and

Sareen (2014) is one such study.

Comparatively, BAGEL and antiSMASH are convenient, free to use web

based tools (Blin et al., 2013; de Jong et al., 2010, 2006). They combine homology

searches with contextual analysis to improve antimicrobial peptide discovery. Thus,

in addition to screening for homology between known bacteriocin ORFs and the

query sequence, they also scan for additional ORFs that are located in the immediate

surroundings of known bacteriocins. These genes mostly code for the proteins that

are required for processing, transport, modification, regulation and producer self-

immunity to the bacteriocin (de Jong et al., 2010, 2006) and include LanR, LanK,

LanM and ABC transporters (see Section 1.5.1) (Arnison et al., 2013; Dirix et al.,

2004; Dischinger et al., 2009; Herzner et al., 2011; Siezen et al., 1996; Singh and

Sareen, 2014). An example result output can be seen in Figure 5.1.

Page 168: Discovery of antimicrobial peptides active against

Chapter 5

168 | P a g e

Introduction

In silico analysis offers a unique angle to antimicrobial peptide research as it

makes it easier to screen hard to grow bacteria, including some toxigenic species

(Eg: Clostridium botulinum), and allows analysis of metagenomic data (Kang and

Brady, 2013).

Anaerobic bacteria are an emerging source for discovery of antimicrobial

peptides. Many studies have proven their capability to produce a plethora of

secondary metabolites and antibiotics; perfrin, clostrubin and ruminococcin C are

only some examples (Crost et al., 2011; Letzel et al., 2014; Pidot et al., 2014;

Timbermont et al., 2014).

In addition, although anaerobes are proving a rich source to mine in silico,

heterologous anaerobic expression of these metabolites will still require extensive

exploration and optimisation of growth conditions. This could be circumvented

through molecular cloning, recombinant expression of the candidate peptide in

aerobic hosts or cell free expressions, which are viable options. In this study, we

Figure 5.1 Graphical result representation of BAGEL3 software analysis. The gene cluster belongs to lead 4 (see below), a putative class II lantibiotic identified during this study and is surrounded by important motifs. LanK and LanR are known to be involved regulation of class II bacteriocins, the ABC transporter is responsible for secretion) and SacCD encodes the SAM enzyme that introduces sulphur bridges.

Page 169: Discovery of antimicrobial peptides active against

Chapter 5

169 | P a g e

Introduction

sought to investigate the ability of BAGEL 3 software to generate reliable leads that

were likely produce an active product following cloning and expression.

Bacteriocins were identified from anaerobic bacteria, mainly focusing on

Clostridium species that are known for their toxigenic nature (Hatheway, 1990) and

Propionibacterium species that have probiotic applications (Darilmaz and Beyatli,

2012; Ekinci and Gurel, 2008; Filya et al., 2004). Having generated genome

sequence data for the two producers isolated during this study (see Chapters 3 and

4), it was decided that these data should also be included. As a proof of concept and

to assess the feasibility and convenience of the direct cloning approach, cloning and

cell free expression of the pumicin NI04 structural gene was attempted, given that

this had been characterised as encoding an active peptide during the current study.

Page 170: Discovery of antimicrobial peptides active against

Chapter 5

170 | P a g e

Results

5.2 RESULTS

5.3 GENOME MINING RESULTS

Mining the genomes of 35 anaerobic bacteria for putative antimicrobial

peptides using BAGEL 3 software yielded 33 possible producer bacteria and 66

predicted putative bacteriocins (Table 5.1). Most of the identified products belonged

to two major groups, class II bacteriocins and sactipeptides. Other identified peptides

included those from class III, class IV, LAPs, bottromycins, lasso peptides and the

lanthipeptides. In addition, analysis of the producer bacteria B. pumilus J1 and K.

pneumoniae A7, identified during this project, yielded seven additional putative

bacteriocin candidates (one has been described and discussed above, in Chapter 4).

Pumicin NI04, described in Chapter 3, was not identified by BAGEL, which is not

surprising given its lack of homology to previously identified bacteriocins.

Closer inspection of the data obtained revealed that 19 out of the 73 putative

bacteriocins were promising leads, as they were either homologs to known

bacteriocins or accompanied by multiple genes that are relevant to bacteriocin

production, including modification, immunity and transport proteins. Five of these

leads only shared significant identities (E value < 0.001) with previously defined

bacteriocins and weren’t located with any other bacteriocin related accessory genes,

except ABC transporters; lead 11 with lactococcin 972 (38%), lead 10 with linocin-

M18 (47.9%), lead 13 and 15 with garvicin ML (both 20%) and leads 16 and 17 with

UviB (25%) (Table 5.1).

Page 171: Discovery of antimicrobial peptides active against

Chapter 5

171 | P a g e

Results

The leads that displayed both homology to known bacteriocins and were also

co-located with bacteriocin related biosynthetic clusters are described in more detail

below. The only exception is lead 18 that was homologous to MccM (71.2%); this

peptide was discovered during the work described in Chapter 4 and is in fact a

truncated MccM like peptide.

Page 172: Discovery of antimicrobial peptides active against

Chapter 5

Results

172 | Pa

ge

Table 5.1 The short list of promising bacteriocin leads identified in the genomes of 35 species of anaerobic bacteria of the genera Clostridium and Propionibacterium and the producer bacteria K. pneumoniae A7 and B. pumilus J1. Unique leads are novel leads that do not share homology with known bacteriocins but are located with bacteriocin related genes.

Lead No Bacteria Location Type Amino acid sequence Significant

homologies Anaerobes

1 Clostridium

acetobutylicum 824 (NC003030)

Chromosome Sactipeptides MNFLKNFYSCVLHNSVSNILCILLYFCLLGNNTIIYYFCKIGKIKKVA Unique

2 Clostridium beijerinckii

NCIMB8052 (NC0090617)

Chromosome ClassIII

MKKQLRKVMILAVSLLMLLGQSPNIVFAATSPNISSSAYNSENIFTQSGYKGQCTWFVWGRAYEKLGIKLNSQFYGNAKQWWNETTYPKGQTPAANSIAVFGNGSAGHVVFIESVSGDTVYFNEANYHVSKAYDGAEENQTVSAFKSRSANFLGFIYLQGNPSEPSDPTQSGFTYPNNAQVSGDFLYVRDSSGNIISGRRVDDGDKITVLDVSYSTQLALIEYPTPNGVRTGYVTNATNIIKYFNDGAWKNGSTSENVYDSNGNVIGSLSPGETATPLYRKNGKLSVVYNTSKGANTKSGFVVYNGGFNKY

Unique

3 Clostridium cellulovorans

743B (NC014393) Chromosome Lant. class_II

MQNYESKAGFISEMELDELVSNKTVGGATTVPCAIAIIGITLSAGICPTSACSKDCPWNN

lichenicidin A2

(35%)

4 Clostridium lentocellum DSM 5427 (NC015275)

Chromosome Sactipeptides MELCTIDGLKLDKFEDLKSTELEEVNGGGAGAVIAVVVVICVIAFAVGVYNGYQDNKK

Unique

5

Clostridium saccharolyticum WM1

(NC014376) Chromosome Sactipeptides MKENLQMHLILPLLIRCGKMNPKEDTAAGQENK

Unique

Page 173: Discovery of antimicrobial peptides active against

Chapter 5

Results

173 | Pa

ge

Lead No Bacteria Location Type Amino acid sequence Significant

homologies

6 Clostridium stercorarium DSM 8532 (NC020887)

Complete Genome

Sactipeptides MKQVQPQGFLNCVLICVGGCLTCISDGPVIIVDFVSGGTAASTGSSA

Unique

7

Clostridium botulinum A strain ATCC 19397

(NC009697) Chromosome LAPs VAPGSCCCCSCCCCVSVSVGGGSASTGGGAAAGQGGN Unique

8 Clostridium perfringens strain 13 (NC003366)

Chromosome Lasso peptide MVSKLISLDSIVSQKEGIDVTELNGEKVMMDLDKGKYFMLNETGSTIWDAINEPKSVSEIIEATIKEYDIDRETCESKVLEYLEKLRHEEIVFIN

Unique

9 Clostridium species

BNL1100 (NC016791) Chromosome Lant. class_II MSHSTIGFAFEDLGEKEMAASQNASSSANSIVTIYSLTGT

CLPSITVTICQVTVPVTITVTAAN Unique

10

Propionibacterium acidipropionici ATCC

4875 (NC019395)

Complete Genome

ClassIII

MNNLHRELAPISEAAWKQIDDEARDTFSLRAAGRRVVDVPEPAGPTLGSVSLGHLETGSQTDGVQTSVYRVQPLVQVRVPFTVSRADIDDVERGAVDLTWDPVDDAVAKLVDTEDTAILHGWEEAGITGLSEASVHQPVQMPAELEQIDDAVSGACNVLRLADVEGPYDLVLPQQLYTQVSETTDHGVPVVDHLTQLLSGGEVLWAPAARCALVVSRRGGDSCLFLGRDVSIGYLSHDAQTVTLYLEESFTFRVHQPDAAVALV

Linocin-M18

(47.9%)

11

Propionibacterium acnes TypeIA2 Pacn17

(NC016512)

Complete Genome

ClassII (unmodified)

MTLSLLLAMKRFSRRSYKEEIVSTVWRHFAAVPITGLLLASTATIAMADTVNPPEGGVWRYGRGYSDYFHGSKKHGSSVQGVNFVRSPCVKAGSWSYARTDAARFGNKAWYRTC

Lactococcin 972 (38%)

Page 174: Discovery of antimicrobial peptides active against

Chapter 5

Results

174 | Pa

ge

Lead No Bacteria Location Type Amino acid sequence Significant

homologies

12

Propionibacterium propionicum F0230a

(NC018142)

Complete Genome Lant. class_II MQSNYDDQILDNLAELSDTEIEELLGAGWGSIFSFTHECN

TNTMTQLFTCCF Ruminococcicin A (38.30%)

Producers

13 B. pumilus J1 Draft Genome Head to tail

cyclised peptide

MTKATDSKFYALLSLSLLAVTLIALVIGNGSLIAANLGVSTGTAFTIVNFLDAWSSVATVITIVGMFTGVGTISAGVAASILAIIKKKEKSKAAAF

Garvicin ML

(20%)

14 B. pumilus J1 Draft Genome Head to tail

cyclized peptide

MTETRNEIKLHVLFGALAVGFLMLALFSFSLQVLPVADLAKEFGIPGSVAAVVLNVVEAGGAVTTIVSILTAVGSGGLSLIAAAGKETIRQYLKNEIKKKGRKAVIAW

enterocin AS-48

(80%)

15 B. pumilus J1 Draft Genome Head to tail

cyclized peptide

MGSSEGIASLASIDFNSGNQFALDLATNLGISRKTAYAAIGVIMTTGDVLTILSLLAVVLGGTGLVTAAMVATAKKLATKHGKKYAAEW

Garvicin ML

(20%)

16 B. pumilus J1 Draft Genome Class 2 VEMDLTQYLMTQGPFAVLFCWVLFYVLNTTKERENKLNEQIEAQNDVLAKFSEKYDVVIDKLDKIERNLK

UviB

(25%)

Page 175: Discovery of antimicrobial peptides active against

Chapter 5

Results

175 | Pa

ge

Lead No Bacteria Location Type Amino acid sequence Significant

homologies

17 B. pumilus J1 Draft Genome Class 2 MVEMDLAQYLMTQGPFAVLFCWVLLYVLNTTKERESKLNEQIEAQNEVLAKFSEKYDVVIDKLDKIERNLK

UviB

(25%)

18 K. pneumoniae A7 Draft Genome Microcin MRKLSENEIKQVSGGDGNDGQAELIAIGAIAGTFLSPGIGSIAGAYVQWSLALEWRSAL

Microcin M

(71.20%)

19 K. pneumoniae A7 Draft Genome Bottromycin MPVPNHRHDENVLSIDEIFQLKLTSAFINKPAKTIAYYLLSHLFSSSG

Unique

Page 176: Discovery of antimicrobial peptides active against

Chapter 5

176 | P a g e

Results

5.3.1 Lead 3

Lead 3 had significant homology with the A2 component of lichenicidin

(35%) (Begley et al., 2009; Letzel et al., 2014), a two peptide class II lanthipeptide

bacteriocin. It also aligned with beta components of two other peptide lantibiotics,

including entrocin W beta (29%) (Sawa et al., 2012) and plantaricin W (26.7%)

(Holo et al., 2001), especially in the C terminal portions (Figure 5.2). While the

cluster contains two sets of LanM class II lanthipeptide modification proteins and

two sets of LanT transporter proteins, BAGEL 3 did not pick up a partner peptide to

lead 3 (Figure 5.2). Although a visual inspection of the intergenic region was not

carried out, Letzel et al. (2014) have identified a companion peptide.

Figure 5.2 A) Amino acid sequence alignments of lead 3 with two component members of the class II lanthipeptide family, including lichenicidin A2 (orf023). B) The biosynthetic cluster of lead 3 as predicted by BAGEL 3 software.

A

B

Page 177: Discovery of antimicrobial peptides active against

Chapter 5

177 | P a g e

Results

5.3.2 Lead 12

Lead 12 is another promising bacteriocin candidate that shares homology

with class II lanthipeptides and strong sequence alignment with ruminococcicin A

(38.3%) (Dabard et al., 2001), lacticin 481 (36%) (Piard et al., 1992) and salivaricin

9 (26%) (Wescombe et al., 2011) (Figure 5.3). The genome cluster also contained an

ABC transporter gene involved bacteriocin secretion, in addition to class II

lanthipeptide modification genes LanC and LanM.

Figure 5.3 A) Lead 12 sequence alignment shows conserved regions with ruminococcicin, lacticin 481 and salivaricin 9. B) The putative bacteriocin is accompanied by LanM and LanC modification components, multiple ABC transporters and two bacteriocin related genes GerE and HisKA 3 with unknown functions, as predicated by BAGEL 3.

A

B

Page 178: Discovery of antimicrobial peptides active against

Chapter 5

178 | P a g e

Results

5.3.3 Lead 14

Lead 14, predicted from the producer B. pumilus J1, is intriguing as it is 80%

identical and has strong C terminal alignments to enterocin AS-48, a circular class

IV inhibitor of Gram positive and negative bacteria (José et al., 2014), as well as

Circularin A (28.95%) (Kemperman et al., 2003b) and Leucocyclicin Q (29.7%)

(Masuda et al., 2011). However, lead 14 has a significantly longer N-terminal

sequence and a higher predicted mass of 9684.398Da compared with 7186.77Da for

enterocin AS-48 (Figure 5.4). The cluster contains two putative modification

proteins and an ABC transporter (Figure 5.4).

5.3.4 Unique putative bacteriocin candidates

The mining efforts also yielded some genes encoding unique putative

bacteriocins that were flanked by multiple bacteriocin related biosynthesis genes and

did not exhibit significant homology with known bacteriocins (E value>0.001); these

were leads 1, 2, 4, 5, 6, 7, 8, 9 from anaerobic bacteria and lead 19 from K.

pneumoniae A7. A selection of genome maps from these predicted bacteriocins are

displayed below (Figure 5.5).

Page 179: Discovery of antimicrobial peptides active against

Chapter 5

179 | P a g e

Results

Figure 5.4 A) the amino acid sequence alignment of lead 14 with circular bacteriocins Enterocin AS-48, Leucocyclicin Q and Circularin A. B) The anticipated biosynthetic gene cluster consists of two modification genes, HttB and HttC and an ABC transporter.

B

A

Page 180: Discovery of antimicrobial peptides active against

Chapter 5

180 | P a g e

Results

Figure 5.5 Unique bacteriocin sequences predicted by BAGEL 3 software. A) Lead 4 is a predicted sactipeptide, B) lead 7 is a predicted LAPs peptide, C) lead 19 is a putative bottromycin and D) lead 9 is a likely class II lanthipeptide family bacteriocin.

A

B

C

D

Page 181: Discovery of antimicrobial peptides active against

Chapter 5

181 | P a g e

Results

5.4 CLONING RESULTS

The outcome of in silico bacteriocin mining work was analysed and a short

list of likely putative bacteriocins was created. However, the only way to confirm the

antimicrobial potential of the identified leads is through direct observation of

bacterial inhibition in response to presence of the suspected bacteriocin lead. This

also requires an understanding of the potential target bacteria, which would be used

in screening assays.

With a robust recombinant or cell free expression system in place that can

generate sufficient quantities of the peptide of interest, identified leads can be

assessed in a cost and time efficient manner. The proposed system in this project was

based around the PURExpress cell free expression system, where expression is

controlled and facilitated by the pET-28a plasmid vector. As a proof of concept, the

efficiency of this system was assessed by attempting the expression of pumicin

NI04.

To initiate the process, pET-28a plasmids were extracted from pre

transformed E. coli XL1 blue cells using the Qiagen miniprep kit. The insert was

amplified from the B. pumilus J1 genome using PCR. Both the plasmid and the insert

were digested using the appropriate restriction enzymes (see Section 2.5). Digested

products were gel purified (Figure 5.6) and ligated. Transformation reactions were

performed using the T4 DNA ligase and plated out on LB/Kan selection plates.

Page 182: Discovery of antimicrobial peptides active against

Chapter 5

182 | P a g e

Results

5.4.1 Insert confirmation and validation following transformation

Following the transformation reactions, 26 colonies were observed for the

pmnA construct on the LB/Kan selection plates. These colonies were checked for the

presence of an insert using the colony PCR method. Results indicated that 22 out of

26 colonies contained an amplifiable product of the pmnA insert (Figure 5.7). The

correct product length was confirmed using the PCR amplified insert DNA from the

producer organism B. pumilus.

Figure 5.6 Digested NI04 structural gene insert (left) and pET28a vector DNA products (right) prior to gel purification.

Page 183: Discovery of antimicrobial peptides active against

Chapter 5

183 | P a g e

Results

Figure 5.7 Gel image displaying the comparison of colony PCR amplified pmnA bands from the clones (CL) to those amplified from the producer. Positive control (+) was insert DNA amplified from producer B. pumilus J1 DNA and the negative control (-) was amplification from pET28a plasmid DNA with no insert (the template vector is visible in the negative control lane). Expected product length was 319bp.

After the confirmation of the clones (CL) carrying the insert from the

selection plate, CL1 and CL2 were picked at random to carry out expression studies.

However, before committing to expression studies, the orientation of the inserts was

validated by substituting the reverse primer with the T7 downstream primer and

performing PCR amplification. The amplicons generated were at the expected ladder

interval, located between 400-500bp markers (Figure 5.8). Once correct orientation

was confirmed, the nucleotide sequence for both clones was validated by sequencing

the insert region within the plasmid.

Page 184: Discovery of antimicrobial peptides active against

Chapter 5

184 | P a g e

Results

Consensus nucleotide sequences generated for each clone were then aligned

to the expected nucleotide sequence. The results indicated that, while the insert

carried by CL2 encoded the correct peptide, the 3’ end of the insert carried by CL1

had errors and encoded a faulty protein (Figure 5.9). As a result, expression studies

were conducted using plasmids generated from the CL2 stocks.

Figure 5.8 Insert orientation was confirmed using pmnA gene forward primers and the T7 promoter region reverse primers. The expected product length was 435bp.

Page 185: Discovery of antimicrobial peptides active against

Chapter 5

Results

185 | Pa

ge

Figure 5.9 CL1 and CL2 insert region sequence alignments with the expected pET28a plasmid T7 region containing the pumicin NI04 gene pmnA. The alignments showed that while CL2 encoded the correct sequence, the CL1 insert had a mutation in the 3’ end and possibly encodes a non-functional peptide

.

Page 186: Discovery of antimicrobial peptides active against

Chapter 5

186 | P a g e

Results

186 | Pa

ge

5.4.2 In vitro expression of pumicin NI04

Following the in vitro expression reaction, the resulting peptide mixtures for

the DHFR control and the experimental sample were filtered using a 100kDa cut off

filter and the filtrate was reverse purified using the Ni-NTA His pure beads. A

sample from the whole mixture, the 100kDa flow through and the final purification

product of each reaction was tested for antimicrobial activity using the agar well

diffusion assay. The bioassay of the final purification products using the well

diffusion assay failed to produce inhibition zones around the wells that contained

both the purified experimental peptide and the control peptide (DHFR). However

when the whole reaction mixture and the 100kDa cut-off filtrate were tested there

were inhibition zones surrounding all test samples including the negative control that

contained only the reaction mixture without the added pET28a-pmnA or the DHFR

plasmid (data not shown). This indicates that the reaction buffer contained an

inhibitor that needed to be excluded prior to antimicrobial bioassays in order to

reveal any activity specific to the recombinant pumicin NI04.

Subsequent SDS- PAGE gel analysis of the samples exposed two bands that

aligned with the expected protein weight marker intervals for both the experimental

and control mixtures (Figure 5.10). However, the expression of the experimental

peptide was observed to produce a considerably lower yield than that of the DHFR

peptide expressed by the control plasmid. Although it is possible that the peptide was

diluted past its MIC during the purification process, it was not possible to test

antimicrobial activity without purification of the product due to the antimicrobial

properties of the reaction components mentioned above.

Page 187: Discovery of antimicrobial peptides active against

Chapter 5

187 | P a g e

Results

187 | Pa

ge

1 2 3 4 5 6Ladder

3

6

14

17

28

38

49

~MW (kDa)

Figure 5.10 Reverse purification of the in vitro expressed pumicin NI04 indicated with red arrows and positive control DHFR indicated with black arrows, (1) Ni-NTA purified DHFR, (2) Ni NTA purified pumicin NI04, (3) DHFR 100kDa ultrafiltration flow through, (4) pumicin NI04 100kDa ultrafiltration flow through, (5) DHFR whole reaction protein, (6) pumicin NI04 whole reaction protein.

Page 188: Discovery of antimicrobial peptides active against

Chapter 5

188 | P a g e

Discussion

5.5 DISCUSSION

Mining of publicly available genomic data for putative bacteriocin candidates

produced by anaerobic bacteria, B. pumilus J1 and K. pneumoniae A7 with BAGEL

3 software has yielded 73 different predicted bacteriocin candidates that are referred

to as the “leads”. Many of the peptides identified belonged to the class II unmodified

and miscellaneous sactipepetide families. However, LAPs, large bacteriocins, lasso

and lanthionine containing peptides were also predicted, demonstrating that a large

spectrum of bacteriocin classes may be accessible from anaerobic bacteria. From the

bacteriocin library created, nineteen candidates were chosen as promising leads and

short listed for possible cloning in future projects. In order to be considered

“promising”, leads had to either be clustered with multiple bacteriocin accessory

genes and/or exhibit homologies with known bacteriocins. The putative bacteriocins

that contained both desirable properties are discussed below.

The first promising lead that stood out was the putative class II lanthipeptide

lead 3. This peptide formed alignments with two-peptide bacteriocins of this group

and specifically with the lichenicidin A2 peptide, instead of the A1 moiety (Begley

et al., 2009; Dischinger et al., 2009; Shenkarev et al., 2010), and with enterocin W

and plantaricin W beta peptides, but not the alpha components (Holo et al., 2001;

Sawa et al., 2012). These results, and the presence of two LanM proteins, indicated

that there should be a second AMP structural gene, but BAGEL 3 had only returned

a single predicted bacteriocin. However, other mining studies have revealed a second

putative bacteriocin, ClocelDRAFT_0418 (equivalent to an anticipated lead 3 A1),

which was identified from the same cluster that precedes the lanM gene and is

located upstream of lead 3 A2 (de Jong et al., 2010; Letzel et al., 2014). The putative

Page 189: Discovery of antimicrobial peptides active against

Chapter 5

189 | P a g e

Discussion

lead 3 A1 (Letzel et al., 2014) is 35% identical to lichenicidin A1 and also forms

alignments with Enterocin W and plantaricin W alpha peptides, especially following

the conserved GG/GA motif where the precursor peptide leader sequence is likely to

be cleaved (Caetano et al., 2014). The putative biosynthetic machinery identified is

also similar to that of lichenicidin, which contains two sets of LanM and one LanT

peptide (Begley et al., 2009; Dischinger et al., 2009; Shenkarev et al., 2010). Thus, a

complete set of prepeptides, modification and transport components are present in

the identified gene cluster for production of an active two-component class II

lanthipeptide. It is quite interesting that BAGEL 3 software did not pick up the

putative A1 component of the lead 3 cluster, especially as a paper published by de

Jong et al. using BAGEL 2, has actually reported the presence of this putative A1

component, but not the A2 homolog (de Jong et al., 2010). The BAGEL3 software

does not discard the previously described bacteriocins.

Lead 12 on the other hand was predicted to be another significant class II

lanthionine containing peptide candidate that was homologous to ruminococcicin A,

an inhibitor of Clostridium perfringens, which is produced by another anaerobic

bacterium Ruminococcus gnavus (Dabard et al., 2001). Ruminococcicin A is an

intriguing peptide that is produced in the presence of trypsin (Gomez et al., 2002;

Marcille et al., 2002). In addition to the presence of a LanM protein for modification

and an ABC transporter for secretion, it is possible that the predicted “unknown

related” peptides GerE and histidine kinase HiskA3, which play roles in signal

transduction and regulation can be involved in regulation of expression of this

putative bacteriocin (Attwood et al., 2007; Ducros et al., 2001; Wolanin et al.,

2002). Salivaricin 9 (Barbour et al., 2013; Wescombe et al., 2011) and lacticin

Page 190: Discovery of antimicrobial peptides active against

Chapter 5

190 | P a g e

Discussion

481(Piard et al., 1992), two potent Gram positive active agents were also related to

lead 12.

Lead 11 on the other hand is a putative class II bacteriocin similar to

lactococcin 972 encoded by P. acnes TypeIA2 Pacn17 that exerts its activity by

binding to lipid II cell wall precursor (Martínez et al., 2008). It was observed in a

previous study that most propionibacteria possessed a lactococcin 972 like peptide

(Letzel et al., 2014), however, BAGEL3 analysis of propionibacteria sequences in

the current study only yielded one such peptide. The reason we have selected this

peptide for the short list is the ease of successful functional cloning of class II

bacteriocins as they don’t require complex modification machineries (Olejnik-

Schmidt et al., 2014; Sandiford and Upton, 2012).

Amongst two putative large class III peptides chosen for the “short list”, the

one identified from P. acidipropionici was the most interesting. Large bacteriocins

have been previously identified from anaerobic bacteria, propionicin SM1 being one

example identified from P. jensenii DF1 (Miescher et al., 2000). However, what

makes P. acidipropionici an intriguing candidate is the fact that the peptide it

encodes is homologous to linocin M18, a large bacteriocin produced by the Gram

positive bacterium Brevibacterium linen, which was isolated from a cheese product

(Valdés-Stauber and Scherer, 1994). Propionibacterium species, including P.

acidipropionici, are commonly found in dairy products, such as yoghurt and cheese

and are also used as additives in animal feed and observed to improve the fungistatic

and antimicrobial properties of these products in multiple studies (Darilmaz and

Beyatli, 2012; Ekinci and Gurel, 2008; Filya et al., 2004). The significance of this

result is that it highlights another application of in silico discovery methods, namely

Page 191: Discovery of antimicrobial peptides active against

Chapter 5

191 | P a g e

Discussion

producer identification. If functional production of putative linocin M18 like

bacteriocin from P. acidipropionici can be confirmed this may aid in its

incorporation in new probiotic products for human or animal use.

The remaining large class III bacteriocin had significant homology to

bacteriocin BCN5 according to the BAGEL 3 software. Production of this

bacteriocin was shown to be activated following DNA damage due to UV exposure

and it will thus be interesting to attempt its cloning to see if it would be possible to

bypass this activation using an inducible promoter (Dupuy et al., 2005).

In addition to those peptides that shared homology with known bacteriocins,

there were also eight putative unique bacteriocin clusters, which were observed to

contain a selection of bacteriocin related genes and some are discussed here. Lead 1

is a good example for sactipeptides; it carries a predicted SacCD region, a radical S-

adenosylme-thionine (SAM) enzyme that is involved in formation of sulphur to

carbon bonds in sactipeptides (Flühe et al., 2012), as well as an ABC transporter. In

addition, lanthipeptide response regulator genes encoding LanK and LanR are visible

upstream that indicate there may be a putative lanthipeptide structural gene present

upstream as well (Dischinger et al., 2009; Herzner et al., 2011; Siezen et al., 1996).

Lead 7 is a predicted LAP that is accompanied by a member of the YcaO like

protein family related to the cyclodehydratase C and D components (LapBotD) in its

cluster. These proteins are required together for the maturation of LAPs (Dunbar et

al., 2014, 2012). Lead 9 on the other hand contains the LanM modification gene that

defines the class II lanthipeptide group. Although these are unique leads for their

respective classes, whether or not they exhibit any antimicrobial traits requires

further analysis.

Page 192: Discovery of antimicrobial peptides active against

Chapter 5

192 | P a g e

Discussion

Another aim of this study was to identify the presence of possible additional

cryptic bacteriocin genes in the producers K. pneumoniae A7 and B. pumilus J1,

which were also analysed using the BAGEL 3 software. Results obtained provided

an interesting contrast to the wet laboratory results reported in previous Chapters.

Two predicted bacteriocins were discovered from the K. pneumoniae A7 genome,

one bottromycin and one microcin candidate. Furthermore, five putative bacteriocin

genes were identified from the producer B. pumilus J1 strain three of which encoded

head to tail cyclised peptides and two class II bacteriocins.

It isn’t uncommon for a single bacterium to produce multiple bacteriocins.

For example some strains of Streptococcus salivarius carry transferable mega

plasmids that encode multiple bacteriocins (Burton et al., 2013; Wescombe et al.,

2006), and strains of E. coli (Gordon and O’Brien, 2006), Enterococcus faecium

(Gaaloul et al., 2015; Izquierdo et al., 2008) and Lactobacillus salivarius (O’Shea et

al., 2011) have also been shown to be multiple bacteriocin producers.

Lead 14 was the most interesting of those identified from B. pumilus J1, as it

may be the elusive peptide NI03 observed during laboratory studies. This lead,

according to the BAGEL 3 software, carries significant homology to the bacteriocin

enterocin AS-48. Like peptide NI03 fraction, enterocin AS-48 is also active against

Gram positive and Gram-negative species (José et al., 2014). In addition its

intriguing that leucocyclicin Q that was revealed to share homologies with lead 14

also possesses Gram-negative activity at high concentrations (Masuda et al., 2011).

Expression of enterocin AS-48 requires the gene cluster as-48ABCC1DD1 EFGH

(Figure 5.11) (Sánchez-Hidalgo et al., 2011; van Belkum et al., 2011). BAGEL 3

also revealed two putative genes that were related to this biosynthesis pathway. One

Page 193: Discovery of antimicrobial peptides active against

Chapter 5

193 | P a g e

Discussion

of them, belongs to the DUF (domain of unknown function) 95 family, that are

observed in most head to tail cyclised peptide bacteriocin clusters (Arnison et al.,

2013; Livermore, 2007; Mu et al., 2014; van Belkum et al., 2011). Httb was also

identified and the corresponding domain in the enterocin AS-48 cluster is encoded

by as-48C and HttB is homologous to the CirB peptide from the Circularin A

biosynthetic cluster that combines with CirD to form an ABC transporter and also

contributes to self-immunity, fulfilling the same function as the AS-48B protein

encoded by the as-48B gene (Kemperman et al., 2003a; van Belkum et al., 2011).

Through manual analysis of the genome, it has been possible to identify another

protein “PROKKA_02416” present in the lead 14 gene cluster that is homologous to

the AS-48C1 protein that follows HttC and precedes the putative ABC transporter.

This protein is also involved in enterocin AS-48 production (Sánchez-Hidalgo et al.,

2011; van Belkum et al., 2011), however it was not annotated by BAGEL 3. It was

not possible to link any other remaining un-annotated genes around lead 14 to the

enterocin AS-48 cluster using PSI-BLAST, but there is sufficient evidence that this

predicted peptide is a potential bacteriocin, with anti-Gram-negative activity to

warrant experimental analysis of the cluster. However, as it has not been confirmed

that there was a single peptide species in the NI03 containing fraction, the link

between NI03 and this putative peptide is purely theoretical and must be supported

by additional data.

Page 194: Discovery of antimicrobial peptides active against

Chapter 5

194 | P a g e

Discussion

Figure 5.11 Enterocin AS-48 and cicularin A gene cluster arrangements [from (van Belkum et al., 2011)]

Analysis of the K. pneumoniae A7 draft genome yielded two putative

bacteriocins, lead 18 produced significant alignments with MccM (Vassiliadis et al.,

2010) and 19 was predicted to be a unique bottromycin candidate. The MccM

homolog, lead 18, is also homologous to MceK found in the MccE492 gene cluster

and is discussed in detail in Chapter 4. In contrast, lead 19 is an unlikely candidate

as bottromycins are known as anti-Gram positive agents, isolated from Gram

positive bacteria (Arnison et al., 2013; Hou et al., 2012; Shimamura et al., 2009;

Tanaka et al., 1966; Waisvisz et al., 1957) and K. pneumoniae A7 did not

demonstrate any antagonism towards Gram positive species, during this project.

However, homology is not a conclusive parameter and it is compelling to suggest

attempted expression of this gene. As Dirix and colleagues (2004) have shown,

microcin V (colicin V) carries motifs that are shared between Gram positive and

Gram negative bacteriocins (Havarstein et al., 1994) that don’t necessarily share a

related spectrum of activity. The LapBotD modification protein identified in the

cluster is observed to play roles in both LAP maturation and in bottromycin

synthesis. However, the bottromycin gene clusters don’t encode the C component of

this peptide. It is believed that during bottromycin synthesis LapBotD plays a role in

Page 195: Discovery of antimicrobial peptides active against

Chapter 5

195 | P a g e

Discussion

macrocyclodehydration and cyclodehydration during thiazol ring formation (Arnison

et al., 2013; Dunbar et al., 2014; Hou et al., 2012; Huo et al., 2012).

As the anaerobic bacteria mentioned above were not available in culture

collections during the current project, to test the feasibility and efficiency of a direct

cloning approach and in parallel to test the achievability of the functional production

of recombinant pumicin NI04, recombinant expression of pumicin NI04 was

attempted. Successful recombinant expression of a functional peptide confers some

significant advantages such as controlled expression and better yields (Ingham et al.,

2005; Olejnik-Schmidt et al., 2014). Expression of active recombinant bacteriocins

has been demonstrated in previous studies, though many of these are unmodified

class II bacteriocins. Some examples are epidermicin NI01 (Sandiford and Upton,

2012), enterocin P (Gutiérrez et al., 2005), enterocin A (Martín et al., 2007) and

divercin AS7 (Olejnik-Schmidt et al., 2014). For this study however, in vitro

expression was attempted as this provides two main advantages over traditional E.

coli based recombinant expression; the first of which is the omission of E. coli cells.

Many non-indigenous peptides are not secreted from the host E. coli and the cells

have to be lysed to recover the peptide, which is less than optimal as cell lysis

introduces multiple impurities that can interfere with stability and complicate the

purification process (Stiege and Erdmann, 1995; Whittaker, 2013). The second

advantage of PURExpress in vitro expression kit used in this study is the fact that it

allows relatively simple reverse purification of an untagged recombinant peptide.

Cell free expression systems are suspensions of specific cell components that are

involved in the ribosomal protein synthesis that include ribosomes, tRNA,

translation factors (10 in PURExpress) and aminoacyl-tRNA synthases (20 in

PURExpress), which, in the presence of ATP are capable of producing protein

Page 196: Discovery of antimicrobial peptides active against

Chapter 5

196 | P a g e

Discussion

(Catherine et al., 2013; Chong, 2014; Shimizu et al., 2005). In the PURExpress

system these components are all individually purified to minimise inhibition due to

nucleases (Shimizu et al., 2005).

Results showed that a peptide was purified, forming bands on SDS PAGE at

the expected mass interval for pumicin NI04. However, bands obtained were faint

and the purified fraction was inactive suggesting a sub-MIC translation reaction

yield. In vitro expression is often associated with low yields or production of a non-

functional peptide due to the in vitro transcription kit lacking the ability to support

possible post translational processes or formation of intramolecular bonds (Gutiérrez

et al 2005). However, as yet, no data are available to suggest that pumicin NI04 has

any intramolecular bonds. The issue of low yield can be addressed using a

continuous exchange system where ATP and other consumable components of the

reaction are continuously replenished during the reaction through a semipermeable

membrane, as these systems are observed to produce higher yields of the functional

product in contrast to the batch kit employed here (Chen et al., 2012). It is also

possible to introduce modifications into the in vitro expression systems such as

chaperone proteins that can enhance total protein yield as well as favourably increase

the proportion of the functional protein produced. In a study by Li and colleagues

(2014), the yield of functional peptide was improved up to 95%, through addition of

chaperons such as GroEL/ES and DnaK/DnaJ/GrpE. However, the resulting effects

were observed to be protein specific rather than universal (J. Li et al., 2014; Zawada

et al., 2011). Nevertheless, commercial systems are still relatively expensive; hence

further exploration of an E. coli based recombinant expression system is still

desirable.

Page 197: Discovery of antimicrobial peptides active against

Chapter 5

197 | P a g e

Discussion

Indeed, it is also possible that pumicin NI04 may require additional

components of its gene cluster to be expressed in order to achieve its full potential. It

is known that YukD carries homologies with ubiquitins that interact with the

proteins; it is associated with marking them for digestion, modification or transport

and is later cleaved from the peptide. While it is suggested to play a role in secretion,

the manner of the interaction is not clear and, thus, YukD may play a further role in

the functionalisation of pumicin NI04. However, this theory must be proven

experimentally through cloning of the whole gene cluster both in E. coli or in a Gram

positive host more closely related to B. pumilus.

The results obtained here prove that there is still a place for use of both in

silico and traditional antimicrobial peptide discovery approaches. While in silico

based lead discovery has proven to be extremely fast, cloning and expression of a

functional peptide, was not possible and has proven to be time consuming in the case

of pumicin NI04. However, in light of the limited success of pumicin NI04 cloning,

it is possible that this could be more successful with another system that produces

higher yields or with another lead. For example, the peptide identified as lead 3 in

the current study, which is a homolog of the two peptide bacteriocin lichenicidin.

Lichenicidin was efficaciously expressed and studied in an E. coli host (Caetano et

al., 2014).

A number of very promising putative bacteriocin leads were observed during the

genome mining efforts that should be carried forward. Nevertheless, in silico natural

product discovery is still ultimately constrained by our understanding of known

structures and motifs. The best example is the incapability of the BAGEL 3 software

to identify the pumicin NI04 gene from the B. pumillus J1 genome as it is encoded

Page 198: Discovery of antimicrobial peptides active against

Chapter 5

198 | P a g e

Discussion

by an unusual gene cluster for bacteriocins. In addition, it should also be noted that

BAGEL 3 failed to identify the well characterised MccE492 gene from the K.

pneumoniae A7 genome, highlighting the continued importance of, and need for,

classical natural product screening.

Page 199: Discovery of antimicrobial peptides active against

Chapter 6

199 | P a g e

CHAPTER 6

CONCLUDING REMARKS

Page 200: Discovery of antimicrobial peptides active against

Chapter 6

200 | P a g e

Conclusion

Natural product drug discovery has entered a new era. The techniques employed

are faster, more sensitive, more capable and more accessible than those used in

previous programmes. During this project, a number of these methods were

employed. MALDI TOF and Orbitrap MS systems were used for quick derepliction

of the peptide and rapid creation of peptide sequence tags through de novo peptide

sequencing. These techniques are quickly replacing Edman degradation to become

the primary amino acid sequencing methods (Medzihradszky and Chalkley, 2015;

Standing, 2003; Taylor and Johnson, 1997) and they are widely employed in

bacteriocin identification (Borrero et al., 2011; Sandiford and Upton, 2012; Tulini et

al., 2014). Next generation genome sequencing was also employed to quickly create

a reliable draft genome to put the peptide tags generated by MS into a genomic

context. In addition, a number of bioinformatic tools were used to analyse the

generated data including; Jalview (Waterhouse et al., 2009), Artemis (Rutherford et

al., 2000), BAGEL 3 (van Heel et al., 2013), CLC Main Workbench, BRIG (Alikhan

et al., 2011), PSI-BLAST (Altschul et al., 1997), BLAST (Altschul et al., 1990),

PROKKA (Seemann, 2014), xBASE (Chaudhuri et al., 2008), MASCOT (Perkins et

al., 1999) and Peaks studio 7 (Zhang et al., 2012).

The product of combining these cutting edge methods was the discovery of

pumicin NI04, a heat stable AMP that showed potent activity towards Gram positive

species including MRSA, VRE and Streptococcus pneumoniae. The 10kDa size of

pumicin NI04 makes it less desirable for development as an antimicrobial but the

activity observed following the trypsin and α-chymotrypsin treatment during enzyme

stability test indicates that truncated versions of pumicin NI04 are likely to produce

active leads for further development, a similar phenomenon was observed with

plantaricin Y (Chen et al., 2014). Also Salvucci and colleagues showed that short

Page 201: Discovery of antimicrobial peptides active against

Chapter 6

201 | P a g e

Conclusion

peptides derived from enterocin CRL35 demonstrated antibacterial activity (Salvucci

et al., 2007). Trypsin is used in MS/MS as it cleaves peptides at very specific

locations, and thus this information can be easily used to guide preliminary

modification studies (Steen and Mann, 2004).

Pumicin NI04 is also an interesting peptide in an evolutionary perspective as

it exhibits a close relationship to an emerging bacterial peptide family, the WXG-100

superfamily. These proteins first attracted attention following the discovery of the

ESAT-6 (EsxA) peptide of the ESX-1 secretion pathway in M. tuberculosis, which

was noted for its immunogenicity (Pallen, 2002; Sørensen et al., 1995). Pumicin

NI04 is the first WXG-100 peptide that displays antimicrobial properties and it is

most similar to peptide YukE, an EsxA homolog found in B. subtilis (Huppert et al.,

2014). The function of YukE, and many other homologs, is as of yet unknown. Thus,

screening pumicin NI04 homologues for antibacterial activity in other bacteria may

yield more AMPs; it may be the first of multiple antimicrobials to come.

Bioinformatic tools have become an integral part of microbiology. These

approaches played an important role in the discovery of pumicin NI04 and were

crucial in the investigations conducted into peptide NI05 and its producer organism

K. pneumoniae A7. Although the results suggest that peptide NI05 is not novel, the

time taken to reach a verdict concerning whether or not to further pursue the peptide

was quite short, although not straightforward; the MS-based dereplication of peptide

NI05 was not successful. However, the information that was extracted from

preliminary characterisation studies, literature research and the genome sequencing

data obtained, enabled the identification of the MccE492 cluster that was most likely

responsible for the antimicrobial activity observed.

Page 202: Discovery of antimicrobial peptides active against

Chapter 6

202 | P a g e

Conclusion

Although not the desired result, it revealed some thought provoking

information concerning the unidentified members of the Ecc492 cluster as it is

wholly conserved including truncated genes, in two K. pneumoniae species that were

separated by a significant time frame, strain RYC492 being recovered prior to 1984

while strain A7 was recovered in 2012. In addition, RYC492 was isolated from a

human stool (Lorenzo, 1984; Marcoleta et al., 2013a) and K. pneumoniae A7 came

from an HVS sample, although it could be argued that HVS flora are likely derived

from faeces.

Indeed, homology based bacteriocin and secondary metabolite search engines

are becoming established as effective in silico discovery tools (Blin et al., 2014; van

Heel et al., 2013). The opportunity presented by these tools for antimicrobial peptide

research is immense and BAGEL 3 is one of the best examples of the currently

available software that identifies RiPP based putative bacteriocins using direct and

contextual homology to known motifs associated with these antimicrobials (van Heel

et al., 2013). This tool was extensively used in the current project and 19 promising

putative bacteriocin candidates were identified from the relatively poorly explored,

hard to grow anaerobic bacteria.

Expression of pumicin NI04 was possible in the cell free expression system

employed, however the band observed on SDS-PAGE was faint, indicating a low

yield and, most likely due to this reason, no antimicrobial activity was observed from

the purified peptide solution. It is likely that an E. coli based recombinant expression

system could be more successful for the expression of pumicin NI04 or the gene

sequence of lead 14 and its accessory genes, which may encode the peptide named

Page 203: Discovery of antimicrobial peptides active against

Chapter 6

203 | P a g e

Conclusion

NI03, however as mentioned further study and data is required to prove this

connection.

This study also helped highlight some of the shortcomings of genomic

analysis and annotation software. It became apparent that major parts of the

MccE492 gene cluster, although clearly defined on the reference genome and in the

literature, was not annotated by either PROKKA software or xBASE. This was also

apparent in the annotation of the B. pumilus J1 genome, as many genes, such as the

pmnA gene, were annotated based on outdated information. This gene was annotated

as esxA of the S. aureus Ess pathway yet it carried higher homology to yukE of B.

subtilis. Also “PROKKA 03052” that was homologous to yukBA was annotated by

PROKKA as eccCa1 prior to manual correction based on UniProt entry “O69735”

that was obsolete and eccCa1 is a lower homology match. Most annotations were

based on genes that carried significantly less homology to the genes of interest than

should have been the case. Furthermore, BAGEL3 also failed to annotate the Ecc492

gene a prominent microcin, though it did pick up the truncated peptide MceK and

most of the accessory genes associated with the MccE492 biosynthetic gene cluster

(Vassiliadis et al., 2010). Overall, this demonstrates that these tools are not perfected

yet, which warrants the need to use a combination of approaches and not rely

exclusively on one tool or method.

The discovery of pumicin NI04 demonstrates the strengths of the “Waksman

platform” empowered with new techniques that were absent during its conception

(Lewis, 2013). However, there are still shortcomings, as one single rule remains

unchanged; for it to succeed the peptide of interest must be expressed under

laboratory conditions and although significantly smaller quantities are needed for

Page 204: Discovery of antimicrobial peptides active against

Chapter 6

204 | P a g e

Conclusion

dereplication (Bouslimani et al., 2014), the yield must be sufficient enough for

efficient purification of the agent. Peptide NI03 and NI05 were not obtained in such

yields under the conditions tested in this study. While it is possible to improve

peptide yield through media optimisation, these approaches are not universal and can

be rather ‘peptide specific’ (Rigali et al., 2008; Sánchez et al., 2010; Zhu et al.,

2014). Without knowing what the peptide is, it is hard to pinpoint what may trigger

its production. Screening and identification of AMPs is still a time consuming

endeavour but the current tools involved in the identification of natural peptides are

amenable to high throughput analysis. High throughput screening systems coupled

with LC-MS identification are likely to be the future of natural AMP screening

(Inglin et al., 2015; Rouse et al., 2007; Sandiford, 2014; Simone et al., 2013).

Supplementation of the “Waksman platform” with in silico tools will also help

alleviate some complications by identifying putative bacteriocins that maybe

produced by the strain under investigation and it is possible to directly express the

peptide in a recombinant host or even in a cell free environment. While tools used

for their identification will change, natural products remain to be the main source for

discovery of novel antibiotics.

Page 205: Discovery of antimicrobial peptides active against

References

205 | P a g e

REFERENCES

Aarestrup, F.M., 2005. Veterinary drug usage and antimicrobial resistance in bacteria of animal origin. Basic Clin. Pharmacol. Toxicol. 96, 271–81.

Abriouel, H., Franz, C.M.A.P., Ben Omar, N., Gálvez, A., 2011. Diversity and applications of Bacillus bacteriocins. FEMS Microbiol. Rev. 35, 201–32.

Akpe San Roman, S., Facey, P.D., Fernandez-Martinez, L., Rodriguez, C., Vallin, C., Del Sol, R., Dyson, P., 2010. A heterodimer of EsxA and EsxB is involved in sporulation and is secreted by a type VII secretion system in Streptomyces coelicolor. Microbiology 156, 1719–29.

Alikhan, N.-F., Petty, N.K., Ben Zakour, N.L., Beatson, S.A., 2011. BLAST Ring Image Generator (BRIG): simple prokaryote genome comparisons. BMC Genomics 12, 402.

Allgaier, H., Jung, G., Werner, R.G., Schneider, U., Zahner, H., 1985. Elucidation of the Structure of Epidermin, a Ribosomally Synthesized, Tetracyclic Heterodetic Polypeptide Antibiotic. Angew. Chemie Int. Ed. English 24, 1051–53.

Altschul, S.F., Gish, W., Miller, W., Myers, E.W., Lipman, D.J., 1990. Basic local alignment search tool. J. Mol. Biol. 215, 403–10.

Altschul, S.F., Madden, T.L., Schäffer, A.A., Zhang, J., Zhang, Z., Miller, W., Lipman, D.J., 1997. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 25, 3389–402.

Amersham Biosciences, n.d. Protein purification handbook.

Aminake, M.N., Schoof, S., Sologub, L., Leubner, M., Kirschner, M., Arndt, H.-D., Pradel, G., 2011. Thiostrepton and derivatives exhibit antimalarial and gametocytocidal activity by dually targeting parasite proteasome and apicoplast. Antimicrob. Agents Chemother. 55, 1338–48.

Anderssen, E.L., Diep, D.B., Nes, I.F., Eijsink, V.G., Nissen-Meyer, J., 1998. Antagonistic activity of Lactobacillus plantarum C11: two new two-peptide bacteriocins, plantaricins EF and JK, and the induction factor plantaricin A. Appl. Environ. Microbiol. 64, 2269–72.

Andrews, J.M., Working, B., Testing, S., 2001. JAC BSAC standardized disc susceptibility testing method 43–57.

Arnesen, T., 2011. Towards a functional understanding of protein N-terminal acetylation. PLoS Biol. 9, e1001074.

Page 206: Discovery of antimicrobial peptides active against

References

206 | P a g e

Arnison, P.G., Bibb, M.J., Bierbaum, G., Bowers, A.A., Bugni, T.S., Bulaj, G., Camarero, J.A., Campopiano, D.J., Challis, G.L., Clardy, J., Cotter, P.D., Craik, D.J., Dawson, M., Dittmann, E., Donadio, S., Dorrestein, P.C., Entian, K.-D., Fischbach, M.A., Garavelli, J.S., Göransson, U., Gruber, C.W., Haft, D.H., Hemscheidt, T.K., Hertweck, C., Hill, C., Horswill, A.R., Jaspars, M., Kelly, W.L., Klinman, J.P., Kuipers, O.P., Link, A.J., Liu, W., Marahiel, M.A., Mitchell, D.A., Moll, G.N., Moore, B.S., Müller, R., Nair, S.K., Nes, I.F., Norris, G.E., Olivera, B.M., Onaka, H., Patchett, M.L., Piel, J., Reaney, M.J.T., Rebuffat, S., Ross, R.P., Sahl, H.-G., Schmidt, E.W., Selsted, M.E., Severinov, K., Shen, B., Sivonen, K., Smith, L., Stein, T., Süssmuth, R.D., Tagg, J.R., Tang, G.-L., Truman, A.W., Vederas, J.C., Walsh, C.T., Walton, J.D., Wenzel, S.C., Willey, J.M., van der Donk, W.A., 2013. Ribosomally synthesized and post-translationally modified peptide natural products: overview and recommendations for a universal nomenclature. Nat. Prod. Rep. 30, 108–60.

Arranz, R., Mercado, G., Martín-Benito, J., Giraldo, R., Monasterio, O., Lagos, R., Valpuesta, J.M., 2012. Structural characterization of microcin E492 amyloid formation: Identification of the precursors. J. Struct. Biol. 178, 54–60.

Asensio, C., Pérez-Díaz, J.C., 1976. A new family of low molecular weight antibiotics from enterobacteria. Biochem. Biophys. Res. Commun. 69, 7–14.

Ates, L.S., Ummels, R., Commandeur, S., van der Weerd, R., Sparrius, M., Weerdenburg, E., Alber, M., Kalscheuer, R., Piersma, S.R., Abdallah, A.M., Abd El Ghany, M., Abdel-Haleem, A.M., Pain, A., Jiménez, C.R., Bitter, W., Houben, E.N.G., 2015. Essential Role of the ESX-5 Secretion System in Outer Membrane Permeability of Pathogenic Mycobacteria. PLOS Genet. 11, e1005190.

Attwood, P. V, Piggott, M.J., Zu, X.L., Besant, P.G., 2007. Focus on phosphohistidine. Amino Acids 32, 145–56.

Aunpad, R., Na-Bangchang, K., 2007. Pumilicin 4, a novel bacteriocin with anti-MRSA and anti-VRE activity produced by newly isolated bacteria Bacillus pumilus strain WAPB4. Curr. Microbiol. 55, 308–13.

Austin, J.W., Sanders, G., Kay, W.W., Collinson, S.K., 1998. Thin aggregative fimbriae enhance Salmonella enteritidis biofilm formation. FEMS Microbiol. Lett. 162, 295–301.

Azpiroz, M.F., Lavina, M., 2004. Involvement of Enterobactin Synthesis Pathway in Production of Microcin H47. Antimicrob. Agents Chemother. 48, 1235–41.

Azpiroz, M.F., Rodríguez, E., Laviña, M., 2001. The structure, function, and origin of the microcin H47 ATP-binding cassette exporter indicate its relatedness to that of colicin V. Antimicrob. Agents Chemother. 45, 969–72.

Page 207: Discovery of antimicrobial peptides active against

References

207 | P a g e

Bagley, M.C., Dale, J.W., Merritt, E.A., Xiong, X., 2005. Thiopeptide antibiotics. Chem. Rev. 105, 685–714.

Banerjee, S., Hansen, J.N., 1988. Structure and expression of a gene encoding the precursor of subtilin, a small protein antibiotic. J. Biol. Chem. 263, 9508–14.

Banerjee, S.P., Dora, K.C., Chowdhury, S., 2013. Detection, partial purification and characterization of bacteriocin produced by Lactobacillus brevis FPTLB3 isolated from freshwater fish: Bacteriocin from Lb. brevis FPTLB3. J. Food Sci. Technol. 50, 17–25.

Bansal, A.K., 2005. Bioinformatics in microbial biotechnology--a mini review. Microb. Cell Fact. 4, 19.

Banskota, A.H., Mcalpine, J.B., Sørensen, D., Aouidate, M., Piraee, M., Alarco, A.-M., Omura, S., Shiomi, K., Farnet, C.M., Zazopoulos, E., 2006a. Isolation and identification of three new 5-alkenyl-3,3(2H)-furanones from two streptomyces species using a genomic screening approach. J. Antibiot. (Tokyo). 59, 168–76.

Banskota, A.H., Mcalpine, J.B., Sørensen, D., Ibrahim, A., Aouidate, M., Piraee, M., Alarco, A.-M., Farnet, C.M., Zazopoulos, E., 2006b. Genomic analyses lead to novel secondary metabolites. Part 3. ECO-0501, a novel antibacterial of a new class. J. Antibiot. (Tokyo). 59, 533–42.

Baptista, C., Barreto, H.C., São-José, C., 2013. High levels of DegU-P activate an Esat-6-like secretion system in Bacillus subtilis. PLoS One 8, e67840.

Barber, M., 1961. Methicillin-resistant staphylococci. J. Clin. Pathol. 14, 385–93.

Barbour, A., Philip, K., Muniandy, S., 2013. Enhanced production, purification, characterization and mechanism of action of salivaricin 9 lantibiotic produced by Streptococcus salivarius NU10. PLoS One 8, e77751.

Barguigua, A., El Otmani, F., Lakbakbi El Yaagoubi, F., Talmi, M., Zerouali, K., Timinouni, M., 2013. First report of a Klebsiella pneumoniae strain coproducing NDM-1, VIM-1 and OXA-48 carbapenemases isolated in Morocco. APMIS 121, 675–7.

Barriere, S.L., 2015. Clinical, economic and societal impact of antibiotic resistance. Expert Opin. Pharmacother. 16, 151–53.

Beatson, S.A., Sloan, G.L., Simmonds, R.S., 1998. Zoocin A immunity factor: a femA -like gene found in a group C streptococcus. FEMS Microbiol. Lett. 163, 73–77.

Begley, M., Cotter, P.D., Hill, C., Ross, R.P., 2009. Identification of a novel two-peptide lantibiotic, lichenicidin, following rational genome mining for LanM proteins. Appl. Environ. Microbiol. 75, 5451–60.

Page 208: Discovery of antimicrobial peptides active against

References

208 | P a g e

Bentur, H.N., Dalzell, A.M., Riordan, F.A.I., 2007. Central venous catheter infection with Bacillus pumilus in an immunocompetent child: a case report. Ann. Clin. Microbiol. Antimicrob. 6, 12.

Bérdy, J., 2012. Thoughts and facts about antibiotics: where we are now and where we are heading. J. Antibiot. (Tokyo). 65, 385–95.

Berrazeg, M., Diene, S., Medjahed, L., Parola, P., Drissi, M., Raoult, D., Rolain, J., 2014. New Delhi Metallo-beta-lactamase around the world: an eReview using Google Maps. Euro Surveill. 19.

Besser, R.E., 2003. Antimicrobial Prescribing in the United States: Good News, Bad News. Ann. Intern. Med. 138, 605.

Bhugaloo-Vial, P., Douliez, J.P., Moll, D., Dousset, X., Boyaval, P., Marion, D., 1999. Delineation of key amino acid side chains and peptide domains for antimicrobial properties of divercin V41, a pediocin-like bacteriocin secreted by Carnobacterium divergens V41. Appl. Environ. Microbiol. 65, 2895–900.

Bieler, S., Estrada, L., Lagos, R., Baeza, M., Castilla, J., Soto, C., 2005. Amyloid formation modulates the biological activity of a bacterial protein. J. Biol. Chem. 280, 26880–5.

Bierbaum, G., Brotz, H., Koller, K.-P., Sahl, H.-G., 1995. Cloning, sequencing and production of the lantibiotic mersacidin. FEMS Microbiol. Lett. 127, 121–26.

Blin, K., Kazempour, D., Wohlleben, W., Weber, T., 2014. Improved lanthipeptide detection and prediction for antiSMASH. PLoS One 9, e89420.

Blin, K., Medema, M.H., Kazempour, D., Fischbach, M.A., Breitling, R., Takano, E., Weber, T., 2013. antiSMASH 2.0--a versatile platform for genome mining of secondary metabolite producers. Nucleic Acids Res. 41, W204–12.

Boakes, S., Cortés, J., Appleyard, A.N., Rudd, B.A.M., Dawson, M.J., 2009. Organization of the genes encoding the biosynthesis of actagardine and engineering of a variant generation system. Mol. Microbiol. 72, 1126–36.

Borrero, J., Brede, D.A., Skaugen, M., Diep, D.B., Herranz, C., Nes, I.F., Cintas, L.M., Hernández, P.E., 2011. Characterization of garvicin ML, a novel circular bacteriocin produced by Lactococcus garvieae DCC43, isolated from mallard ducks (Anas platyrhynchos). Appl. Environ. Microbiol. 77, 369–73.

Bou, G., Cerveró, G., Domínguez, M.A., Quereda, C., Martínez-Beltrán, J., 2000. Characterization of a nosocomial outbreak caused by a multiresistant Acinetobacter baumannii strain with a carbapenem-hydrolyzing enzyme: high-level carbapenem resistance in A. baumannii is not due solely to the presence of beta-lactamases. J. Clin. Microbiol. 38, 3299–305.

Page 209: Discovery of antimicrobial peptides active against

References

209 | P a g e

Boucher, H.W., Talbot, G.H., Bradley, J.S., Edwards, J.E., Gilbert, D., Rice, L.B., Scheld, M., Spellberg, B., Bartlett, J., 2009. Bad bugs, no drugs: no ESKAPE! An update from the Infectious Diseases Society of America. Clin. Infect. Dis. 48, 1–12.

Boucher, H.W., Talbot, G.H., Jr, D.K.B., Bradley, J., Guidos, R.J., Jones, R.N., Benjamin, D.K., Bradley, J., Guidos, R.J., Jones, R.N., Murray, B.E., Bonomo, R. a, Gilbert, D., 2013. 10 x ’20 Progress--development of new drugs active against gram-negative bacilli: an update from the Infectious Diseases Society of America. Clin. Infect. Dis. 56, 1685–94.

Bouslimani, A., Sanchez, L.M., Garg, N., Dorrestein, P.C., 2014. Mass spectrometry of natural products: current, emerging and future technologies. Nat. Prod. Rep. 31, 718–29.

Braun, V., Patzer, S.I., Hantke, K., 2002. Ton-dependent colicins and microcins: modular design and evolution. Biochimie 84, 365–80.

Braun, V., Pilsl, H., Gross, P., 1994. Colicins: structures, modes of action, transfer through membranes, and evolution. Arch. Microbiol. 161, 199–206.

Brede, D.A., Faye, T., Johnsborg, O., Odegård, I., Nes, I.F., Holo, H., 2004. Molecular and genetic characterization of propionicin F, a bacteriocin from Propionibacterium freudenreichii. Appl. Environ. Microbiol. 70, 7303–10.

Breton, R.C., Reynolds, W.F., 2013. Using NMR to identify and characterize natural products. Nat. Prod. Rep. 30, 501–24.

Burroughs, A.M., Iyer, L.M., Aravind, L., 2011. Functional diversification of the RING finger and other binuclear treble clef domains in prokaryotes and the early evolution of the ubiquitin system. Mol. Biosyst. 7, 2261–77.

Burton, J.P., Wescombe, P.A., Macklaim, J.M., Chai, M.H.C., Macdonald, K., Hale, J.D.F., Tagg, J., Reid, G., Gloor, G.B., Cadieux, P.A., 2013. Persistence of the oral probiotic Streptococcus salivarius M18 is dose dependent and megaplasmid transfer can augment their bacteriocin production and adhesion characteristics. PLoS One 8, e65991.

Burts, M.L., DeDent, A.C., Missiakas, D.M., 2008. EsaC substrate for the ESAT-6 secretion pathway and its role in persistent infections of Staphylococcus aureus. Mol. Microbiol. 69, 736–46.

Burts, M.L., Williams, W. a, DeBord, K., Missiakas, D.M., 2005. EsxA and EsxB are secreted by an ESAT-6-like system that is required for the pathogenesis of Staphylococcus aureus infections. Proc. Natl. Acad. Sci. U. S. A. 102, 1169–74.

Bush, K., 2010. Alarming β-lactamase-mediated resistance in multidrug-resistant Enterobacteriaceae. Curr. Opin. Microbiol. 13, 558–64.

Page 210: Discovery of antimicrobial peptides active against

References

210 | P a g e

Butler, M.S., Blaskovich, M.A., Cooper, M.A., 2013. Antibiotics in the clinical pipeline in 2013. J. Antibiot. (Tokyo). 66, 571–91.

Caetano, T., Barbosa, J., Möesker, E., Süssmuth, R.D., Mendo, S., 2014. Bioengineering of lanthipeptides in Escherichia coli: assessing the specificity of lichenicidin and haloduracin biosynthetic machinery. Res. Microbiol. 165, 600–4.

Cantón, R., González-Alba, J.M., Galán, J.C., 2012. CTX-M Enzymes: Origin and Diffusion. Front. Microbiol. 3, 110.

Cascales, E., Buchanan, S.K., Duché, D., Kleanthous, C., Lloubès, R., Postle, K., Riley, M., Slatin, S., Cavard, D., 2007. Colicin biology. Microbiol. Mol. Biol. Rev. 71, 158–229.

Catherine, C., Lee, K.-H., Oh, S.-J., Kim, D.-M., 2013. Cell-free platforms for flexible expression and screening of enzymes. Biotechnol. Adv. 31, 797–803.

Cavalcante, F.S., Ferreira, D. de C., Chamon, R.C., da Costa, T.M., Maia, F., Barros, E.M., Dantas, T.S., Dos Santos, K.R.N., 2014. Daptomycin and methicillin-resistant Staphylococcus aureus isolated from a catheter-related bloodstream infection: a case report. BMC Res. Notes 7, 759.

Cavera, V.L., Arthur, T.D., Kashtanov, D., Chikindas, M.L., 2015. Bacteriocins and their position in the next wave of conventional antibiotics. Int. J. Antimicrob. Agents 46, 494–501.

Challis, G.L., Ravel, J., 2000. Coelichelin, a new peptide siderophore encoded by the Streptomyces coelicolor genome: structure prediction from the sequence of its non-ribosomal peptide synthetase. FEMS Microbiol. Lett. 187, 111–14.

Chambers, H.F., Deleo, F.R., 2009. Waves of resistance: Staphylococcus aureus in the antibiotic era. Nat. Rev. Microbiol. 7, 629–41.

Champion, P.A.D., Stanley, S.A., Champion, M.M., Brown, E.J., Cox, J.S., 2006. C-terminal signal sequence promotes virulence factor secretion in Mycobacterium tuberculosis. Science 313, 1632–6.

Chan, Y.-C., Wu, J.-L., Wu, H.-P., Tzeng, K.-C., Chuang, D.-Y., 2011. Cloning, purification, and functional characterization of Carocin S2, a ribonuclease bacteriocin produced by Pectobacterium carotovorum. BMC Microbiol. 11, 99.

Chapman, M.R., Robinson, L.S., Pinkner, J.S., Roth, R., Heuser, J., Hammar, M., Normark, S., Hultgren, S.J., 2002. Role of Escherichia coli curli operons in directing amyloid fiber formation. Science 295, 851–5.

Chaudhuri, R.R., Loman, N.J., Snyder, L.A.S., Bailey, C.M., Stekel, D.J., Pallen, M.J., 2008. xBASE2: a comprehensive resource for comparative bacterial

Page 211: Discovery of antimicrobial peptides active against

References

211 | P a g e

genomics. Nucleic Acids Res. 36, D543–6.

Chen, H., Yan, X., Tian, F., Song, Y., Chen, Y.Q., Zhang, H., Chen, W., 2012. Cloning, expression, and identification of a novel class IIa bacteriocin in the Escherichia coli cell-free protein expression system. Biotechnol. Lett. 34, 359–64.

Chen, Y.S., Wang, Y.C., Chow, Y.S., Yanagida, F., Liao, C.C., Chiu, C.M., 2014. Purification and characterization of plantaricin Y, a novel bacteriocin produced by Lactobacillus plantarum 510. Arch. Microbiol. 196, 193–99.

Chong, S., 2014. Overview of cell-free protein synthesis: historic landmarks, commercial systems, and expanding applications. Curr. Protoc. Mol. Biol. 108, 16.30.1–16.30.11.

Cintas, L.M., Casaus, P., Håvarstein, L.S., Hernández, P.E., Nes, I.F., 1997. Biochemical and genetic characterization of enterocin P, a novel sec-dependent bacteriocin from Enterococcus faecium P13 with a broad antimicrobial spectrum. Appl. Environ. Microbiol. 63, 4321–30.

Claessen, D., Rink, R., de Jong, W., Siebring, J., de Vreugd, P., Boersma, F.G.H., Dijkhuizen, L., Wosten, H.A.B., 2003. A novel class of secreted hydrophobic proteins is involved in aerial hyphae formation in Streptomyces coelicolor by forming amyloid-like fibrils. Genes Dev. 17, 1714–26.

Clatworthy, A.E., Pierson, E., Hung, D.T., 2007. Targeting virulence: a new paradigm for antimicrobial therapy. Nat. Chem. Biol. 3, 541–8.

CLSI, 2012. Methods for dilution antimicrobial susceptibility tests for bacteria that grow aerobically; approved standard—ninth edition. CLSI Doc.

Coates, A.R.M., Halls, G., Hu, Y., 2011. Novel classes of antibiotics or more of the same? Br. J. Pharmacol. 163, 184–94.

Cochrane, S.A., Vederas, J.C., 2014. Lipopeptides from Bacillus and Paenibacillus spp.: A Gold Mine of Antibiotic Candidates. Med. Res. Rev.

Cole, C., Barber, J.D., Barton, G.J., 2008. The Jpred 3 secondary structure prediction server. Nucleic Acids Res. 36, W197–201.

Collin, F., Thompson, R.E., Jolliffe, K.A., Payne, R.J., Maxwell, A., 2013. Fragments of the bacterial toxin microcin B17 as gyrase poisons. PLoS One 8, e61459.

Converse, S.E., Cox, J.S., 2005. A protein secretion pathway critical for Mycobacterium tuberculosis virulence is conserved and functional in Mycobacterium smegmatis. J. Bacteriol. 187, 1238–45.

Cornaglia, G., 2009. Fighting infections due to multidrug-resistant Gram-positive

Page 212: Discovery of antimicrobial peptides active against

References

212 | P a g e

pathogens. Clin. Microbiol. Infect. 15, 209–11.

Cortes, J., Haydock, S.F., Roberts, G.A., Bevitt, D.J., Leadlay, P.F., 1990. An unusually large multifunctional polypeptide in the erythromycin-producing polyketide synthase of Saccharopolyspora erythraea. Nature 348, 176–8.

Cosgrove, S.E., 2006. The relationship between antimicrobial resistance and patient outcomes: mortality, length of hospital stay, and health care costs. Clin. Infect. Dis. 42 Suppl 2, S82–9.

Cotter, P.D., Hill, C., Ross, R.P., 2005. Bacteriocins: developing innate immunity for food. Nat. Rev. Microbiol. 3, 777–88.

Cotter, P.D., Ross, R.P., Hill, C., 2013. Bacteriocins - a viable alternative to antibiotics? Nat. Rev. Microbiol. 11, 95–105.

Crone, W.J.K., Leeper, F.J., Truman, A.W., 2012. Identification and characterisation of the gene cluster for the anti-MRSA antibiotic bottromycin: expanding the biosynthetic diversity of ribosomal peptides. Chem. Sci. 3, 3516.

Crost, E.H., Ajandouz, E.H., Villard, C., Geraert, P. a, Puigserver, a, Fons, M., 2011. Ruminococcin C, a new anti-Clostridium perfringens bacteriocin produced in the gut by the commensal bacterium Ruminococcus gnavus E1. Biochimie 93, 1487–94.

Cuff, J.A., Barton, G.J., 2000. Application of multiple sequence alignment profiles to improve protein secondary structure prediction. Proteins 40, 502–11.

Cui, Y., Zhang, C., Wang, Y., Shi, J., Zhang, L., Ding, Z., Qu, X., Cui, H., 2012. Class IIa bacteriocins: diversity and new developments. Int. J. Mol. Sci. 13, 16668–707.

Cuozzo, S.A., Sesma, F., Palacios, J.M., RuÃ-z Holgado, Aã-.P., Raya, R.R., 2000. Identification and nucleotide sequence of genes involved in the synthesis of lactocin 705, a two-peptide bacteriocin from Lactobacillus casei CRL 705. FEMS Microbiol. Lett. 185, 157–61.

Curtis, T.P., Head, I.M., Lunn, M., Woodcock, S., Schloss, P.D., Sloan, W.T., 2006. What is the extent of prokaryotic diversity? Philos. Trans. R. Soc. Lond. B. Biol. Sci. 361, 2023–37.

Dabard, J., Bridonneau, C., Phillipe, C., Anglade, P., Molle, D., Nardi, M., Ladiré, M., Girardin, H., Marcille, F., Gomez, A., Fons, M., 2001. Ruminococcin A, a new lantibiotic produced by a Ruminococcus gnavus strain isolated from human feces. Appl. Environ. Microbiol. 67, 4111–8.

Darilmaz, D.O., Beyatli, Y., 2012. Acid-bile, antibiotic resistance and inhibitory properties of propionibacteria isolated from Turkish traditional home-made

Page 213: Discovery of antimicrobial peptides active against

References

213 | P a g e

cheeses. Anaerobe 18, 122–7.

Davies, D.L., Falkiner, F.R., Hardy, K.G., 1981. Colicin V production by clinical isolates of Escherichia coli. Infect. Immun. 31, 574–9.

Davies, J., 2011. How to discover new antibiotics: harvesting the parvome. Curr. Opin. Chem. Biol. 15, 5–10.

Davies, J., Davies, D., 2010. Origins and evolution of antibiotic resistance. Microbiol. Mol. Biol. Rev. 74, 417–33.

de Jong, A., van Heel, A.J., Kok, J., Kuipers, O.P., 2010. BAGEL2: mining for bacteriocins in genomic data. Nucleic Acids Res. 38, W647–51.

de Jong, A., van Hijum, S. a F.T., Bijlsma, J.J.E., Kok, J., Kuipers, O.P., 2006. BAGEL: a web-based bacteriocin genome mining tool. Nucleic Acids Res. 34, W273–9.

De Jonge, M.I., Pehau-Arnaudet, G., Fretz, M.M., Romain, F., Bottai, D., Brodin, P., Honoré, N., Marchal, G., Jiskoot, W., England, P., Cole, S.T., Brosch, R., 2007. ESAT-6 from Mycobacterium tuberculosis dissociates from its putative chaperone CFP-10 under acidic conditions and exhibits membrane-lysing activity. J. Bacteriol. 189, 6028–34.

De Vriese, A.S., Vandecasteele, S.J., 2014. Vancomycin: the tale of the vanquisher and the pyrrhic victory. Perit. Dial. Int. 34, 154–61.

Del Fabbro, C., Scalabrin, S., Morgante, M., Giorgi, F.M., 2013. An extensive evaluation of read trimming effects on Illumina NGS data analysis. PLoS One 8, e85024.

Delcour, A.H., 2009. Outer membrane permeability and antibiotic resistance. Biochim. Biophys. Acta 1794, 808–16.

Demirev, P.A., Lin, J.S., Pineda, F.J., Fenselau, C., 2001. Bioinformatics and Mass Spectrometry for Microorganism Identification: Proteome-Wide Post-Translational Modifications and Database Search Algorithms for Characterization of Intact H. p ylori. Anal. Chem. 73, 4566–73.

Destoumieux-Garzón, D., Peduzzi, J., Thomas, X., Djediat, C., Rebuffat, S., 2006. Parasitism of iron-siderophore receptors of Escherichia coli by the siderophore-peptide microcin E492m and its unmodified counterpart. Biometals 19, 181–91.

Destoumieux-Garzón, D., Thomas, X., Santamaria, M., Goulard, C., Barthélémy, M., Boscher, B., Bessin, Y., Molle, G., Pons, A.-M., Letellier, L., Peduzzi, J., Rebuffat, S., 2003. Microcin E492 antibacterial activity: evidence for a TonB-dependent inner membrane permeabilization on Escherichia coli. Mol. Microbiol. 49, 1031–41.

Page 214: Discovery of antimicrobial peptides active against

References

214 | P a g e

Devi, K.P., Nisha, S.A., Sakthivel, R., Pandian, S.K., 2010. Eugenol (an essential oil of clove) acts as an antibacterial agent against Salmonella typhi by disrupting the cellular membrane. J. Ethnopharmacol. 130, 107–15.

Diep, D.B., Skaugen, M., Salehian, Z., Holo, H., Nes, I.F., 2007. Common mechanisms of target cell recognition and immunity for class II bacteriocins. Proc. Natl. Acad. Sci. U. S. A. 104, 2384–9.

Dirix, G., Monsieurs, P., Dombrecht, B., Daniels, R., Marchal, K., Vanderleyden, J., Michiels, J., 2004. Peptide signal molecules and bacteriocins in Gram-negative bacteria: a genome-wide in silico screening for peptides containing a double-glycine leader sequence and their cognate transporters. Peptides 25, 1425–40.

Dischinger, J., Josten, M., Szekat, C., Sahl, H.-G., Bierbaum, G., 2009. Production of the novel two-peptide lantibiotic lichenicidin by Bacillus licheniformis DSM 13. PLoS One 4, e6788.

Dobson, C.M., 1999. Protein misfolding, evolution and disease. Trends Biochem. Sci. 24, 329–32.

Dobson, C.M., 2004. Chemical space and biology. Nature 432, 824–8.

Donadio, S., Staver, M.J., McAlpine, J.B., Swanson, S.J., Katz, L., 1991. Modular organization of genes required for complex polyketide biosynthesis. Science 252, 675–9.

Dortet, L., Poirel, L., Nordmann, P., 2014. Worldwide dissemination of the NDM-type carbapenemases in Gram-negative bacteria. Biomed Res. Int. 2014, 249856.

Doumith, M., Dhanji, H., Ellington, M.J., Hawkey, P., Woodford, N., 2012. Characterization of plasmids encoding extended-spectrum β-lactamases and their addiction systems circulating among Escherichia coli clinical isolates in the UK. J. Antimicrob. Chemother. 67, 878–85.

Ducros, V.M., Lewis, R.J., Verma, C.S., Dodson, E.J., Leonard, G., Turkenburg, J.P., Murshudov, G.N., Wilkinson, A.J., Brannigan, J.A., 2001. Crystal structure of GerE, the ultimate transcriptional regulator of spore formation in Bacillus subtilis. J. Mol. Biol. 306, 759–71.

Dueholm, M.S., Søndergaard, M.T., Nilsson, M., Christiansen, G., Stensballe, A., Overgaard, M.T., Givskov, M., Tolker-Nielsen, T., Otzen, D.E., Nielsen, P.H., 2013. Expression of Fap amyloids in Pseudomonas aeruginosa, P. fluorescens, and P. putida results in aggregation and increased biofilm formation. Microbiologyopen 2, 365–82.

Dunbar, K.L., Chekan, J.R., Cox, C.L., Burkhart, B.J., Nair, S.K., Mitchell, D.A., 2014. Discovery of a new ATP-binding motif involved in peptidic azoline

Page 215: Discovery of antimicrobial peptides active against

References

215 | P a g e

biosynthesis. Nat. Chem. Biol. 10, 823–9.

Dunbar, K.L., Melby, J.O., Mitchell, D.A., 2012. YcaO domains use ATP to activate amide backbones during peptide cyclodehydrations. Nat. Chem. Biol. 8, 569–75.

Dupuy, B., Mani, N., Katayama, S., Sonenshein, A.L., 2005. Transcription activation of a UV-inducible Clostridium perfringens bacteriocin gene by a novel sigma factor. Mol. Microbiol. 55, 1196–206.

Duquesne, S., Destoumieux-Garzón, D., Peduzzi, J., Rebuffat, S., 2007a. Microcins, gene-encoded antibacterial peptides from enterobacteria. Nat. Prod. Rep. 24, 708–34.

Duquesne, S., Petit, V., Peduzzi, J., Rebuffat, S., 2007b. Structural and functional diversity of microcins, gene-encoded antibacterial peptides from enterobacteria. J. Mol. Microbiol. Biotechnol. 13, 200–9.

Eberhart, L.J., Deringer, J.R., Brayton, K.A., Sawant, A.A., Besser, T.E., Call, D.R., 2012. Characterization of a novel microcin that kills enterohemorrhagic Escherichia coli O157:H7 and O26. Appl. Environ. Microbiol. 78, 6592–9.

Eijsink, V.G., Skeie, M., Middelhoven, P.H., Brurberg, M.B., Nes, I.F., 1998. Comparative studies of class IIa bacteriocins of lactic acid bacteria. Appl. Environ. Microbiol. 64, 3275–81.

Ekinci, F.Y., Gurel, M., 2008. Effect of using propionic acid bacteria as an adjunct culture in yogurt production. J. Dairy Sci. 91, 892–9.

Eumkeb, G., Siriwong, S., Thumanu, K., 2012. Synergistic activity of luteolin and amoxicillin combination against amoxicillin-resistant Escherichia coli and mode of action. J. Photochem. Photobiol. B Biol. 117, 247–53.

Felnagle, E.E.A., Jackson, E.E.E., Chan, Y.A., Podevels, A.M., Berti, A.D., McMahon, M.D., Thomas, M.G., 2008. Nonribosomal peptide synthetases involved in the production of medically relevant natural products. Mol. Pharm. 5, 191–211.

Fessler, A.T., Calvo, N., Gutiérrez, N., Bellido, J.L.M., Fajardo, M., Garduno, E., Monecke, S., Ehricht, R., Kadlec, K., Schwarz, S., 2014. Cfr-mediated linezolid resistance in methicillin-resistant Staphylococcus aureus and Staphylococcus haemolyticus associated with clinical infections in humans: Two case reports. J. Antimicrob. Chemother. 69, 268–70.

Filya, I., Sucu, E., Karabulut, A., 2004. The effect of Propionibacterium acidipropionici, with or without Lactobacillus plantarum, on the fermentation and aerobic stability of wheat, sorghum and maize silages. J. Appl. Microbiol. 97, 818–26.

Page 216: Discovery of antimicrobial peptides active against

References

216 | P a g e

Fischbach, M.A., Lin, H., Liu, D.R., Walsh, C.T., 2006. How pathogenic bacteria evade mammalian sabotage in the battle for iron. Nat. Chem. Biol. 2, 132–8.

Fischetti, V. a, 2010. Bacteriophage endolysins: a novel anti-infective to control Gram-positive pathogens. Int. J. Med. Microbiol. 300, 357–62.

Flühe, L., Knappe, T.A., Gattner, M.J., Schäfer, A., Burghaus, O., Linne, U., Marahiel, M.A., 2012. The radical SAM enzyme AlbA catalyzes thioether bond formation in subtilosin A. Nat. Chem. Biol. 8, 350–7.

Franz, C.M.A.P., Grube, A., Herrmann, A., Abriouel, H., Starke, J., Lombardi, A., Tauscher, B., Holzapfel, W.H., 2002. Biochemical and Genetic Characterization of the Two-Peptide Bacteriocin Enterocin 1071 Produced by Enterococcus faecalis FAIR-E 309. Appl. Environ. Microbiol. 68, 2550–4.

Frottin, F., Martinez, A., Peynot, P., Mitra, S., Holz, R.C., Giglione, C., Meinnel, T., 2006. The proteomics of N-terminal methionine cleavage. Mol. Cell. Proteomics 5, 2336–49.

Fujita, K., Ichimasa, S., Zendo, T., Koga, S., Yoneyama, F., Nakayama, J., Sonomoto, K., 2007. Structural analysis and characterization of lacticin Q, a novel bacteriocin belonging to a new family of unmodified bacteriocins of gram-positive bacteria. Appl. Environ. Microbiol. 73, 2871–7.

Fyans, J.K., Bignell, D., Loria, R., Toth, I., Palmer, T., 2013. The ESX/type VII secretion system modulates development, but not virulence, of the plant pathogen Streptomyces scabies. Mol. Plant Pathol. 14, 119–30.

Gaaloul, N., ben Braiek, O., Hani, K., Volski, A., Chikindas, M.L., Ghrairi, T., 2015. Isolation and characterization of large spectrum and multiple bacteriocin-producing Enterococcus faecium strain from raw bovine milk. J. Appl. Microbiol. 118, 343–55.

Gabrielsen, C., Brede, D.A., Nes, I.F., Diep, D.B., 2014. Circular bacteriocins: biosynthesis and mode of action. Appl. Environ. Microbiol. 80, 6854–62.

Gaillard-Gendron, S., Vignon, D., Cottenceau, G., Graber, M., Zorn, N., van Dorsselaer, A., AM, P., 2000. Isolation, purification and partial amino acid sequence of a highly hydrophobic new microcin named microcin L produced by Escherichia coli. FEMS Microbiol. Lett. 193, 95–8.

Gao, L.-Y., Guo, S., McLaughlin, B., Morisaki, H., Engel, J.N., Brown, E.J., 2004. A mycobacterial virulence gene cluster extending RD1 is required for cytolysis, bacterial spreading and ESAT-6 secretion. Mol. Microbiol. 53, 1677–93.

Garces, A., Atmakuri, K., Chase, M.R., Woodworth, J.S., Krastins, B., Rothchild, A.C., Ramsdell, T.L., Lopez, M.F., Behar, S.M., Sarracino, D.A., Fortune, S.M., 2010. EspA acts as a critical mediator of ESX1-dependent virulence in

Page 217: Discovery of antimicrobial peptides active against

References

217 | P a g e

Mycobacterium tuberculosis by affecting bacterial cell wall integrity. PLoS Pathog. 6, e1000957.

Garcia De Gonzalo, C. V, Zhu, L., Oman, T.J., van der Donk, W.A., 2014. NMR structure of the S-linked glycopeptide sublancin 168. ACS Chem. Biol. 9, 796–801.

Garneau, S., Martin, N.I., Vederas, J.C., 2002. Two-peptide bacteriocins produced by lactic acid bacteria. Biochimie 84, 577–92.

Genilloud, O., 2014. The re-emerging role of microbial natural products in antibiotic discovery. Antonie van Leeuwenhoek, Int. J. Gen. Mol. Microbiol. 106, 173–88.

Gillor, O., Kirkup, B., Riley, M., 2004. Colicins and microcins: the next generation antimicrobials. Adv. Appl. Microbiol. 54, 129–46.

Godballe, T., Nilsson, L.L., Petersen, P.D., Jenssen, H., 2011. Antimicrobial β-peptides and α-peptoids. Chem. Biol. Drug Des. 77, 107–16.

Gomez, A., Ladiré, M., Marcille, F., Fons, M., 2002. Trypsin mediates growth phase-dependent transcriptional tegulation of genes involved in biosynthesis of ruminococcin A, a lantibiotic produced by a Ruminococcus gnavus strain from a human intestinal microbiota. J. Bacteriol. 184, 18–28.

Gomez-Escribano, J.P., Song, L., Bibb, M.J., Challis, G.L., 2012. Posttranslational β-methylation and macrolactamidination in the biosynthesis of the bottromycin complex of ribosomal peptide antibiotics. Chem. Sci. 3, 3522.

Gonzalez, C.F., Kunka, B.S., 1987. Plasmid-Associated Bacteriocin Production and Sucrose Fermentation in Pediococcus acidilactici. Appl. Environ. Microbiol. 53, 2534–8.

Gordon, D.M., O’Brien, C.L., 2006. Bacteriocin diversity and the frequency of multiple bacteriocin production in Escherichia coli. Microbiology 152, 3239–44.

Goto, Y., Li, B., Claesen, J., Shi, Y., Bibb, M.J., van der Donk, W.A., 2010. Discovery of unique lanthionine synthetases reveals new mechanistic and evolutionary insights. PLoS Biol. 8, e1000339.

Goto, Y., Ökesli, A., van der Donk, W.A., 2011. Mechanistic studies of Ser/Thr dehydration catalyzed by a member of the LanL lanthionine synthetase family. Biochemistry 50, 891–8.

Gravesen, A., Ramnath, M., Rechinger, K.B., Andersen, N., Jansch, L., Hechard, Y., Hastings, J.W., Knochel, S., 2002. High-level resistance to class IIa bacteriocins is associated with one general mechanism in Listeria monocytogenes.

Page 218: Discovery of antimicrobial peptides active against

References

218 | P a g e

Microbiology 148, 2361–9.

Green, M.R., Sambrook, J., 2001. Molecular cloning: A laboratory manual, 3rd ed. Cold spring harbour laboratory press, Cold Spring Harbor, NY.

Gross, F., Luniak, N., Perlova, O., Gaitatzis, N., Jenke-Kodama, H., Gerth, K., Gottschalk, D., Dittmann, E., Müller, R., 2006. Bacterial type III polyketide synthases: phylogenetic analysis and potential for the production of novel secondary metabolites by heterologous expression in pseudomonads. Arch. Microbiol. 185, 28–38.

Guan, L., Liu, Q., Li, C., Zhang, Y., 2013. Development of a Fur-dependent and tightly regulated expression system in Escherichia coli for toxic protein synthesis. BMC Biotechnol. 13, 25.

Guinn, K.M., Hickey, M.J., Mathur, S.K., Zakel, K.L., Grotzke, J.E., Lewinsohn, D.M., Smith, S., Sherman, D.R., 2004. Individual RD1-region genes are required for export of ESAT-6/CFP-10 and for virulence of Mycobacterium tuberculosis. Mol. Microbiol. 51, 359–70.

Guo, J., Jia, R., 2014. A novel inducible expression system for the functional study of toxic gene in bacteria. World J. Microbiol. Biotechnol. 30, 1527–31.

Gutiérrez, J., Criado, R., Citti, R., Martín, M., Herranz, C., Nes, I.F., Cintas, L.M., Hernández, P.E., 2005. Cloning, production and functional expression of enterocin P, a sec-dependent bacteriocin produced by Enterococcus faecium P13, in Escherichia coli. Int. J. Food Microbiol. 103, 239–50.

Hammami, R., Bédard, F., Gomaa, A., Subirade, M., Biron, E., Fliss, I., 2015. Lasso-inspired peptides with distinct antibacterial mechanisms. Amino Acids 47, 417–28.

Hammami, R., Zouhir, A., Le Lay, C., Ben Hamida, J., Fliss, I., 2010. BACTIBASE second release: a database and tool platform for bacteriocin characterization. BMC Microbiol. 10, 22.

Händel, N., Otte, S., Jonker, M., Brul, S., ter Kuile, B.H., 2015. Factors that affect transfer of the IncI1 β-lactam resistance plasmid pESBL-283 between E. coli strains. PLoS One 10, e0123039.

Harbarth, S., Cosgrove, S., Carmeli, Y., 2002. Effects of antibiotics on nosocomial epidemiology of vancomycin-resistant enterococci. Antimicrob. Agents Chemother. 46, 1619–28.

Harrigan, G.G., Goetz, G.H., 2005. Chemical and biological integrity in natural products screening. Comb. Chem. High Throughput Screen. 8, 529–34.

Harvey, A.L., 2008. Natural products in drug discovery. Drug Discov. Today 13,

Page 219: Discovery of antimicrobial peptides active against

References

219 | P a g e

894–901.

Hatheway, C.L., 1990. Toxigenic clostridia. Clin. Microbiol. Rev. 3, 66–98.

Havarstein, L.S., Holo, H., Nes, I.F., 1994. The leader peptide of colicin V shares consensus sequences with leader peptides that are common among peptide bacteriocins produced by Gram-positive bacteria. Microbiology 140, 2383–9.

Héchard, Y., Sahl, H.-G., 2002. Mode of action of modified and unmodified bacteriocins from Gram-positive bacteria. Biochimie 84, 545–57.

Hede, K., 2014. Antibiotic resistance: An infectious arms race. Nature 509, S2–3.

Helbig, A.O., Gauci, S., Raijmakers, R., van Breukelen, B., Slijper, M., Mohammed, S., Heck, A.J.R., 2010. Profiling of N-acetylated protein termini provides in-depth insights into the N-terminal nature of the proteome. Mol. Cell. Proteomics 9, 928–39.

Heng, N.C.K., Tagg, J.R., 2006. What’s in a name? Class distinction for bacteriocins. Nat. Rev. Microbiol. 4.

Hershko, A., Ciechanover, A., 1998. The ubiquitin system. Annu. Rev. Biochem. 67, 425–79.

Herzner, A.M., Dischinger, J., Szekat, C., Josten, M., Schmitz, S., Yakéléba, A., Reinartz, R., Jansen, A., Sahl, H.-G., Piel, J., Bierbaum, G., 2011. Expression of the lantibiotic mersacidin in Bacillus amyloliquefaciens FZB42. PLoS One 6, e22389.

Hobbs, J.K., Miller, K., O’Neill, A.J., Chopra, I., 2008. Consequences of daptomycin-mediated membrane damage in Staphylococcus aureus. J. Antimicrob. Chemother. 62, 1003–8.

Holo, H., Jeknic, Z., Daeschel, M., Stevanovic, S., Nes, I.F., 2001. Plantaricin W from Lactobacillus plantarum belongs to a new family of two-peptide lantibiotics. Microbiology 147, 643–51.

Hou, Y., Tianero, M.D.B., Kwan, J.C., Wyche, T.P., Michel, C.R., Ellis, G.A., Vazquez-Rivera, E., Braun, D.R., Rose, W.E., Schmidt, E.W., Bugni, T.S., 2012. Structure and biosynthesis of the antibiotic bottromycin D. Org. Lett. 14, 5050–3.

Hsu, T., Hingley-Wilson, S.M., Chen, B., Chen, M., Dai, A.Z., Morin, P.M., Marks, C.B., Padiyar, J., Goulding, C., Gingery, M., Eisenberg, D., Russell, R.G., Derrick, S.C., Collins, F.M., Morris, S.L., King, C.H., Jacobs, W.R., 2003. The primary mechanism of attenuation of bacillus Calmette-Guerin is a loss of secreted lytic function required for invasion of lung interstitial tissue. Proc. Natl. Acad. Sci. U. S. A. 100, 12420–5.

Page 220: Discovery of antimicrobial peptides active against

References

220 | P a g e

Hsueh, P.-R., 2012. Study for Monitoring Antimicrobial Resistance Trends (SMART) in the Asia-Pacific region, 2002-2010. Int. J. Antimicrob. Agents 40 Suppl, S1–3.

Huo, L., Rachid, S., Stadler, M., Wenzel, S.C., Müller, R., 2012. Synthetic biotechnology to study and engineer ribosomal bottromycin biosynthesis. Chem. Biol. 19, 1278–87.

Huppert, L. a., Ramsdell, T.L., Chase, M.R., Sarracino, D. a., Fortune, S.M., Burton, B.M., 2014. The ESX system in Bacillus subtilis mediates protein secretion. PLoS One 9, e96267.

Ibrahim, A., Yang, L., Johnston, C., Liu, X., Ma, B., Magarvey, N.A., 2012. Dereplicating nonribosomal peptides using an informatic search algorithm for natural products (iSNAP) discovery. Proc. Natl. Acad. Sci. U. S. A. 109, 19196–201.

Infectious Diseases Society of America, 2010. The 10 x ’20 Initiative: pursuing a global commitment to develop 10 new antibacterial drugs by 2020. Clin. Infect. Dis. 50, 1081–3.

Ingham, A.B., Sproat, K.W., Tizard, M.L. V, Moore, R.J., 2005. A versatile system for the expression of nonmodified bacteriocins in Escherichia coli. J. Appl. Microbiol. 98, 676–83.

Inglin, R.C., Stevens, M.J.A., Meile, L., Lacroix, C., Meile, L., 2015. High-throughput screening assays for antibacterial and antifungal activities of Lactobacillus species. J. Microbiol. Methods 114, 26–9.

Iwatani, S., Zendo, T., Yoneyama, F., Nakayama, J., Sonomoto, K., 2007. Characterization and structure analysis of a novel bacteriocin, lacticin Z, produced by Lactococcus lactis QU 14. Biosci. Biotechnol. Biochem. 71, 1984–92.

Iwatsuki, M., Uchida, R., Takakusagi, Y., Matsumoto, A., Jiang, C.-L., Takahashi, Y., Arai, M., Kobayashi, S., Matsumoto, M., Inokoshi, J., Tomoda, H., Omura, S., 2007. Lariatins, novel anti-mycobacterial peptides with a lasso structure, produced by Rhodococcus jostii K01-B0171. J. Antibiot. (Tokyo). 60, 357–63.

Iyer, L.M., Burroughs, A.M., Aravind, L., 2006. The prokaryotic antecedents of the ubiquitin-signaling system and the early evolution of ubiquitin-like beta-grasp domains. Genome Biol. 7, R60.

Izquierdo, E., Bednarczyk, A., Schaeffer, C., Cai, Y., Marchioni, E., Van Dorsselaer, A., Ennahar, S., 2008. Production of enterocins L50A, L50B, and IT, a new enterocin, by Enterococcus faecium IT62, a strain isolated from Italian ryegrass in Japan. Antimicrob. Agents Chemother. 52, 1917–23.

Page 221: Discovery of antimicrobial peptides active against

References

221 | P a g e

Jack, R.W., Tagg, J.R., Ray, B., 1995. Bacteriocins of gram-positive bacteria. Microbiol. Rev. 59, 171–200.

Jacoby, G.A., 2005. Mechanisms of resistance to quinolones. Clin. Infect. Dis. 41, S120–6.

Jean, S.-S., Hsueh, P.-R., 2011. High burden of antimicrobial resistance in Asia. Int. J. Antimicrob. Agents 37, 291–5.

Joerger, M.C., Klaenhammer, T.R., 1986. Characterization and purification of helveticin J and evidence for a chromosomally determined bacteriocin produced by Lactobacillus helveticus 481. J. Bacteriol. 167, 439–46.

Johnson, L.M., Gellman, S.H., 2013. α-Helix mimicry with α/β-peptides. Methods Enzymol. 523, 407–29.

José, M., Burgos, G., Pulido, R.P., López, C., Grande Burgos, M.J., Pulido, R.P., Del Carmen López Aguayo, M., Gálvez, A., Lucas, R., 2014. The Cyclic Antibacterial Peptide Enterocin AS-48: Isolation, Mode of Action, and Possible Food Applications. Int. J. Mol. Sci. 15, 22706–27.

Joshi, N.A., Fass, J.N., 2011. Sickle: A sliding-window, adaptive, quality-based trimming tool for FastQ files [Software].

Just-Baringo, X., Albericio, F., Álvarez, M., 2014. Thiopeptide antibiotics: retrospective and recent advances. Mar. Drugs 12, 317–51.

Kaiser, A.L., Montville, T.J., 1996. Purification of the bacteriocin bavaricin MN and characterization of its mode of action against Listeria monocytogenes Scott A cells and lipid vesicles. Appl. Environ. Microbiol. 62, 4529–35.

Kalmokoff, M.L., Cyr, T.D., Hefford, M.A., Whitford, M.F., Teather, R.M., 2003. Butyrivibriocin AR10, a new cyclic bacteriocin produced by the ruminal anaerobe Butyrivibrio fibrisolvens AR10: characterization of the gene and peptide. Can. J. Microbiol. 49, 763–73.

Kang, H.-S., Brady, S.F., 2013. Arimetamycin A: improving clinically relevant families of natural products through sequence-guided screening of soil metagenomes. Angew. Chem. Int. Ed. Engl. 52, 11063–7.

Karakas Sen, A., Narbad, A., Horn, N., Dodd, H.M., Parr, A.J., Colquhoun, I., Gasson, M.J., 1999. Post-translational modification of nisin. The involvement of NisB in the dehydration process. Eur. J. Biochem. 261, 524–32.

Katz, M.L., Mueller, L. V, Polyakov, M., Weinstock, S.F., 2006. Where have all the antibiotic patents gone? Nat. Biotechnol. 24, 1529–31.

Kawai, Y., Saito, T., Kitazawa, H., Itoh, T., 1998. Gassericin A; an uncommon cyclic bacteriocin produced by Lactobacillus gasseri LA39 linked at N- and C-

Page 222: Discovery of antimicrobial peptides active against

References

222 | P a g e

terminal ends. Biosci. Biotechnol. Biochem. 62, 2438–40.

Kawulka, K., Sprules, T., McKay, R.T., Mercier, P., Diaper, C.M., Zuber, P., Vederas, J.C., 2003. Structure of subtilosin A, an antimicrobial peptide from Bacillus subtilis with unusual posttranslational modifications linking cysteine sulfurs to alpha-carbons of phenylalanine and threonine. J. Am. Chem. Soc. 125, 4726–7.

Kemperman, R., Jonker, M., Nauta, A., Kuipers, O.P., Kok, J., 2003a. Functional Analysis of the Gene Cluster Involved in Production of the Bacteriocin Circularin A by Clostridium beijerinckii ATCC 25752. Appl. Environ. Microbiol. 69, 5839–48.

Kemperman, R., Kuipers, A., Karsens, H., Nauta, A., Kuipers, O., Kok, J., 2003b. Identification and characterization of two novel clostridial bacteriocins, circularin A and closticin 574. Appl. Environ. Microbiol. 69, 1589–97.

Kjos, M., Borrero, J., Opsata, M., Birri, D.J., Holo, H., Cintas, L.M., Snipen, L., Hernández, P.E., Nes, I.F., Diep, D.B., 2011. Target recognition, resistance, immunity and genome mining of class II bacteriocins from Gram-positive bacteria. Microbiology 157, 3256–67.

Klaenhammer, T.R., 1993. Genetics of bacteriocins produced by lactic acid bacteria. FEMS Microbiol. Rev. 12, 39–85.

Kluskens, L.D., Kuipers, A., Rink, R., De Boef, E., Fekken, S., Driessen, a. J.M., Kuipers, O.P., Moll, G.N., 2005. Post-translational modification of therapeutic peptides by NisB, the dehydratase of the lantibiotic nisin. Biochemistry 44, 12827–34.

Knerr, P.J., van der Donk, W.A., 2012. Discovery, Biosynthesis, and Engineering of Lantipeptides. Annu. Rev. Biochem. 81, 479–505.

Kneuper, H., Cao, Z.P., Twomey, K.B., Zoltner, M., Jäger, F., Cargill, J.S., Chalmers, J., van der Kooi-Pol, M.M., van Dijl, J.M., Ryan, R.P., Hunter, W.N., Palmer, T., 2014. Heterogeneity in ess transcriptional organization and variable contribution of the Ess/Type VII protein secretion system to virulence across closely related Staphylocccus aureus strains. Mol. Microbiol. 93, 928–43.

Kobayashi, Y., Ichioka, M., Hirose, T., Nagai, K., Matsumoto, A., Matsui, H., Hanaki, H., Masuma, R., Takahashi, Y., Omura, S., Sunazuka, T., 2010. Bottromycin derivatives: efficient chemical modifications of the ester moiety and evaluation of anti-MRSA and anti-VRE activities. Bioorg. Med. Chem. Lett. 20, 6116–20.

Koch, G., Yepes, A., Förstner, K.U., Wermser, C., Stengel, S.T., Modamio, J.,

Page 223: Discovery of antimicrobial peptides active against

References

223 | P a g e

Ohlsen, K., Foster, K.R., Lopez, D., 2014. Evolution of Resistance to a Last-Resort Antibiotic in Staphylococcus aureus via Bacterial Competition. Cell 158, 1060–71.

Kodani, S., Hudson, M.E., Durrant, M.C., Buttner, M.J., Nodwell, J.R., Willey, J.M., 2004. The SapB morphogen is a lantibiotic-like peptide derived from the product of the developmental gene ramS in Streptomyces coelicolor. Proc. Natl. Acad. Sci. U. S. A. 101, 11448–53.

Kodani, S., Lodato, M.A., Durrant, M.C., Picart, F., Willey, J.M., 2005. SapT, a lanthionine-containing peptide involved in aerial hyphae formation in the streptomycetes. Mol. Microbiol. 58, 1368–80.

Koponen, O., Tolonen, M., Qiao, M., Wahlstrom, G., Helin, J., Saris, P.E.J., 2002. NisB is required for the dehydration and NisC for the lanthionine formation in the post-translational modification of nisin. Microbiology 148, 3561–8.

Krawczyk, B., Völler, G.H., Völler, J., Ensle, P., Süssmuth, R.D., 2012. Curvopeptin: a new lanthionine-containing class III lantibiotic and its co-substrate promiscuous synthetase. Chembiochem 13, 2065–71.

Kumarasamy, K.K., Toleman, M.A., Walsh, T.R., Bagaria, J., Butt, F., Balakrishnan, R., Chaudhary, U., Doumith, M., Giske, C.G., Irfan, S., Krishnan, P., Kumar, A. V, Maharjan, S., Mushtaq, S., Noorie, T., Paterson, D.L., Pearson, A., Perry, C., Pike, R., Rao, B., Ray, U., Sarma, J.B., Sharma, M., Sheridan, E., Thirunarayan, M.A., Turton, J., Upadhyay, S., Warner, M., Welfare, W., Livermore, D.M., Woodford, N., 2010. Emergence of a new antibiotic resistance mechanism in India, Pakistan, and the UK: a molecular, biological, and epidemiological study. Lancet Infect. Dis. 10, 597–602.

Kuo, Y.-C., Liu, C.-F., Lin, J.-F., Li, A.-C., Lo, T.-C., Lin, T.-H., 2013. Characterization of putative class II bacteriocins identified from a non-bacteriocin-producing strain Lactobacillus casei ATCC 334. Appl. Microbiol. Biotechnol. 97, 237–46.

Lagos, R., Baeza, M., Corsini, G., Hetz, C., Strahsburger, E., Castillo, J.A., Vergara, C., Monasterio, O., 2001. Structure, organization and characterization of the gene cluster involved in the production of microcin E492, a channel-forming bacteriocin. Mol. Microbiol. 42, 229–43.

Lagos, R., Tello, M., Mercado, G., García, V., Monasterio, O., 2009. Antibacterial and antitumorigenic properties of microcin E492, a pore-forming bacteriocin. Curr. Pharm. Biotechnol. 10, 74–85.

Lam, K.S., 2007. New aspects of natural products in drug discovery. Trends Microbiol. 15, 279–89.

Page 224: Discovery of antimicrobial peptides active against

References

224 | P a g e

Larkin, M.A., Blackshields, G., Brown, N.P., Chenna, R., McGettigan, P.A., McWilliam, H., Valentin, F., Wallace, I.M., Wilm, A., Lopez, R., Thompson, J.D., Gibson, T.J., Higgins, D.G., 2007. Clustal W and Clustal X version 2.0. Bioinformatics 23, 2947–8.

Lashuel, H.A., Hartley, D., Petre, B.M., Walz, T., Lansbury, P.T., 2002. Neurodegenerative disease: amyloid pores from pathogenic mutations. Nature 418, 291.

Lautru, S., Deeth, R.J., Bailey, L.M., Challis, G.L., 2005. Discovery of a new peptide natural product by Streptomyces coelicolor genome mining. Nat. Chem. Biol. 1, 265–9.

Lawton, E.M., Cotter, P.D., Hill, C., Ross, R.P., 2007. Identification of a novel two-peptide lantibiotic, haloduracin, produced by the alkaliphile Bacillus halodurans C-125. FEMS Microbiol. Lett. 267, 64–71.

Laxminarayan, R., Duse, A., Wattal, C., Zaidi, A.K.M., Wertheim, H.F.L., Sumpradit, N., Vlieghe, E., Hara, G.L., Gould, I.M., Goossens, H., Greko, C., So, A.D., Bigdeli, M., Tomson, G., Woodhouse, W., Ombaka, E., Peralta, A.Q., Qamar, F.N., Mir, F., Kariuki, S., Bhutta, Z.A., Coates, A., Bergstrom, R., Wright, G.D., Brown, E.D., Cars, O., 2013. Antibiotic resistance-the need for global solutions. Lancet Infect. Dis. 13, 1057–98.

Leach, K.L., Brickner, S.J., Noe, M.C., Miller, P.F., 2011. Linezolid, the first oxazolidinone antibacterial agent. Ann. N. Y. Acad. Sci. 1222, 49–54.

Lee, G.C., Reveles, K.R., Attridge, R.T., Lawson, K.A., Mansi, I.A., Lewis, J.S., Frei, C.R., 2014. Outpatient antibiotic prescribing in the United States: 2000 to 2010. BMC Med. 12, 96.

Lee, J.W., Park, S.C., Kim, M.K., Cheon, M.W., Kim, G.Y., Cho, J.-H., 2014. Prevalence of antimicrobial resistance in normal rectal flora of patients undergoing transrectal ultrasonography-guided prostate biopsy in Korea. Int. J. Urol. 21, 811–4.

Lee, S.W., Mitchell, D.A., Markley, A.L., Hensler, M.E., Gonzalez, D., Wohlrab, A., Dorrestein, P.C., Nizet, V., Dixon, J.E., 2008. Discovery of a widely distributed toxin biosynthetic gene cluster. Proc. Natl. Acad. Sci. U. S. A. 105, 5879–84.

Leeds, J. a, Schmitt, E.K., Krastel, P., 2006. Recent developments in antibacterial drug discovery: microbe-derived natural products--from collection to the clinic. Expert Opin. Investig. Drugs 15, 211–26.

Letzel, A.-C., Pidot, S.J., Hertweck, C., 2014. Genome mining for ribosomally synthesized and post-translationally modified peptides (RiPPs) in anaerobic bacteria. BMC Genomics 15, 983.

Page 225: Discovery of antimicrobial peptides active against

References

225 | P a g e

Lewis, K., 2013. Platforms for antibiotic discovery. Nat. Rev. Drug Discov. 12, 371–87.

Li, B., Yu, J.P.J., Brunzelle, J.S., Moll, G.N., van der Donk, W.A., Nair, S.K., 2006. Structure and mechanism of the lantibiotic cyclase involved in nisin biosynthesis. Science 311, 1464–7.

Li, J., Gu, L., Aach, J., Church, G.M., 2014. Improved cell-free RNA and protein synthesis system. PLoS One 9, e106232.

Li, W., Tailhades, J., O’Brien-Simpson, N.M., Separovic, F., Otvos, L., Hossain, M.A., Wade, J.D., 2014. Proline-rich antimicrobial peptides: potential therapeutics against antibiotic-resistant bacteria. Amino Acids 46, 2287–94.

Li, Y.M., Milne, J.C., Madison, L.L., Kolter, R., Walsh, C.T., 1996. From peptide precursors to oxazole and thiazole-containing peptide antibiotics: microcin B17 synthase. Science 274, 1188–93.

Lincke, T., Behnken, S., Ishida, K., Roth, M., Hertweck, C., 2010. Closthioamide: an unprecedented polythioamide antibiotic from the strictly anaerobic bacterium Clostridium cellulolyticum. Angew. Chem. Int. Ed. Engl. 49, 2011–3.

Ling, L.L., Schneider, T., Peoples, A.J., Spoering, A.L., Engels, I., Conlon, B.P., Mueller, A., Schäberle, T.F., Hughes, D.E., Epstein, S., Jones, M., Lazarides, L., Steadman, V.A., Cohen, D.R., Felix, C.R., Fetterman, K.A., Millett, W.P., Nitti, A.G., Zullo, A.M., Chen, C., Lewis, K., 2015. A new antibiotic kills pathogens without detectable resistance. Nature 517, 455–9.

Lipinski, C.A., Lombardo, F., Dominy, B.W., Feeney, P.J., 2001. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings1PII of original article: S0169-409X(96)00423-1. The article was originally published in Advanced Drug Delivery Reviews 23 (1997) 3. Adv. Drug Deliv. Rev. 46, 3–26.

Liu, S.Y., Lin, J.Y., Chu, C., Su, L.H., Lin, T.Y., Chiu, C.H., 2006. Integron-associated imipenem resistance in Acinetobacter baumannii isolated from a regional hospital in Taiwan. Int. J. Antimicrob. Agents 27, 81–4.

Livermore, D.M., 2007. Introduction: the challenge of multiresistance. Int. J. Antimicrob. Agents 29 Suppl 3, S1–7.

Livermore, D.M., 2011. Discovery research: the scientific challenge of finding new antibiotics. J. Antimicrob. Chemother. 66, 1941–4.

Livermore, D.M., Mushtaq, S., Warner, M., Zhang, J., Maharjan, S., Doumith, M., Woodford, N., 2011. Activities of NXL104 combinations with ceftazidime and aztreonam against carbapenemase-Producing Enterobacteriaceae. Antimicrob. Agents Chemother. 55, 390–4.

Page 226: Discovery of antimicrobial peptides active against

References

226 | P a g e

Lloubès, R., Goemaere, E., Zhang, X., Cascales, E., Duché, D., 2012. Energetics of colicin import revealed by genetic cross-complementation between the Tol and Ton systems. Biochem. Soc. Trans. 40, 1480–5.

Lohans, C.T., Vederas, J.C., 2013. Structural characterization of thioether-bridged bacteriocins. J. Antibiot. (Tokyo). 67, 23–30.

López-Rojas, R., Docobo-Pérez, F., Pachón-Ibáñez, M.E., de la Torre, B.G., Fernández-Reyes, M., March, C., Bengoechea, J. a, Andreu, D., Rivas, L., Pachón, J., 2011. Efficacy of cecropin A-melittin peptides on a sepsis model of infection by pan-resistant Acinetobacter baumannii. Eur. J. Clin. Microbiol. Infect. Dis. 30, 1391–8.

Lorenzo, V. De, 1984. Isolation and characterization of microcin E 492 from Klebsiella pneumoniae. Arch. Microbiol. 139, 72–5.

Louis, K.S., Siegel, A.C., 2011. Cell viability analysis using trypan blue: manual and automated methods. Methods Mol. Biol. 740, 7–12.

Lubelski, J., Rink, R., Khusainov, R., Moll, G.N., Kuipers, O.P., 2008. Biosynthesis, immunity, regulation, mode of action and engineering of the model lantibiotic nisin. Cell. Mol. Life Sci. 65, 455–76.

Mach, B., Reich, E., Tatum, E.L., 1963. Separation of the biosynthesis of the antibiotic polypeptide tyrocidine from protein biosynthesis. Proc. Natl. Acad. Sci. U. S. A. 50, 175–81.

Maksimov, M.O., Pan, S.J., James Link, A., 2012. Lasso peptides: structure, function, biosynthesis, and engineering. Nat. Prod. Rep. 29, 996–1006.

Mantovani, H.C., Russell, J.B., 2001. Nisin resistance of Streptococcus bovis. Appl. Environ. Microbiol. 67, 808–13.

Marahiel, M.A., Stachelhaus, T., Mootz, H.D., 1997. Modular Peptide Synthetases Involved in Nonribosomal Peptide Synthesis. Chem. Rev. 97, 2651–74.

Marchesi, J.R., Sato, T., Weightman, A.J., Martin, T. a., Fry, J.C., Hiom, S.J., Wade, W.G., 1998. Design and evaluation of useful bacterium-specific PCR primers that amplify genes coding for bacterial 16S rRNA. Appl. Environ. Microbiol. 64, 795–9.

Marchler-Bauer, A., Derbyshire, M.K., Gonzales, N.R., Lu, S., Chitsaz, F., Geer, L.Y., Geer, R.C., He, J., Gwadz, M., Hurwitz, D.I., Lanczycki, C.J., Lu, F., Marchler, G.H., Song, J.S., Thanki, N., Wang, Z., Yamashita, R.A., Zhang, D., Zheng, C., Bryant, S.H., 2014. CDD: NCBI’s conserved domain database. Nucleic Acids Res. 43, D222–6.

Marcille, F., Gomez, A., Joubert, P., Ladiré, M., Veau, G., Clara, A., Gavini, F.,

Page 227: Discovery of antimicrobial peptides active against

References

227 | P a g e

Willems, A., Fons, M., 2002. Distribution of genes encoding the trypsin-dependent lantibiotic ruminococcin A among bacteria isolated from human fecal microbiota. Appl. Environ. Microbiol. 68, 3424–31.

Marcoleta, A., Gutiérrez-Cortez, S., Maturana, D., Monasterio, O., Lagos, R., 2013a. Whole-Genome Sequence of the Microcin E492-Producing Strain Klebsiella pneumoniae RYC492. Genome Announc. 1.

Marcoleta, A., Marín, M., Mercado, G., Valpuesta, J.M., Monasterio, O., Lagos, R., 2013b. Microcin e492 amyloid formation is retarded by posttranslational modification. J. Bacteriol. 195, 3995–4004.

Martín, M., Gutiérrez, J., Criado, R., Herranz, C., Cintas, L.M., Hernández, P.E., 2007. Cloning, production and expression of the bacteriocin enterocin A produced by Enterococcus faecium PLBC21 in Lactococcus lactis. Appl. Microbiol. Biotechnol. 76, 667–75.

Martin, N.I., Sprules, T., Carpenter, M.R., Cotter, P.D., Hill, C., Ross, R.P., Vederas, J.C., 2004. Structural characterization of lacticin 3147, a two-peptide lantibiotic with synergistic activity. Biochemistry 43, 3049–56.

Martínez, B., Böttiger, T., Schneider, T., Rodríguez, A., Sahl, H.-G., Wiedemann, I., 2008. Specific interaction of the unmodified bacteriocin Lactococcin 972 with the cell wall precursor lipid II. Appl. Environ. Microbiol. 74, 4666–70.

Martínez, J.L., Baquero, F., 2014. Emergence and spread of antibiotic resistance: setting a parameter space. Ups. J. Med. Sci. 119, 68–77.

Martin-Visscher, L.A., Gong, X., Duszyk, M., Vederas, J.C., 2009. The three-dimensional structure of carnocyclin A reveals that many circular bacteriocins share a common structural motif. J. Biol. Chem. 284, 28674–81.

Masuda, Y., Ono, H., Kitagawa, H., Ito, H., Mu, F., Sawa, N., Zendo, T., Sonomoto, K., 2011. Identification and characterization of leucocyclicin Q, a novel cyclic bacteriocin produced by Leuconostoc mesenteroides TK41401. Appl. Environ. Microbiol. 77, 8164–70.

Mathavan, I., Beis, K., 2012. The role of bacterial membrane proteins in the internalization of microcin MccJ25 and MccB17. Biochem. Soc. Trans. 40, 1539–43.

Mathavan, I., Zirah, S., Mehmood, S., Choudhury, H.G., Goulard, C., Li, Y., Robinson, C. V, Rebuffat, S., Beis, K., 2014. Structural basis for hijacking siderophore receptors by antimicrobial lasso peptides. Nat. Chem. Biol. 10, 340–2.

Mathur, H., O’Connor, P.M., Cotter, P.D., Hill, C., Ross, R.P., 2014. Heterologous expression of thuricin CD immunity genes in Listeria monocytogenes.

Page 228: Discovery of antimicrobial peptides active against

References

228 | P a g e

Antimicrob. Agents Chemother. 58, 3421–8.

Matsumura, Y., Yamamoto, M., Nagao, M., Ito, Y., Takakura, S., Ichiyama, S., 2013. Association of fluoroquinolone resistance, virulence genes, and IncF plasmids with extended-spectrum-β-lactamase-producing Escherichia coli sequence type 131 (ST131) and ST405 clonal groups. Antimicrob. Agents Chemother. 57, 4736–42.

Mba Medie, F., Champion, M.M., Williams, E.A., Champion, P.A.D., 2014. Homeostasis of N-α-terminal acetylation of EsxA correlates with virulence in Mycobacterium marinum. Infect. Immun. 82, 4572–86.

McAuliffe, O., Ross, R.P., Hill, C., 2001. Lantibiotics: structure, biosynthesis and mode of action. FEMS Microbiol. Rev. 25, 285–308.

McAuliffe, O., Ryan, M., Ross, R., 1998. Lacticin 3147, a broad-spectrum bacteriocin which selectively dissipates the membrane potential. Appl. Environ. Microbiol. 64, 439–45.

McClerren, A.L., Cooper, L.E., Quan, C., Thomas, P.M., Kelleher, N.L., van der Donk, W.A., 2006. Discovery and in vitro biosynthesis of haloduracin, a two-component lantibiotic. Proc. Natl. Acad. Sci. U. S. A. 103, 17243–8.

McDaniel, R., Welch, M., Hutchinson, C.R., 2005. Genetic approaches to polyketide antibiotics. 1. Chem. Rev. 105, 543–58.

McElvania TeKippe, E., Thomas, B.S., Ewald, G.A., Lawrence, S.J., Burnham, C.-A.D., 2014. Rapid emergence of daptomycin resistance in clinical isolates of Corynebacterium striatum… a cautionary tale. Eur. J. Clin. Microbiol. Infect. Dis. 33, 2199–205.

McIntosh, J.A., Donia, M.S., Schmidt, E.W., 2009. Ribosomal peptide natural products: bridging the ribosomal and nonribosomal worlds. Nat. Prod. Rep. 26, 537–59.

McLaughlin, B., Chon, J.S., MacGurn, J. a., Carlsson, F., Cheng, T.L., Cox, J.S., Brown, E.J., 2007. A mycobacterium ESX-1-secreted virulence factor with unique requirements for export. PLoS Pathog. 3, 1051–61.

Medzihradszky, K.F., Chalkley, R.J., 2015. Lessons in de novo peptide sequencing by tandem mass spectrometry. Mass Spectrom. Rev. 34, 43–63.

Mercado, G., Tello, M., Marín, M., Monasterio, O., Lagos, R., 2008. The production in vivo of microcin E492 with antibacterial activity depends on salmochelin and EntF. J. Bacteriol. 190, 5464–71.

Meziane-Cherif, D., Stogios, P.J., Evdokimova, E., Savchenko, A., Courvalin, P., 2014. Structural basis for the evolution of vancomycin resistance D,D-

Page 229: Discovery of antimicrobial peptides active against

References

229 | P a g e

peptidases. Proc. Natl. Acad. Sci. U. S. A. 111, 5872–7.

Miescher, S., Stierli, M.P., Teuber, M., Meile, L., 2000. Propionicin SM1, a bacteriocin from Propionibacterium jensenii DF1: isolation and characterization of the protein and its gene. Syst. Appl. Microbiol. 23, 174–84.

Miller, S.M., Simon, R.J., Ng, S., Zuckermann, R.N., Kerr, J.M., Moos, W.H., 1994. Proteolytic studies of homologous peptide and N-substituted glycine peptoid oligomers. Bioorg. Med. Chem. Lett. 4, 2657–62.

Mills, S.D., 2006. When will the genomics investment pay off for antibacterial discovery? Biochem. Pharmacol. 71, 1096–102.

Mitchell, J.M., Clasman, J.R., June, C.M., Kaitany, K.-C.J., LaFleur, J.R., Taracila, M.A., Klinger, N. V, Bonomo, R.A., Wymore, T., Szarecka, A., Powers, R.A., Leonard, D.A., 2015. Structural basis of activity against aztreonam and extended spectrum cephalosporins for two carbapenem-hydrolyzing class D β-lactamases from Acinetobacter baumannii. Biochemistry 54, 1976–87.

Montero, C.I., Stock, F., Murray, P.R., 2008. Mechanisms of resistance to daptomycin in Enterococcus faecium. Antimicrob. Agents Chemother. 52, 1167–70.

Mora, L., de Zamaroczy, M., 2014. In vivo processing of DNase colicins E2 and E7 is required for their import into the cytoplasm of target cells. PLoS One 9, e96549.

Morrison, K.C., Hergenrother, P.J., 2014. Natural products as starting points for the synthesis of complex and diverse compounds. Nat. Prod. Rep. 31, 6–14.

Mu, F., Masuda, Y., Zendo, T., Ono, H., Kitagawa, H., Ito, H., Nakayama, J., Sonomoto, K., 2014. Biological function of a DUF95 superfamily protein involved in the biosynthesis of a circular bacteriocin, leucocyclicin Q. J. Biosci. Bioeng. 117, 158–64.

Müller, W.M., Schmiederer, T., Ensle, P., Süssmuth, R.D., 2010. In vitro biosynthesis of the prepeptide of type-III lantibiotic labyrinthopeptin A2 including formation of a C-C bond as a post-translational modification. Angew. Chem. Int. Ed. Engl. 49, 2436–40.

Naclerio, G., Ricca, E., Sacco, M., De Felice, M., 1993. Antimicrobial activity of a newly identified bacteriocin of Bacillus cereus. Appl. Environ. Microbiol. 59, 4313–6.

Newman, D.J., 2008. Natural products as leads to potential drugs: an old process or the new hope for drug discovery? J. Med. Chem. 51, 2589–99.

Ng, C.L., Lang, K., Meenan, N.A.G., Sharma, A., Kelley, A.C., Kleanthous, C.,

Page 230: Discovery of antimicrobial peptides active against

References

230 | P a g e

Ramakrishnan, V., 2010. Structural basis for 16S ribosomal RNA cleavage by the cytotoxic domain of colicin E3. Nat. Struct. Mol. Biol. 17, 1241–6.

Nichols, D., Cahoon, N., Trakhtenberg, E.M., Pham, L., Mehta, A., Belanger, A., Kanigan, T., Lewis, K., Epstein, S.S., 2010. Use of ichip for high-throughput in situ cultivation of “uncultivable” microbial species. Appl. Environ. Microbiol. 76, 2445–50.

Nielsen, K.F., Månsson, M., Rank, C., Frisvad, J.C., Larsen, T.O., 2011. Dereplication of microbial natural products by LC-DAD-TOFMS. J. Nat. Prod. 74, 2338–48.

Nishie, M., Nagao, J.-I., Sonomoto, K., 2012. Antibacterial peptides “bacteriocins”: an overview of their diverse characteristics and applications. Biocontrol Sci. 17, 1–16.

Nissen-Meyer, J., Rogne, P., Oppegård, C., Haugen, H.S., Kristiansen, P.E., 2009. Structure-function relationships of the non-lanthionine-containing peptide (class II) bacteriocins produced by gram-positive bacteria. Curr. Pharm. Biotechnol. 10, 19–37.

Nolan, E.M., Fischbach, M.A., Koglin, A., Walsh, C.T., 2007. Biosynthetic tailoring of microcin E492m: post-translational modification affords an antibacterial siderophore-peptide conjugate. J. Am. Chem. Soc. 129, 14336–47.

Nolan, E.M., Walsh, C.T., 2008. Investigations of the MceIJ-catalyzed posttranslational modification of the microcin E492 C-terminus: linkage of ribosomal and nonribosomal peptides to form “trojan horse” antibiotics. Biochemistry 47, 9289–99.

Nordmann, P., 2014. Carbapenemase-producing Enterobacteriaceae: overview of a major public health challenge. Médecine Mal. Infect. 44, 51–6.

O’Neill, J., 2014. Antimicrobial Resistance: Tackling a Crisis for the Future Health and Wealth of Nationsle, The Review on Antimicrobial Resistance.

O’Shea, E.F., O’Connor, P.M., Raftis, E.J., O’Toole, P.W., Stanton, C., Cotter, P.D., Ross, R.P., Hill, C., 2011. Production of multiple bacteriocins from a single locus by gastrointestinal strains of Lactobacillus salivarius. J. Bacteriol. 193, 6973–82.

Oberstein, T.J., Spitzer, P., Klafki, H.-W., Linning, P., Neff, F., Knölker, H.-J., Lewczuk, P., Wiltfang, J., Kornhuber, J., Maler, J.M., 2015. Astrocytes and microglia but not neurons preferentially generate N-terminally truncated Aβ peptides. Neurobiol. Dis. 73, 24–35.

Okkels, L.M., Müller, E.-C., Schmid, M., Rosenkrands, I., Kaufmann, S.H.E., Andersen, P., Jungblut, P.R., 2004. CFP10 discriminates between nonacetylated

Page 231: Discovery of antimicrobial peptides active against

References

231 | P a g e

and acetylated ESAT-6 of Mycobacterium tuberculosis by differential interaction. Proteomics 4, 2954–60.

Olejnik-Schmidt, A.K., Schmidt, M.T., Sip, A., Szablewski, T., Grajek, W., 2014. Expression of bacteriocin divercin AS7 in Escherichia coli and its functional analysis. Ann. Microbiol. 64, 1197–202.

Oman, T.J., Boettcher, J.M., Wang, H., Okalibe, X.N., van der Donk, W.A., 2011. Sublancin is not a lantibiotic but an S-linked glycopeptide. Nat. Chem. Biol. 7, 78–80.

Ooi, N., Miller, K., Hobbs, J., Rhys-Williams, W., Love, W., Chopra, I., 2009. XF-73, a novel antistaphylococcal membrane-active agent with rapid bactericidal activity. J. Antimicrob. Chemother. 64, 735–40.

Oppegard, C., Rogne, P., Emanuelsen, L., Kristiansen, P.E., Fimland, G., Nissen-Meyer, J., 2007. The two-peptide class II bacteriocins: Structure, production, and mode of action. J. Mol. Microbiol. Biotechnol. 13, 210–19.

Page, M.G.P., Heim, J., 2009. Prospects for the next anti-Pseudomonas drug. Curr. Opin. Pharmacol. 9, 558–65.

Pallen, M.J., 2002. The ESAT-6/WXG100 superfamily - And a new Gram-positive secretion system? Trends Microbiol. 10, 209–212.

Pandit, B., Gartel, A.L., 2011. Thiazole antibiotic thiostrepton synergize with bortezomib to induce apoptosis in cancer cells. PLoS One 6, e17110.

Parker, C.E., Mocanu, V., Mocanu, M., Dicheva, N., Warren, M.R., 2010. Mass Spectrometry for Post-Translational Modifications, in: Neuroproteomics. CRC Press, p. Chapter 6. Available from: http://www.ncbi.nlm.nih.

Pasteris, S.E., Vera Pingitore, E., Ale, C.E., Nader-Macías, M.E.F., 2014. Characterization of a bacteriocin produced by Lactococcus lactis subsp. lactis CRL 1584 isolated from a Lithobates catesbeianus hatchery. World J. Microbiol. Biotechnol. 30, 1053–62.

Patzer, S.I., Baquero, M.R., Bravo, D., Moreno, F., Hantke, K., 2003. The colicin G, H and X determinants encode microcins M and H47, which might utilize the catecholate siderophore receptors FepA, Cir, Fiu and IronN. Microbiology 149, 2557–70.

Payne, D.J., Gwynn, M.N., Holmes, D.J., Pompliano, D.L., 2007. Drugs for bad bugs: confronting the challenges of antibacterial discovery. Nat. Rev. Drug Discov. 6, 29–40.

Payne, D.J., Gwynn, M.N., Holmes, D.J., Rosenberg, M., 2004. Genomic approaches to antibacterial discovery. Methods Mol. Biol. 266, 231–59.

Page 232: Discovery of antimicrobial peptides active against

References

232 | P a g e

Perkins, D.N., Pappin, D.J., Creasy, D.M., Cottrell, J.S., 1999. Probability-based protein identification by searching sequence databases using mass spectrometry data. Electrophoresis 20, 3551–67.

Perry, J.D., Naqvi, S.H., Mirza, I.A., Alizai, S.A., Hussain, A., Ghirardi, S., Orenga, S., Wilkinson, K., Woodford, N., Zhang, J., Livermore, D.M., Abbasi, S.A., Raza, M.W., 2011. Prevalence of faecal carriage of Enterobacteriaceae with NDM-1 carbapenemase at military hospitals in Pakistan, and evaluation of two chromogenic media. J. Antimicrob. Chemother. 66, 2288–94.

Piard, J.C., Muriana, P.M., Desmazeaud, M.J., Klaenhammer, T.R., 1992. Purification and Partial Characterization of Lacticin 481, a Lanthionine-Containing Bacteriocin Produced by Lactococcus lactis subsp. lactis CNRZ 481. Appl. Environ. Microbiol. 58, 279–84.

Pickart, C.M., 2001. Mechanisms underlying ubiquitination. Annu. Rev. Biochem. 70, 503–33.

Pidot, S., Ishida, K., Cyrulies, M., Hertweck, C., 2014. Discovery of clostrubin, an exceptional polyphenolic polyketide antibiotic from a strictly anaerobic bacterium. Angew. Chem. Int. Ed. Engl. 53, 7856–9.

Pilsl, H., Braun, V., 1995. Strong function-related homology between the pore-forming colicins K and 5. J. Bacteriol. 177, 6973–7.

Poey, M.E., Azpiroz, M.F., Laviña, M., 2006. Comparative analysis of chromosome-encoded microcins. Antimicrob. Agents Chemother. 50, 1411–8.

Pons, A.-M., Lanneluc, I., Cottenceau, G., Sable, S., 2002. New developments in non-post translationally modified microcins. Biochimie 84, 531–7.

Potterat, O., Hamburger, M., 2013. Concepts and technologies for tracking bioactive compounds in natural product extracts: generation of libraries, and hyphenation of analytical processes with bioassays. Nat. Prod. Rep. 30, 546–64.

Poulsen, C., Panjikar, S., Holton, S.J., Wilmanns, M., Song, Y.-H.H., 2014. WXG100 protein superfamily consists of three subfamilies and exhibits an α-helical C-terminal conserved residue pattern. PLoS One 9, e89313.

Projan, S.J., 2003. Why is big Pharma getting out of antibacterial drug discovery? Curr. Opin. Microbiol. 6, 427–30.

Pucci, M., 2007. Novel Genetic Techniques and Approaches in the Microbial Genomics Era. Drugs R D 8, 201–12.

Pucci, M.J., Bush, K., 2013. Investigational antimicrobial agents of 2013. Clin. Microbiol. Rev. 26, 792–821.

Quail, M.A., Smith, M., Coupland, P., Otto, T.D., Harris, S.R., Connor, T.R.,

Page 233: Discovery of antimicrobial peptides active against

References

233 | P a g e

Bertoni, A., Swerdlow, H.P., Gu, Y., 2012. A tale of three next generation sequencing platforms: comparison of Ion Torrent, Pacific Biosciences and Illumina MiSeq sequencers. BMC Genomics 13, 341.

Ramnath, M., Beukes, M., Tamura, K., Hastings, J.W., 2000. Absence of a putative mannose-specific phosphotransferase system enzyme IIAB component in a leucocin a-resistant strain of Listeria monocytogenes, as shown by two-dimensional sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Appl. Environ. Microbiol. 66, 3098–101.

Ramsdell, T.L., Huppert, L.A., Sysoeva, T.A., Fortune, S.M., Burton, B.M., 2015. Linked domain architectures allow for specialization of function in the FtsK/SpoIIIE ATPases of ESX secretion systems. J. Mol. Biol. 427, 1119–32.

Rasheed, J.K., Kitchel, B., Zhu, W., Anderson, K.F., Clark, N.C., Ferraro, M.J., Savard, P., Humphries, R.M., Kallen, A.J., Limbago, B.M., 2013. New Delhi metallo-β-lactamase-producing Enterobacteriaceae, United States. Emerg. Infect. Dis. 19, 870–8.

Rea, M.C., Sit, C.S., Clayton, E., O’Connor, P.M., Whittal, R.M., Zheng, J., Vederas, J.C., Ross, R.P., Hill, C., 2010. Thuricin CD, a posttranslationally modified bacteriocin with a narrow spectrum of activity against Clostridium difficile. Proc. Natl. Acad. Sci. U. S. A. 107, 9352–7.

Rebuffat, S., 2012. Microcins in action: amazing defence strategies of Enterobacteria. Biochem. Soc. Trans. 40, 1456–62.

Repetto, G., del Peso, A., Zurita, J.L., 2008. Neutral red uptake assay for the estimation of cell viability/cytotoxicity. Nat. Protoc. 3, 1125–31.

Ribera, A., Fernández-Cuenca, F., Beceiro, A., Bou, G., Martínez-Martínez, L., Pascual, A., Cisneros, J.M., Rodríguez-Baño, J., Pachón, J., Vila, J., 2004. Antimicrobial susceptibility and mechanisms of resistance to quinolones and beta-lactams in Acinetobacter genospecies 3. Antimicrob. Agents Chemother. 48, 1430–2.

Rice, L.B., 2008. Federal funding for the study of antimicrobial resistance in nosocomial pathogens: no ESKAPE. J. Infect. Dis. 197, 1079–81.

Rigali, S., Titgemeyer, F., Barends, S., Mulder, S., Thomae, A.W., Hopwood, D.A., van Wezel, G.P., 2008. Feast or famine: the global regulator DasR links nutrient stress to antibiotic production by Streptomyces. EMBO Rep. 9, 670–5.

Robilotti, E., Deresinski, S., 2014. Carbapenemase-producing Klebsiella pneumoniae. F1000Prime Rep. 6, 80.

Rodríguez-Hernández, M.J., Saugar, J., Docobo-Pérez, F., de la Torre, B.G., Pachón-Ibáñez, M.E., García-Curiel, A., Fernández-Cuenca, F., Andreu, D., Rivas, L.,

Page 234: Discovery of antimicrobial peptides active against

References

234 | P a g e

Pachón, J., 2006. Studies on the antimicrobial activity of cecropin A-melittin hybrid peptides in colistin-resistant clinical isolates of Acinetobacter baumannii. J. Antimicrob. Chemother. 58, 95–100.

Rogers, L.A., 1928. The inhibiting effect of Streptococcus lactis on Lactobacillus bulgaricus. J. Bacteriol. 16, 321–5.

Rosén, J., Gottfries, J., Muresan, S., Backlund, A., Oprea, T.I., 2009. Novel chemical space exploration via natural products. J. Med. Chem. 52, 1953–62.

Rotem, S., Mor, A., 2009. Antimicrobial peptide mimics for improved therapeutic properties. Biochim. Biophys. Acta 1788, 1582–92.

Rouse, S., Sun, F., Vaughan, A., Sinderen, D., 2007. High-Throughput Isolation of Bacteriocin-Producing Lactic Acid Bacteria, with Potential Application in the Brewing Industry. J. Inst. Brew. 113, 256–62.

Rutherford, K., Parkhill, J., Crook, J., Horsnell, T., Rice, P., Rajandream, M.-A., Barrell, B., 2000. Artemis: sequence visualization and annotation. Bioinformatics 16, 944–5.

Ryan, M.P., Jack, R.W., Josten, M., Sahl, H.G., Jung, G., Ross, R.P., Hill, C., 1999. Extensive post-translational modification, including serine to d-alanine conversion, in the two-component lantibiotic, lacticin 3147. J. Biol. Chem. 274, 37544–50.

Saha, M., Sarkar, S., Sarkar, B., Sharma, B.K., Bhattacharjee, S., Tribedi, P., 2015. Microbial siderophores and their potential applications: a review. Environ. Sci. Pollut. Res. Int. [epub].

Salomón, R.A., Farías, R.N., 1992. Microcin 25, a novel antimicrobial peptide produced by Escherichia coli. J. Bacteriol. 174, 7428–35.

Salvucci, E., Saavedra, L., Sesma, F., 2007. Short peptides derived from the NH2-terminus of subclass IIa bacteriocin enterocin CRL35 show antimicrobial activity. J. Antimicrob. Chemother. 59, 1102–8.

Sánchez, S., Chávez, A., Forero, A., García-Huante, Y., Romero, A., Sánchez, M., Rocha, D., Sánchez, B., Avalos, M., Guzmán-Trampe, S., Rodríguez-Sanoja, R., Langley, E., Ruiz, B., 2010. Carbon source regulation of antibiotic production. J. Antibiot. (Tokyo). 63, 442–59.

Sánchez-Hidalgo, M., Montalbán-López, M., Cebrián, R., Valdivia, E., Martínez-Bueno, M., Maqueda, M., 2011. AS-48 bacteriocin: close to perfection. Cell. Mol. Life Sci. 68, 2845–57.

Sandiford, S., Upton, M., 2012. Identification, characterization, and recombinant expression of epidermicin NI01, a novel unmodified bacteriocin produced by

Page 235: Discovery of antimicrobial peptides active against

References

235 | P a g e

Staphylococcus epidermidis that displays potent activity against staphylococci. Antimicrob. Agents Chemother. 56, 1539–47.

Sandiford, S.K., 2014. Advances in the arsenal of tools available enabling the discovery of novel lantibiotics with therapeutic potential. Expert Opin. Drug Discov. 9, 283–97.

São-José, C., Baptista, C., Santos, M.A., 2004. Bacillus subtilis operon encoding a membrane receptor for bacteriophage SPP1. J. Bacteriol. 186, 8337–46.

São-José, C., Lhuillier, S., Lurz, R., Melki, R., Lepault, J., Santos, M.A., Tavares, P., 2006. The ectodomain of the viral receptor YueB forms a fiber that triggers ejection of bacteriophage SPP1 DNA. J. Biol. Chem. 281, 11464–70.

Saruwatari, T., Praseuth, A.P., Sato, M., Torikai, K., Noguchi, H., Watanabe, K., 2011. A comprehensive overview on genomically directed assembly of aromatic polyketides and macrolide lactones using fungal megasynthases. J. Antibiot. (Tokyo). 64, 9–17.

Sawa, N., Wilaipun, P., Kinoshita, S., Zendo, T., Leelawatcharamas, V., Nakayama, J., Sonomoto, K., 2012. Isolation and characterization of enterocin W, a novel two-peptide lantibiotic produced by Enterococcus faecalis NKR-4-1. Appl. Environ. Microbiol. 78, 900–3.

Sawa, N., Zendo, T., Kiyofuji, J., Fujita, K., Himeno, K., Nakayama, J., Sonomoto, K., 2009. Identification and characterization of lactocyclicin Q, a novel cyclic bacteriocin produced by Lactococcus sp. strain QU 12. Appl. Environ. Microbiol. 75, 1552–8.

Scherlach, K., Hertweck, C., 2009. Triggering cryptic natural product biosynthesis in microorganisms. Org. Biomol. Chem. 7, 1753–60.

Schmidt, E.W., Nelson, J.T., Rasko, D.A., Sudek, S., Eisen, J.A., Haygood, M.G., Ravel, J., 2005. Patellamide A and C biosynthesis by a microcin-like pathway in Prochloron didemni, the cyanobacterial symbiont of Lissoclinum patella. Proc. Natl. Acad. Sci. U. S. A. 102, 7315–20.

Schmitz, J.E., Ossiprandi, M.C., Rumah, K.R., Fischetti, V. a, 2011. Lytic enzyme discovery through multigenomic sequence analysis in Clostridium perfringens. Appl. Microbiol. Biotechnol. 89, 1783–95.

Scholz, R., Molohon, K.J., Nachtigall, J., Vater, J., Markley, A.L., Süssmuth, R.D., Mitchell, D.A., Borriss, R., 2011. Plantazolicin, a novel microcin B17/streptolysin S-like natural product from Bacillus amyloliquefaciens FZB42. J. Bacteriol. 193, 215–24.

Schwarzer, D., Finking, R., Marahiel, M.A., 2003. Nonribosomal peptides: from genes to products. Nat. Prod. Rep. 20, 275.

Page 236: Discovery of antimicrobial peptides active against

References

236 | P a g e

Seemann, T., 2014. Prokka: rapid prokaryotic genome annotation. Bioinformatics 30, 2068–9.

Shallcross, L.J., Davies, D.S.C., 2014. Antibiotic overuse: a key driver of antimicrobial resistance. Br. J. Gen. Pract. 64, 604–5.

Shenkarev, Z.O., Finkina, E.I., Nurmukhamedova, E.K., Balandin, S. V., Mineev, K.S., Nadezhdin, K.D., Yakimenko, Z. a., Tagaev, A. a., Temirov, Y. V., Arseniev, A.S., Ovchinnikova, T. V., 2010. Isolation, structure elucidation, and synergistic antibacterial activity of a novel two-component lantibiotic lichenicidin from Bacillus licheniformis VK21. Biochemistry 49, 6462–72.

Shimamura, H., Gouda, H., Nagai, K., Hirose, T., Ichioka, M., Furuya, Y., Kobayashi, Y., Hirono, S., Sunazuka, T., Omura, S., 2009. Structure determination and total synthesis of bottromycin A2: a potent antibiotic against MRSA and VRE. Angew. Chem. Int. Ed. Engl. 48, 914–7.

Shimizu, Y., Kanamori, T., Ueda, T., 2005. Protein synthesis by pure translation systems. Methods 36, 299–304.

Sieber, S., Marahiel, M., 2003. Learning from nature’s drug factories: nonribosomal synthesis of macrocyclic peptides. J. Bacteriol. 185, 7036–43.

Siezen, R.J., Kuipers, O.P., de Vos, W.M., 1996. Comparison of lantibiotic gene clusters and encoded proteins. Antonie Van Leeuwenhoek 69, 171–84.

Silver, L.L., 2011. Challenges of antibacterial discovery. Clin. Microbiol. Rev. 24, 71–109.

Simeone, R., Bobard, A., Lippmann, J., Bitter, W., Majlessi, L., Brosch, R., Enninga, J., 2012. Phagosomal rupture by Mycobacterium tuberculosis results in toxicity and host cell death. PLoS Pathog. 8, e1002507.

Simone, M., Monciardini, P., Gaspari, E., Donadio, S., Maffioli, S.I., 2013. Isolation and characterization of NAI-802, a new lantibiotic produced by two different Actinoplanes strains. J. Antibiot. (Tokyo). 66, 73–8.

Singh, M., Sareen, D., 2014. Novel LanT associated lantibiotic clusters identified by genome database mining. PLoS One 9, e91352.

Sit, C.S., van Belkum, M.J., McKay, R.T., Worobo, R.W., Vederas, J.C., 2011. The 3D solution structure of thurincin H, a bacteriocin with four sulfur to α-carbon crosslinks. Angew. Chem. Int. Ed. Engl. 50, 8718–21.

Smith, P.K., Krohn, R.I., Hermanson, G.T., Mallia, A.K., Gartner, F.H., Provenzano, M.D., Fujimoto, E.K., Goeke, N.M., Olson, B.J., Klenk, D.C., 1985. Measurement of protein using bicinchoninic acid. Anal. Biochem. 150, 76–85.

Snyder, A.B., Worobo, R.W., 2014. Chemical and genetic characterization of

Page 237: Discovery of antimicrobial peptides active against

References

237 | P a g e

bacteriocins: antimicrobial peptides for food safety. J. Sci. Food Agric. 94, 28–44.

Söding, J., Biegert, A., Lupas, A.N., 2005. The HHpred interactive server for protein homology detection and structure prediction. Nucleic Acids Res. 33, W244–8.

Sørensen, A.L., Nagai, S., Houen, G., Andersen, P., Andersen, A.B., 1995. Purification and characterization of a low-molecular-mass T-cell antigen secreted by Mycobacterium tuberculosis. Infect. Immun. 63, 1710–7.

Spellberg, B., Blaser, M., Guidos, R.J., Boucher, H.W., Bradley, J.S., Eisenstein, B.I., Gerding, D., Lynfield, R., Reller, L.B., Rex, J., Schwartz, D., Septimus, E., Tenover, F.C., Gilbert, D.N., 2011. Combating antimicrobial resistance: policy recommendations to save lives. Clin. Infect. Dis. 52 Suppl 5, S397–428.

Standing, K., 2003. Peptide and protein de novo sequencing by mass spectrometry. Curr. Opin. Struct. Biol. 13, 595–601.

Steen, H., Mann, M., 2004. The ABC’s (and XYZ's) of peptide sequencing. Nat. Rev. Mol. Cell Biol. 5, 699–711.

Stein, T., 2005. Bacillus subtilis antibiotics: structures, syntheses and specific functions. Mol. Microbiol. 56, 845–57.

Stepper, J., Shastri, S., Loo, T.S., Preston, J.C., Novak, P., Man, P., Moore, C.H., Havlíček, V., Patchett, M.L., Norris, G.E., 2011. Cysteine S-glycosylation, a new post-translational modification found in glycopeptide bacteriocins. FEBS Lett. 585, 645–50.

Stiege, W., Erdmann, V.A., 1995. The potentials of the in vitro protein biosynthesis system. J. Biotechnol. 41, 81–90.

Strahsburger, E., Baeza, M., Monasterio, O., Lagos, R., 2005. Cooperative uptake of microcin E492 by receptors FepA, Fiu, and Cir and inhibition by the siderophore enterochelin and its dimeric and trimeric hydrolysis products. Antimicrob. Agents Chemother. 49, 3083–6.

Strasser de Saad, a M., Manca de Nadra, M.C., 1993. Characterization of bacteriocin produced by Pediococcus pentosaceus from wine. J. Appl. Bacteriol. 74, 406–10.

Su, H.N., Chen, Z.H., Song, X.Y., Chen, X.L., Shi, M., Zhou, B.C., Zhao, X., Zhang, Y.Z., 2012. Antimicrobial Peptide Trichokonin VI-Induced Alterations in the Morphological and Nanomechanical Properties of Bacillus subtilis. PLoS One 7, 1–10.

SU, T.L., 1948. Micrococcin, an antibacterial substance formed by a strain of Micrococcus. Br. J. Exp. Pathol. 29, 473–81.

Page 238: Discovery of antimicrobial peptides active against

References

238 | P a g e

Sundaramoorthy, R., Fyfe, P.K., Hunter, W.N., 2008. Structure of Staphylococcus aureus EsxA suggests a contribution to virulence by action as a transport chaperone and/or adaptor Protein. J. Mol. Biol. 383, 603–14.

Swe, P.M., Cook, G.M., Tagg, J.R., Jack, R.W., 2009. Mode of action of dysgalacticin: a large heat-labile bacteriocin. J. Antimicrob. Chemother. 63, 679–86.

Sysoeva, T.A., Zepeda-Rivera, M.A., Huppert, L.A., Burton, B.M., 2014. Dimer recognition and secretion by the ESX secretion system in Bacillus subtilis. Proc. Natl. Acad. Sci. U. S. A. 111, 7653–8.

Szychowski, J., Truchon, J.-F., Bennani, Y.L., 2014. Natural products in medicine: transformational outcome of synthetic chemistry. J. Med. Chem. 57, 9292–308.

Tagg, J.R., Bannister, L. V, 1979. “Fingerprinting” beta-haemolytic streptococci by their production of and sensitivity to bacteriocine-like inhibitors. J. Med. Microbiol. 12, 397–411.

Tambadou, F., Lanneluc, I., Sablé, S., Klein, G.L., Doghri, I., Sopéna, V., Didelot, S., Barthélémy, C., Thiéry, V., Chevrot, R., 2014. Novel nonribosomal peptide synthetase (NRPS) genes sequenced from intertidal mudflat bacteria. FEMS Microbiol. Lett. 357, 123–30.

Tan, S.W.B., Chai, C.L.L., Moloney, M.G., Thompson, A.L., 2015. Synthesis of Mimics of Pramanicin from Pyroglutamic acid, and their Antibacterial Activity. J. Org. Chem. 80, 2661–75.

Tanaka, N., Sashikata, K., Yamaguchi, H., Umezawa, H., 1966. Inhibition of protein synthesis by bottromycin A2 and its hydrazide. J. Biochem. 60, 405–10.

Taylor, J.A., Johnson, R.S., 1997. Sequence database searches via de novo peptide sequencing by tandem mass spectrometry. Rapid Commun. Mass Spectrom. 11, 1067–75.

Taylor, P.L., Wright, G.D., 2008. Novel approaches to discovery of antibacterial agents. Anim. Health Res. Rev. 9, 237–46.

Telang, N. V, Satpute, M.G., Dhakephalkar, P.K., Niphadkar, K.B., Joshi, S.G., 2011. Fulminating septicemia due to persistent pan-resistant community-acquired metallo-β-lactamase (IMP-1)-positive Acinetobacter baumannii. Indian J. Pathol. Microbiol. 54, 180–2.

Tena, D., Martinez-Torres, J.A., Perez-Pomata, M.T., Sáez-Nieto, J.A., Rubio, V., Bisquert, J., 2007. Cutaneous infection due to Bacillus pumilus: report of 3 cases. Clin. Infect. Dis. 44, e40–2.

Thomas, X., Destoumieux-Garzón, D., Peduzzi, J., Afonso, C., Blond, A., Birlirakis,

Page 239: Discovery of antimicrobial peptides active against

References

239 | P a g e

N., Goulard, C., Dubost, L., Thai, R., Tabet, J.-C., Rebuffat, S., 2004. Siderophore peptide, a new type of post-translationally modified antibacterial peptide with potent activity. J. Biol. Chem. 279, 28233–42.

Thompson, R.E., Collin, F., Maxwell, A., Jolliffe, K.A., Payne, R.J., 2014. Synthesis of full length and truncated microcin B17 analogues as DNA gyrase poisons. Org. Biomol. Chem. 12, 1570–8.

Thumm, G., Gotz, F., 1997. Studies on prolysostaphin processing and characterization of the lysostaphin immunity factor (Lif) of Staphylococcus simulans biovar staphylolyticus. Mol. Microbiol. 23, 1251–5.

Timbermont, L., De Smet, L., Van Nieuwerburgh, F., Parreira, V.R., Van Driessche, G., Haesebrouck, F., Ducatelle, R., Prescott, J., Deforce, D., Devreese, B., Van Immerseel, F., 2014. Perfrin, a novel bacteriocin associated with netB positive Clostridium perfringens strains from broilers with necrotic enteritis. Acta Vet. Scand. 45, 40.

Trent, M.S., Worsham, L.M.S., Ernst-Fonberg, M. Lou, 1999. HlyC, the Internal Protein Acyltransferase That Activates Hemolysin Toxin: Role of Conserved Histidine, Serine, and Cysteine Residues in Enzymatic Activity As Probed by Chemical Modification and Site-Directed Mutagenesis. Biochemistry 38, 3433–9.

Tripathi, N., Cotten, C.M., Smith, P.B., 2012. Antibiotic use and misuse in the neonatal intensive care unit. Clin. Perinatol. 39, 61–8.

Tulini, F.L., Lohans, C.T., Bordon, K.C.F., Zheng, J., Arantes, E.C., Vederas, J.C., De Martinis, E.C.P., 2014. Purification and characterization of antimicrobial peptides from fish isolate Carnobacterium maltaromaticum C2: Carnobacteriocin X and carnolysins A1 and A2. Int. J. Food Microbiol. 173, 81–8.

US Congress. 2011. Generating Antibiotic Incentives Now Act of 2011. 112th Cong., H.R. 2182/S. 1734.

Valdés-Stauber, N., Scherer, S., 1994. Isolation and characterization of Linocin M18, a bacteriocin produced by Brevibacterium linens. Appl. Environ. Microbiol. 60, 3809–14.

van Belkum, M.J., Martin-Visscher, L. a, Vederas, J.C., 2011. Structure and genetics of circular bacteriocins. Trends Microbiol. 19, 411–8.

van den Ent, F., Löwe, J., 2005. Crystal structure of the ubiquitin-like protein YukD from Bacillus subtilis. FEBS Lett. 579, 3837–41.

van Heel, A.J., de Jong, A., Montalbán-López, M., Kok, J., Kuipers, O.P., 2013. BAGEL3: Automated identification of genes encoding bacteriocins and (non-

Page 240: Discovery of antimicrobial peptides active against

References

240 | P a g e

)bactericidal posttranslationally modified peptides. Nucleic Acids Res. 41, W448–53.

Van Looveren, M., Goossens, H., 2004. Antimicrobial resistance of Acinetobacter spp. in Europe. Clin. Infect. Dis. 10, 684–704.

Vassiliadis, G., Destoumieux-Garzón, D., Lombard, C., Rebuffat, S., Peduzzi, J., 2010. Isolation and characterization of two members of the siderophore-microcin family, microcins M and H47. Antimicrob. Agents Chemother. 54, 288–97.

Vassiliadis, G., Peduzzi, J., Zirah, S., Thomas, X., Rebuffat, S., Destoumieux-Garzón, D., 2007. Insight into siderophore-carrying peptide biosynthesis: Enterobactin is a precursor for microcin E492 posttranslational modification. Antimicrob. Agents Chemother. 51, 3546–53.

Vaucher, R.A., da Motta, A.D.S., Brandelli, A., 2010. Evaluation of the in vitro cytotoxicity of the antimicrobial peptide P34. Cell Biol. Int. 34, 317–23.

Venugopal, H., Edwards, P.J.B., Schwalbe, M., Claridge, J.K., Libich, D.S., Stepper, J., Loo, T., Patchett, M.L., Norris, G.E., Pascal, S.M., 2011. Structural, dynamic, and chemical characterization of a novel S-glycosylated bacteriocin. Biochemistry 50, 2748–55.

Völler, G.H., Krawczyk, J.M., Pesic, A., Krawczyk, B., Nachtigall, J., Süssmuth, R.D., 2012. Characterization of new class III lantibiotics--erythreapeptin, avermipeptin and griseopeptin from Saccharopolyspora erythraea, Streptomyces avermitilis and Streptomyces griseus demonstrates stepwise N-terminal leader processing. Chembiochem 13, 1174–83.

Vondenhoff, G.H.M., Van Aerschot, A., 2011. Microcin C: biosynthesis, mode of action, and potential as a lead in antibiotics development. Nucleosides. Nucleotides Nucleic Acids 30, 465–74.

Waisvisz, J.M., van der Hoeven, M.G., van Peppen, J., Zwennis, W.C.M., 1957. Bottromycin. I. A New Sulfur-containing Antibiotic. J. Am. Chem. Soc. 79, 4520–1.

Walsh, C.T., Fischbach, M.A., 2010. Natural products version 2.0: Connecting genes to molecules. J. Am. Chem. Soc. 132, 2469–93.

Walsh, T.R., Toleman, M.A., 2012. The emergence of pan-resistant Gram-negative pathogens merits a rapid global political response. J. Antimicrob. Chemother. 67, 1–3.

Wang, H., Fewer, D.P., Sivonen, K., 2011. Genome mining demonstrates the widespread occurrence of gene clusters encoding bacteriocins in cyanobacteria. PLoS One 6, e22384.

Page 241: Discovery of antimicrobial peptides active against

References

241 | P a g e

Watanabe, K., Praseuth, A.P., Wang, C.C.C., 2007. A comprehensive and engaging overview of the type III family of polyketide synthases. Curr. Opin. Chem. Biol. 11, 279–86.

Waterhouse, A.M., Procter, J.B., Martin, D.M.A., Clamp, M., Barton, G.J., 2009. Jalview Version 2--a multiple sequence alignment editor and analysis workbench. Bioinformatics 25, 1189–91.

Wescombe, P. a., Burton, J.P., Cadieux, P. a., Klesse, N. a., Hyink, O., Heng, N.C.K., Chilcott, C.N., Reid, G., Tagg, J.R., 2006. Megaplasmids encode differing combinations of lantibiotics in Streptococcus salivarius. Antonie van Leeuwenhoek, Int. J. Gen. Mol. Microbiol. 90, 269–80.

Wescombe, P.A., Upton, M., Renault, P., Wirawan, R.E., Power, D., Burton, J.P., Chilcott, C.N., Tagg, J.R., 2011. Salivaricin 9, a new lantibiotic produced by Streptococcus salivarius. Microbiology 157, 1290–9.

White, A.R., 2011. Effective antibacterials: at what cost? The economics of antibacterial resistance and its control. J. Antimicrob. Chemother. 66, 1948–53.

Whittaker, J.W., 2013. Cell-free protein synthesis: the state of the art. Biotechnol. Lett. 35, 143–52.

Wolanin, P., Thomason, P., Stock, J., 2002. Histidine protein kinases: key signal transducers outside the animal kingdom. Genome Biol. 3, reviews3013.1–reviews3013.8.

Wong, K., Ma, J., Rothnie, A., Biggin, P.C., Kerr, I.D., 2014. Towards understanding promiscuity in multidrug efflux pumps. Trends Biochem. Sci. 39, 8–16.

Woo, P.C.Y., Lau, S.K.P., Teng, J.L.L., Tse, H., Yuen, K.-Y., 2008. Then and now: use of 16S rDNA gene sequencing for bacterial identification and discovery of novel bacteria in clinical microbiology laboratories. Clin. Microbiol. Infect. 14, 908–34.

Wright, G.D., 2014. Something old, something new: revisiting natural products in antibiotic drug discovery. Can. J. Microbiol. 60, 147–54.

Yang, J.Y., Sanchez, L.M., Rath, C.M., Liu, X., Boudreau, P.D., Bruns, N., Glukhov, E., Wodtke, A., de Felicio, R., Fenner, A., Wong, W.R., Linington, R.G., Zhang, L., Debonsi, H.M., Gerwick, W.H., Dorrestein, P.C., 2013. Molecular networking as a dereplication strategy. J. Nat. Prod. 76, 1686–99.

Yildirim, Z., Yildirim, M., Johnson, M.G., 2004. Mode of action of lactococcin R produced by Lactococcus lactis R. Nahrung - Food 48, 145–8.

Zawada, J.F., Yin, G., Steiner, A.R., Yang, J., Naresh, A., Roy, S.M., Gold, D.S.,

Page 242: Discovery of antimicrobial peptides active against

References

242 | P a g e

Heinsohn, H.G., Murray, C.J., 2011. Microscale to manufacturing scale-up of cell-free cytokine production--a new approach for shortening protein production development timelines. Biotechnol. Bioeng. 108, 1570–8.

Zazopoulos, E., Huang, K., Staffa, A., Liu, W., Bachmann, B.O., Nonaka, K., Ahlert, J., Thorson, J.S., Shen, B., Farnet, C.M., 2003. A genomics-guided approach for discovering and expressing cryptic metabolic pathways. Nat. Biotechnol. 21, 187–90.

Zhang, C., Occi, J., Masurekar, P., Barrett, J.F., Zink, D.L., Smith, S., Onishi, R., Ha, S., Salazar, O., Genilloud, O., Basilio, A., Vicente, F., Gill, C., Hickey, E.J., Dorso, K., Motyl, M., Singh, S.B., 2008a. Isolation, Structure, and Antibacterial Activity of Philipimycin, A Thiazolyl Peptide Discovered from Actinoplanes philippinensis MA7347. J. Am. Chem. Soc. 130, 12102–10.

Zhang, C., Zink, D.L., Ushio, M., Burgess, B., Onishi, R., Masurekar, P., Barrett, J.F., Singh, S.B., 2008b. Isolation, structure, and antibacterial activity of thiazomycin A, a potent thiazolyl peptide antibiotic from Amycolatopsis fastidiosa. Bioorg. Med. Chem. 16, 8818–23.

Zhang, J., Xin, L., Shan, B., Chen, W., Xie, M., Yuen, D., Zhang, W., Zhang, Z., Lajoie, G.A., Ma, B., 2012. PEAKS DB: De Novo Sequencing Assisted Database Search for Sensitive and Accurate Peptide Identification. Mol. Cell. Proteomics 11, M111.010587.

Zhang, R., Eggleston, K., Rotimi, V., Zeckhauser, R.J., 2006. Antibiotic resistance as a global threat: evidence from China, Kuwait and the United States. Global. Health 2, 6.

Zhou, K., Zhou, W., Li, P., Liu, G., Zhang, J., Dai, Y., 2008. Mode of action of pentocin 31-1: An antilisteria bacteriocin produced by Lactobacillus pentosus from Chinese traditional ham. Food Control 19, 817–22.

Zhu, F., Ma, X.H., Qin, C., Tao, L., Liu, X., Shi, Z., Zhang, C.L., Tan, C.Y., Chen, Y.Z., Jiang, Y.Y., 2012. Drug discovery prospect from untapped species: indications from approved natural product drugs. PLoS One 7, e39782.

Zhu, H., Swierstra, J., Wu, C., Girard, G., Choi, Y.H., Van Wamel, W., Sandiford, S.K., van Wezel, G.P., 2014. Eliciting antibiotics active against the ESKAPE pathogens in a collection of actinomycetes isolated from mountain soils. Microbiol. (United Kingdom) 160, 1714–26.

Zschüttig, A., Zimmermann, K., Blom, J., Goesmann, A., Pöhlmann, C., Gunzer, F., 2012. Identification and characterization of microcin S, a new antibacterial peptide produced by probiotic Escherichia coli G3/10. PLoS One 7, e33351.

Zuckermann, R.N., Kerr, J.M., Kent, S.B.H., Moos, W.H., 1992. Efficient method

Page 243: Discovery of antimicrobial peptides active against

References

243 | P a g e

for the preparation of peptoids [oligo(N-substituted glycines)] by submonomer solid-phase synthesis. J. Am. Chem. Soc. 114, 10646–7.

Zukher, I., Novikova, M., Tikhonov, A., Nesterchuk, M. V, Osterman, I.A., Djordjevic, M., Sergiev, P. V, Sharma, C.M., Severinov, K., 2014. Ribosome-controlled transcription termination is essential for the production of antibiotic microcin C. Nucleic Acids Res. 42, 11891–902.