86
Heterologous expression and characterization of the antibacterial lasso peptide LP2006 by Gaelen Moore A thesis submitted in conformity with the requirements for the degree of Master of Science Department of Biochemistry University of Toronto © Copyright by Gaelen Moore, 2019

Heterologous expression and characterization of the

  • Upload
    others

  • View
    11

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Heterologous expression and characterization of the

Heterologous expression and characterization of the antibacterial lasso peptide LP2006

by

Gaelen Moore

A thesis submitted in conformity with the requirements for the degree of Master of Science

Department of Biochemistry University of Toronto

© Copyright by Gaelen Moore, 2019

Page 2: Heterologous expression and characterization of the

ii

Heterologous expression and characterization of the antibacterial

lasso peptide LP2006

Gaelen Moore

Master of Science

Department of Biochemistry University of Toronto

2019

Abstract

The lasso peptides are a class of ribosomally synthesized peptide natural products with diverse

bioactivities and structures resembling a lasso. Although the targets of several antibacterial lasso

peptides have been investigated to date, the majority remain uncharacterized. Among those that

have been characterized, the antibacterial lasso peptides have diverse targets and unique

mechanisms of action. One antibacterial lasso peptide with a unique structure, LP2006, is the

only member of the class IV peptides. Currently the target and mechanism of action of LP2006

remains unknown. The aim of this study is to develop a system for the heterologous expression

of LP2006 to allow for the study of its target and mode of action. I demonstrate that LP2006 can

be heterologously expressed using Streptomyces coelicolor M1146, and that purified LP2006

does not appear to activate the cell wall stress response gene liaI.

Page 3: Heterologous expression and characterization of the

iii

Acknowledgments

The past two years of my Master’s research project would not have been possible without the

support of many people.

Firstly, I would like to thank Dr. Justin Nodwell for providing me with the opportunity to work

in his laboratory. Despite his busy schedule, he always manages to make plenty of time to meet

with his students and never fails to inspire his students. I would also like to thank my committee

members, Dr. Karen Maxwell and Dr. Alex Ensminger for their input and thoroughness over the

course of my project.

I am deeply grateful to Dr. Sheila Pimental-Elardo for her guidance and support in conducting

this project. I have thoroughly enjoyed our discussions about marine natural products in addition

to our non-scientific discussions. Sheila's kindness and empathy have had a very positive

influence on the lab. I am very thankful to have had the opportunity to meet and work with all of

the Nodwell Lab members. I am appreciative of not only their scientific suggestions, but also

their friendship, which has made my time in the lab highly enjoyable.

I would also like to thank my undergraduate thesis project mentor, Sohee Yun, for her guidance

and enthusiasm. She inspired me to pursue research and she was always a positive presence

during my time as her trainee. Finally, I would like to thank my parents for their steadfast

support over the years.

Page 4: Heterologous expression and characterization of the

iv

Table of Contents

Acknowledgments.......................................................................................................................... iii

Table of Contents ........................................................................................................................... iv

List of Tables ................................................................................................................................ vii

List of Figures .............................................................................................................................. viii

Chapter 1 Introduction .....................................................................................................................1

Introduction .................................................................................................................................1

1.1 RiPPs ....................................................................................................................................1

1.1.1 Thiopeptides .............................................................................................................4

1.1.1.1 Biosynthesis of Thiopeptides ....................................................................4

1.1.1.2 Antibiotic Targets of Thiopeptides ............................................................5

1.1.2 Lanthipeptides ..........................................................................................................6

1.1.2.1 Biosynthesis of lanthipeptides ...................................................................9

1.1.2.2 Activity of lanthipeptides ........................................................................11

1.1.3 Lasso peptides ........................................................................................................12

1.1.3.1 Classification of lasso peptides ...............................................................13

1.1.3.2 Biosynthesis of lasso peptides .................................................................17

1.1.3.3 Activity of lasso peptides ........................................................................20

1.1.3.4 Discovery and Heterologous expression of lasso peptides ......................24

1.2 Aim of this work ................................................................................................................26

Chapter 2 Materials and Methods ..................................................................................................27

Materials and Methods ..............................................................................................................27

2.1 General experimental procedures ......................................................................................27

2.1.1 Materials ................................................................................................................27

2.1.2 Strains and plasmids used ......................................................................................27

2.1.3 Primers Used ..........................................................................................................28

Page 5: Heterologous expression and characterization of the

v

2.1.4 Culture conditions ..................................................................................................29

2.1.5 Heterologous expression of LP2006 ......................................................................29

2.2 Isolation and purification of bioactive metabolites ............................................................30

2.2.1 Metabolite extraction .............................................................................................30

2.2.2 Flash chromatography purification ........................................................................30

2.2.3 High-performance liquid chromatography purification .........................................31

2.2.4 Liquid chromatography mass spectrometry analysis .............................................31

2.3 Susceptibility testing ..........................................................................................................32

2.3.1 Disk diffusion assays .............................................................................................32

2.3.2 Broth microtiter dilution assay...............................................................................32

2.4 Target identification ...........................................................................................................32

2.4.1 LacZ reporter assay ................................................................................................32

Results and discussion ..............................................................................................................33

3.1 Nocardiopsis sp. HB141 extract testing.............................................................................33

3.1.1 Extracts of Nocardiopsis sp. HB141 have antibacterial activity ...........................33

3.1.2 Nocardiopsis sp. HB141 is a producer of an antibacterial lasso peptide, LP2006 ...................................................................................................................34

3.1.3 Purification of LP2006 ...........................................................................................35

3.2 Heterologous expression of LP2006 ..................................................................................38

3.2.1 Heterologous expression in Escherichia coli .........................................................38

3.2.2 Heterologous expression in Streptomyces coelicolor M1146 ................................43

3.3 Bioactivity of LP2006 ........................................................................................................46

Conclusions and future directions .............................................................................................48

References ......................................................................................................................................50

Appendix 1 Screen for novel bioactive natural products from marine bacteria ...........................62

Appendix 1 ................................................................................................................................62

Page 6: Heterologous expression and characterization of the

vi

5.1 Introduction ........................................................................................................................62

5.2 Methods..............................................................................................................................62

5.2.1 Bioactivity screen...................................................................................................62

5.2.1.1 Collection and Isolation of maritime strains ...........................................62

5.2.1.2 Culture conditions ...................................................................................63

5.2.1.3 Broth microtiter dilution and disk diffusion assay ..................................63

5.2.2 Isolation of the Marinobacter sp. N33 bioactive metabolite(s) .............................64

5.2.2.1 Metabolite extraction ...............................................................................64

5.2.2.2 Flash chromatography and HPLC purification ........................................64

5.2.3 Genomic studies .....................................................................................................65

5.2.3.1 Genomic DNA extraction ........................................................................65

5.2.3.2 Phylogenetic analysis ..............................................................................65

5.2.3.3 Whole genome sequencing of Marinobacter sp. N33 .............................66

5.3 Results and discussion .......................................................................................................66

5.3.1 Screen of marine bacteria .......................................................................................66

5.3.1.1 Phylogenetic analysis ..............................................................................66

5.3.1.2 Bioactivity screening ...............................................................................70

5.3.2 Marinobacter sp. N33 extract testing ....................................................................72

5.3.2.1 Bioactivity testing ....................................................................................72

5.3.2.2 Genome sequencing and genome mining ................................................73

5.3.2.3 Bioactivity guided fractionation and purification ....................................75

Copyright Acknowledgements.......................................................................................................77

Page 7: Heterologous expression and characterization of the

vii

List of Tables

Table 1.1. Characterized lasso peptides and their tested bioactivity. ........................................... 14

Table 1.2. Heterologously produced lasso peptides. ..................................................................... 25

Table 2.1. Strains and plasmids used in this study. ...................................................................... 27

Table 2.2. Primers used in this study. ........................................................................................... 28

Table 3.1. Ion comparison of HB141 mass and LP2006 .............................................................. 35

Table 3.2. Sequence identity of LP2006 biosynthetic proteins of Nocardiopsis alba ATCC BAA-

2165 and Nocardiopsis sp. TP-A0876 compared to Nocardiopsis alba DSM 43377. .................. 39

Table 3.3. Comparison of LP2006 masses from the Nocardiopsis sp. HB141 and those produced

by heterologous expression in S. coelicolor M1146. .................................................................... 46

Table 5.1. Identity and characteristics of maritime strains ........................................................... 67

Table 5.2. Comparison of Marinobacter sp. N33 genome statistics to close relatives................. 75

Page 8: Heterologous expression and characterization of the

viii

List of Figures

Figure 1.1. Structures of selected commercially used RiPPs.......................................................... 2

Figure 1.2. Generalized RiPP Biosynthesis .................................................................................... 3

Figure 1.3. Biosynthesis of the thiomuracin core scaffold ............................................................. 5

Figure 1.4. Post-translational modifications characteristic in lanthipeptides ................................. 7

Figure 1.5. Distribution of lasso peptide, lanthipeptide and thiopeptide clusters in Streptomyces

genomes .......................................................................................................................................... 8

Figure 1.6. Mechanisms of lanthipeptide synthesis ...................................................................... 10

Figure 1.7. Classes of lanthionine synthetases.............................................................................. 11

Figure 1.8. Lasso peptide cyclization and structure ...................................................................... 13

Figure 1.9. Classes of lasso peptides ............................................................................................ 14

Figure 1.10. Lasso peptide gene clusters and biosynthesis ........................................................... 18

Figure 1.11. Proposed mechanism of microcin J25 biosynthesis ................................................. 19

Figure 1.12. Antibacterial targets of lasso peptides ...................................................................... 20

Figure 3.1. Disk diffusion assay of Nocardiopsis sp. HB141 extract ........................................... 33

Figure 3.2. Extracted ion chromatogram of LP2006 M+2H mass ................................................ 34

Figure 3.3. Mass spectrum of the Nocardiopsis sp. HB141 mass of 1002.9335 m/z ................... 35

Figure 3.4. Disk diffusion assay from flash chromatography fractionated Nocardiopsis sp.

HB141 extract ............................................................................................................................... 36

Figure 3.5. HPLC chromatogram of first-round purification of LP2006 ..................................... 37

Figure 3.6. HPLC chromatogram of second-round purification of LP2006 ................................. 37

Page 9: Heterologous expression and characterization of the

ix

Figure 3.7. LP2006 biosynthetic gene cluster from N. alba DSM 43377 .................................... 39

Figure 3.8. Vectors used for E. coli heterologous expression ...................................................... 40

Figure 3.9. Protein expression testing of LpeCEB ....................................................................... 42

Figure 3.10. Construct used for S. coelicolor heterologous expression ........................................ 44

Figure 3.11. Detection of heterologously expressed LP2006 by mass spectrometry ................... 45

Figure 3.12. Antibacterial MIC testing of pure LP2006 ............................................................... 47

Figure 3.13. LP2006 does not activate the cell wall stress response gene liaI in B. subtilis 1A980

....................................................................................................................................................... 47

Figure 5.1. Phylogenetic tree of Nodwell Maritime Collection strains. ....................................... 69

Figure 5.2. Screen of Nodwell Maritime Collection strains ......................................................... 70

Figure 5.3. Distribution of growth inhibition values. ................................................................... 71

Figure 5.4. Phylogenetic tree and antibacterial activity of Nodwell Maritime Collection strains

tested against B. subtilis ................................................................................................................ 72

Figure 5.5. Disk diffusion assay of Marinobacter sp. N33 crude extract using B. subtilis JH642.

....................................................................................................................................................... 73

Figure 5.6. Circular representation of the Marinobacter sp. N33 genome ................................... 74

Figure 5.7. UV chromatogram of the purification of fraction 19 from flash chromatography by

HPLC ............................................................................................................................................ 76

Page 10: Heterologous expression and characterization of the

1

Chapter 1 Introduction

Introduction

The discovery of antibiotics has produced enormous benefits for both human and animal

medicine. Today, many routine medical procedures would not be possible without effective

antibiotics to prevent post-procedural infections. In the clinic, natural products and derivatives

remain the primary source of novel antibiotics, even decades after the golden era of natural

product-based antibiotic discovery of the 1940s to 1960s1. With the advent of next-generation

sequencing, it has become apparent that our knowledge of natural product chemistry and biology

is incredibly incomplete. Further exploration of the natural product chemical space will

undoubtedly lead to new therapeutics.

With the escalating threat of antibiotic resistance, it is important to continue to develop novel

antibiotics to prevent the emergence of pathogens that are untreatable with our existing antibiotic

arsenal. While antibiotics synthesized by nonribosomal peptide synthases and polyketide

synthases are familiar in medicine, one overlooked class of natural products are the ribosomally

synthesized and post-translationally modified peptides (RiPPs) which have a long history of

documented antibiotic activity but have not been thoroughly explored for their therapeutic and

industrial applications.

1.1 RiPPs

Peptide synthesis can occur either ribosomally or nonribosomally. Nonribosomal peptides are

familiar to many – the class includes medically important compounds such as the beta-lactams,

the glycopeptides, the depsipeptides, cyclosporine and daptomycin. The nonribosomal peptides

have had an enormous impact on science and medicine. In contrast, ribosomally synthesized and

post-translationally modified peptides (RiPPs) are less well-known but possess impressive

structural diversity and potent bioactivity2. The RiPPs, which are produced by all domains of

life, are comparatively understudied in spite of their interesting bioactivity and therapeutic

potential. Commercialized RiPPs include thiostrepton, an antibacterial used in veterinary

Page 11: Heterologous expression and characterization of the

2

medicine, nisin, a food preservative, and omega-conotoxin MVIIA, which has been developed

into the synthetic antinociceptive drug, Ziconotide (Figure 1.1)3–6.

Figure 1.1. Structures of selected commercially used RiPPs. A. Structure of thiostrepton. B.

Structure of Ziconotide, the synthetic analogue of the omega-conotoxin MVIIA. C. Structure of

nisin. The post-translationally modified residues are highlighted in red.

The RiPPs share a biosynthetic process, unifying the family which would otherwise be

unrelated2. Like other classes of microbial natural products, the genes required for RiPP

biosynthesis are clustered within the genome. RiPPs are genetically encoded in the form of a

Page 12: Heterologous expression and characterization of the

3

precursor peptide, which consists of an N-terminal leader sequence and a C-terminal core

sequence (Figure 1.2). The core sequence undergoes a series of post-translational modifications,

which are introduced by modification enzymes encoded in the biosynthetic gene cluster of the

RiPP. RiPP biosynthetic gene clusters also encode a leader peptidase, which cleaves the N-

terminal leader sequence from the C-terminal core, ultimately yielding the mature RiPP. RiPP

clusters may sometimes encode genes for the export of the mature RiPPs or the regulation of the

production of the RiPPs.

Figure 1.2. Generalized RiPP Biosynthesis. A. General biosynthetic cluster of a RiPP natural

product. B. Generalized biosynthetic logic of RiPP biosynthesis. Adapted from Tan et al.,

Antibiotics, 20197.

The RiPPs possess a unique ability to generate substantial chemical diversity at a low metabolic

cost. The RiPP model has evolved to produce structural diversity as many of the enzymes that

introduce post-translational modifications are permissive to mutations in the core sequence and

primarily recognize the leader sequence. In fact, certain classes of RiPPs contain core regions

that are naturally hypervariable8. The broad chemical diversity of RiPPs, combined with their

suitability for genome mining, heterologous expression and their potential for engineering proves

advantageous in their therapeutic development. In the following sections I will discuss the

Page 13: Heterologous expression and characterization of the

4

structural characteristics, biosynthesis and bioactivity of three RiPP classes which have a large

body of literature and well-documented antibiotic activity.

1.1.1 Thiopeptides

Few classes of RiPPs have been studied as thoroughly as the thiopeptides. The thiopeptides are

macrocyclic peptides characterized by sulfur-rich heterocycles and a central 6-membered

nitrogen-containing oxidized ring. There are currently more than 100 known thiopeptides, which

are classified into five categories based on the structure and oxidation state of the central

nitrogen-containing ring9. The first thiopeptide discovered was micrococcin in 1948, although

the first thiopeptide gene clusters were only identified in 200910–12. Most thiopeptides are potent

antibacterial translation inhibitors and some have long been used in agricultural feed and

veterinary treatments13–16.

1.1.1.1 Biosynthesis of Thiopeptides

For more than half a century after the discovery of the micrococcin in 1948 it was unclear

whether the thiopeptides were synthesized ribosomally or nonribosomally. Finally, in 2009,

several thiopeptide gene clusters were published, revealing that thiopeptides indeed belong to the

RiPP family10,11,17,18.

After translation, thiopeptide biosynthesis typically involves three steps: installation of the

sulfur-rich heterocycles, installation of the central nitrogen-containing ring, and leader peptide

removal by the leader peptidase, producing the mature thiopeptide. The first step involves the

dehydration of Cys and sometimes Ser/Thr residues to produce the sulfur and oxygen-containing

heterocycles, thiazol(in)e or oxazol(in)e (Figure 1.3)19. This step is catalyzed by a trimeric

heterocycle synthetase, and is shared with the closely related class of RiPPs the linear azol(in)e-

containing peptides, of which the DNA-gyrase inhibitor microcin B17 is a member20,21. Next,

unmodified Ser/Thr residues are dehydrated by a LanB-like dehydratase to yield dehydroalanine

(Dha) and dehydrobutyrine (Dhb)22. Finally, a [4 + 2] cycloaddition reaction occurs between two

Dha residues and an amide backbone to produce a 6-membered nitrogen-containing ring, which

constitutes part of the macrocycle23. Subsequent modifications such as hydroxylation,

methylation, or the addition of a tryptophan-derived macrocycle have been reported for certain

Page 14: Heterologous expression and characterization of the

5

thiopeptides11,18,24,25. Finally, after all post-translational modifications the leader peptide is

cleaved from the core peptide, resulting in the mature thiopeptide.

Figure 1.3. Biosynthesis of the thiomuracin core scaffold. Thiazole heterocycles are depicted

in purple; dehydroalanine residues are depicted in green; and the central nitrogen-containing

heterocycle is depicted in orange. Figure adapted from Hudson et al., 201526.

1.1.1.2 Antibiotic Targets of Thiopeptides

The thiopeptides have long been known for their potent antibacterial activity, but their

development into clinical therapeutics has been hampered by poor bioavailability and high rates

of emergence of resistance27. Thiopeptides often have antibacterial activity with nanomolar

potency against Gram-positive bacteria but lack activity against Gram-negatives due to their

inability to bypass the outer membrane12. Thiopeptides are mostly protein synthesis inhibitors

Page 15: Heterologous expression and characterization of the

6

and can inhibit translation through one of two mechanisms, dependent on the macrocycle size.

Thiopeptides containing 26 or 32-atom macrocycles, such as thiostrepton and micrococcin,

inhibit protein translation by binding directly to the ribosome, at the interface of the 23s rRNA

and the N-terminal domain of ribosomal protein L1128. This results in stabilization of the N-

terminal domain of ribosomal protein L11, restricting the interaction with elongation factor G

and preventing the conformational shifts necessary for translocation to occur29. Resistance to this

class of thiopeptides can occur through the three mechanisms: the deletion or mutation of the

gene encoding the ribosomal protein L11, mutation of A1067 or A1095 of 23S rRNA (E. coli

nomenclature) or methylation of the 23S rRNA30–35.

Thiopeptides that contain 29-atom macrocycles, such as GE2270A, inhibit translation by binding

to elongation factor Tu (EF-Tu), where they prevent the binding of amino-acyl tRNAs to the

protein. GE2270A partially occludes the binding of the aminoacyl-tRNA and GTP, but not GDP.

Resistance to EF-Tu-targeting thiopeptides increases the affinity of elongation factor Tu for

aminoacyl-tRNA in the presence of EF-Tu-targeting thiopeptides36,37.

Although the vast majority of thiopeptides are bacterial translation inhibitors, there are several

for which different activities have been reported. Notably, the cyclothiazomycins are reported to

inhibit RNA polymerase, while lactazole has no reported antibiotic activity38,39.

1.1.2 Lanthipeptides

The lanthipeptides are another relatively well-studied class of RiPPs which contain the

characteristic thioether amino acids lanthionine and methyllanthionine. The (methyl)lanthionine

bridges are introduced via Michael addition between free cysteines and the alkene-containing

amino acids, dehydroalanine (Dha) and dehydrobutyrine (Dhb) (Figure 1.4). This class of RiPPs

has been the subject of much research interest not only because of the interesting biology and

bioactivity, but also because of the potential of industrial and medical applications. Much of the

interest has been surrounding the prototypical lanthipeptide nisin, which was first discovered in

1928 and has been widely used as a food preservative for several decades40,41. Nisin is also

currently being pursued as a treatment for bovine mastitis42. Encouragingly, in spite of its heavy

use in food preservation, widespread resistance to nisin has yet to emerge. Beyond nisin, the

Page 16: Heterologous expression and characterization of the

7

lanthipeptide NVB302 was found to have strong in vitro results in the treatment of Clostridioides

difficile infections when compared to vancomycin43.

Figure 1.4. Post-translational modifications characteristic in lanthipeptides. The

(methyl)lanthionine, dehydroalanine and dehydrobutyrine modifications are characteristic of all

classes of lanthipeptides, while the labionin modification is found only in the class III

lanthipeptides. Figure adapted from Knerr and van der Donk, 201244.

Lanthipeptides, along with other classes of RiPPs, are produced across many bacterial lineages,

particularly by Actinobacteria. In fact, the Streptomyces, which are known producers of many

important secondary metabolites of polyketide and nonribosomal peptide origin, appear to be

prolific producers of RiPPs as well (Figure 1.5)7. In a random sampling of 50 complete

Streptomyces genomes, the genomes were found to encode as many as 8 RiPPs, with

lanthipeptide clusters occurring more frequently compared to lasso peptide and thiopeptide

clusters. Some lanthipeptides are known to be important for Streptomyces development, which

may partially explain the frequency at which lanthipeptide clusters are found in the genus45.

Page 17: Heterologous expression and characterization of the

8

Figure 1.5. Distribution of lasso peptide, lanthipeptide and thiopeptide clusters in

Streptomyces genomes. 50 randomly-selected Streptomyces genomes, were analyzed using

AntiSMASH 5.0 to detect for RiPP biosynthetic gene clusters. Lasso peptide clusters are shown

in red, lanthipeptide clusters in blue, and thiopeptide clusters in green. From each genome, the

genes atpD, gyrA, recA, rpoB, trpB and the 16S rRNA gene were compiled for cladogram

construction using FastTree 2.0 and visualized using the interactive Tree of Life.46–48 Figure

from Tan et al., 20197.

Page 18: Heterologous expression and characterization of the

9

Lanthipeptides were originally categorized into two types: type A, including nisin and subtilisin

which are long, flexible molecules and type B, including duramycin and cinnamycin, which are

globular molecules49. With the discovery of lanthipeptides that did not fit well into these

categories, a new classification scheme was proposed based on the enzymes used to synthesize

lanthipeptides50,51. To date, there are four distinct classes of lanthipeptide biosynthetic enzymes,

which will be discussed in the following section.

1.1.2.1 Biosynthesis of lanthipeptides

The characteristic lanthionine or methyllanthionine cross-bridges are introduced in lanthipeptides

in two steps. First, Ser and Thr residues are selectively dehydrated to dehydroalanine or

dehydrobutyrine, respectively. Next, the unsaturated Dha and Dhb amino acids undergo

nucleophilic attack by select cysteines through a Michael reaction to generate lanthionine and

methyllanthionines and form cross-links in the peptide structure (Figure 1.6).

Page 19: Heterologous expression and characterization of the

10

Figure 1.6. Mechanisms of lanthipeptide synthesis. A. Mechanism of dehydration in LanB-

type enzymes that synthesize class I lanthipeptides. B. Mechanism of dehydration in the

synthesis of class II-IV lanthipeptides. C. Mechanism of cyclization via Michael addition used in

lanthipeptide synthesis. Figure adapted from Ortega and van der Donk, 201652.

There are four classes of lanthipeptides, which are categorized according to the enzymes that

catalyze (methyl)lanthionine installation (Figure 1.7)44. Class I lanthipeptides have dedicated

proteins for dehydration and cyclization, named LanB and LanC, respectively. Class II have

fused N-terminal dehydratase and C-terminal cyclase enzyme functionalities, whereby the

enzymes are termed LanM. The N-terminal dehydratase domain does not display sequence

homology with other lanthionine synthetases, although the C-terminal cyclase domain is

homologous to the LanC cyclase protein of class I lanthionine synthetases. Class III and IV

enzymes both contain lyase and kinase domains and C-terminal cyclase domains that are

homologous to LanC cyclases, although the class III cyclase domain is missing several

Page 20: Heterologous expression and characterization of the

11

conserved metal binding residues. The class III enzymes are the only class able to introduce

labionin structures, an additional carbon-carbon crosslink of the lanthionine amino acid53.

Figure 1.7. Classes of lanthionine synthetases. The conserved motifs are highlighted. Figure

adapted from Knerr and van der Donk, 201244.

Ser/Thr dehydration can occur through one of two mechanisms, which is dependent on the

biosynthetic machinery performing the dehydration. Class I enzymes convert Ser/Thr to

Dha/Dhb in a tRNA-dependent process54. The LanB enzymes transfer a glutamate from tRNAGlu

to Ser/Thr, thereby activating it for dehydration (Figure 1.6). In contrast, the class II-IV enzymes

transfer the ɣ-phosphate of ATP to activate Ser/Thr for dehydration55. After Cys-dependent

Michael addition to Dha/Dhb to produce (methyl)lanthionine residues, additional post-

translational modifications are sometimes introduced including decarboxylation, hydroxylation

and additional cross links between two amino acids56,57.

1.1.2.2 Activity of lanthipeptides

To date, only class I and II lanthipeptides are known to have antimicrobial activity. Nisin, the

prototypical lanthipeptide, has been shown to inhibit Gram-positive bacteria at single-digit

nanomolar concentrations which comparable to the potency of many clinical antibiotics58. Many

experiments have been performed to investigate the mechanism of antibacterial activity of nisin.

Nisin has since been found to target the bacterial cell wall by binding to the diphosphate of lipid

II and forming pores in the membrane, ultimately resulting in a loss of membrane potential58–60.

Page 21: Heterologous expression and characterization of the

12

In addition to its pore-forming ability, nisin was later found to have a second activity, whereby it

sequesters lipid II at the cell division site, blocking cell wall synthesis and cell division61.

Certain class II lanthipeptides such as mersacidin complex with lipid II, but do not form pores62.

It is believed that lanthipeptides of this class bind to a region of lipid II that encompasses the N-

acetylglucosamine and the diphosphate, in contrast to nisin which has been shown to bind to the

diphosphate and not the N-acetylglucosamine. Importantly, lanthipeptides have also been

reported to have activities other than antibiotic. The SapB, SapT and catenulipeptin peptides,

discovered from S. coelicolor, S. tendae and Catenulispora acidiphila, respectively, are

morphogens in Streptomyces, enabling the formation of aerial mycelium45,63,64.

Although in vivo antimicrobial lanthipeptide resistance remains very rare, many mechanisms of

resistance to lanthipeptides have been reported in vitro65. In lanthipeptide-producing organisms,

self-resistance typically occurs through the expression of an ABC transporter or an immunity

protein, which localizes to the membrane66. In vitro, several groups have reported nisin

resistance occurring through changes in the phospholipid composition or rigidity of the

membrane67,68. Lysine esterification of membrane lipids through MprF (multiple peptide

resistance factor), reduces the net negative charge of the membrane, providing resistance against

many cationic antimicrobial peptides including nisin69. A plasmid-encoded nisin resistance

protein from certain strains of Lactococcus lactis was found to specifically degrade nisin through

proteolysis70,71. In addition to these mechanisms, a number of other genes have been associated

with lanthipeptide resistance including cell wall modification systems and two-component

systems65.

1.1.3 Lasso peptides

Lasso peptides are a unique class of RiPPs with a simple structure, yet interesting and diverse

bioactivities. The structure of the lasso peptides consists of a macrocycle formed by a

macrolactam linkage between the N-terminal amine of the peptide and a glutamate or aspartate

residue in the 7th to 9th position. The C-terminal tail of the peptide is threaded through the N-

terminal ring, producing a structure which resembles a lasso (Figure 1.8). The lasso peptide

structure is often stabilized by disulfide bridges or bulky residues such as Trp, Tyr or Phe, which

are positioned above and below the ring to prevent unthreading. As a result, the lasso peptides

Page 22: Heterologous expression and characterization of the

13

are surprisingly stable both to heat and proteolytic degradation. In fact, some lasso peptides can

withstand temperatures up to 95°C for 8 hours, while others can withstand an autoclave cycle,

although this feature is not universal72–74.

Figure 1.8. Lasso peptide cyclization and structure. The teal circles denote amino acids that

form the N-terminal macrolactam ring, and grey circles represent amino acids that form the C-

terminal tail. The reaction is catalyzed by an ATP-dependent lasso cyclase. Figure from Tan et

al., 20197.

1.1.3.1 Classification of lasso peptides

The lasso peptides are classified into four structural classes based on the location of disulfide

bridges in their structure (Figure 1.9).Class I peptides contain two disulfide bridges; class II

contains no disulfide bridges; class III contains a single disulfide bridge joining the N-terminal

ring and the C-terminal tail; while class IV peptides contain a single intra-tail disulfide. An

estimate in 2017 stated that 96% of bioinformatically-predicted lasso peptides belong to class

II75. To date, 69 lasso peptides have been characterized, 47 of which belong to a unique family

(Table 1.1).

While, the vast majority of characterized lasso peptides belong to class II, there is currently only

a single member of the class IV peptides, LP200675. LP2006 was identified using the genome

mining algorithm RODEO and was reported to have antimicrobial activity against several Gram-

positive bacteria. These included Bacillus anthracis, vancomycin-resistant Enterococcus faecium

Page 23: Heterologous expression and characterization of the

14

and Mycobacterium smegmatis which were inhibited at 6.25, 12.5 and 12.5 μM concentrations of

LP2006, respectively. Although LP2006 has antibacterial activity, it is not known what the target

of the antibiotic is, and whether its mechanism of action differs from those of other lasso

peptides because of its unique structure.

Figure 1.9. Classes of lasso peptides. The teal circles represent amino acids that form the

macrolactam ring, and grey circles represent amino acids that form the C-terminal tail. The four

classes differ in their disulfide bonds, which are depicted in yellow. Figure from Tan et al.,

20197.

Table 1.1. Characterized lasso peptides and their tested bioactivity.

Peptide Name Producing Organism Sequence Antibacterial

Activity?

Class I

Humidimycin MDN 0010

Streptomyces humidus F-100.629

CLGIGSCDDFAGCGYAIVCFW Not Tested

Specialicin Streptomyces sp. CLGVGSCVDFAGCGYAVVCFW Yes

Siamycin-1 Streptomyces sp. CLGVGSCNDFAGCGYAIVCFW Yes

Siamycin-2 Streptomyces sp. CLGIGSCNDFAGCGYAIVCFW Not Tested

Siamycin-3 Streptomyces sp. CLGIGSCNDFAGCGYAVVCFW Yes

SSV 2083/Sviceucin Streptomyces sviceus CVWGGDCTDFLGCGTAWICV Yes

Class II

Albusnodin Streptomyces albus GQGGGQSEDKRRAYNC Not Tested

Page 24: Heterologous expression and characterization of the

15

Achromosin Streptomyces

achromogenes GIGSQTWDTIWLWD Yes

Acinetodin Acinetobacter

gyllenbergii GGKGPIFETWVTEGNYYG Yes

Actinokineosin Actinokineospora

spheciospongiae GYPFWDNRDIFGGYTFIG Yes

Astexin-1 (23 residue) Asticcacaulis

excentricus GLSQGVEPDIGQTYFEESRINQD Not Tested

Astexin-1 (19 residue) Asticcacaulis

excentricus GLSQGVEPDIGQTYFEESR No

Astexin-2 Asticcacaulis

excentricus GLTQIQALDSVSGQFRDQLGL Not Tested

Astexin-3 Asticcacaulis

excentricus GPTPMVGLDSVSGQYWDQHAPL Not Tested

Anantin A Streptomyces

coerulescens GFIGWGNDIFGHYSGDF Not Tested

Anantin B1 Streptomyces sp. NRRL S-146

GFIGWGNDIFGHYSGGF No

Anantin B2 Streptomyces sp. NRRL S-146

GFIGWGNDIFGHYSGD Yes

Benenodin-1 Asticcacaulis

benevestitus GVGFGRPDSILTQEQAKPMGLDRD Not Tested

Brevunsin Brevundimonas

diminuta DGMGEEFIEGLVRDSLYPPAG No

Burhizin Burkholderia

rhizoxinica GGAGQYKEVEAGRWSDR Not Tested

Capistruin Burkholderia

thailandensis GTPGFQTPDARVISRFGFN Yes

Caulonodulin-1 Caulobacter sp. GDVLNAPEPGIGREPTG Not Tested

Caulonodulin-2 Caulobacter sp. GDVLFAPEPGVGRPPMG Not Tested

Caulonodulin-3 Caulobacter sp. GQIYDHPEVGIGAYGCE Not Tested

Caulonodulin-4 Caulobacter sp. SFDVGTIKEGLVSQYYFA Not Tested

Caulonodulin-5 Caulobacter sp. SIGDSGLRESMSSQTYWP Not Tested

Caulonodulin-6 Caulobacter sp. AGTGVLLPETNQIKRYDPA Not Tested

Caulonodulin-7 Caulobacter sp. SGIGDVFPEPNMVRRWD Not Tested

Caulosegnin-1 Caulobacter segnis GAFVGQPEAVNPLGREIQG No

Caulosegnin-2 Caulobacter segnis GTLTPGLPEDFLPGHYMPG No

Caulosegnin-3 Caulobacter segnis GALVGLLLEDITVARYDPM No

Chaxapeptin Streptomyces

leeuwenhoekii C58 GFGSKPLDSFGLNFF Yes

Citrocin Citrobacter pasteurii GGVGKIIEYFIGGGVGRYG Yes

Citrulassin Streptomyces albulus LLGLAGNDRLVLSKN No

Page 25: Heterologous expression and characterization of the

16

Fusilassin/Fuscanodin Thermobifida fusca WYTAEWGLELIFVFPRFI No

Klebsidin Klebsiella pneumoniae GSDGPIIEFFNPNGVMHYG Yes

Lariatin A Rhodococcus sp. K01-BI0171

GSQLVYREWVGHSNVIKP Yes

Lariatin B Rhodococcus sp. K01-BI0171

GSQLVYREWVGHSNVIKPGP Yes

Lagmysin Streptomyces sp. LAGQGSPDLLGGHSLL No

Lassomycin Lentzea kentuckyensis GFIGWGNDIFGHYSGDF Yes

Microcin J25 Escherichia coli AY25 GGAGHVPEYFVGIGTPISFYG Yes

Cattlecin/Moomysin Streptomyces cattleya SYHWGDYHDWHHGWYGWWDD No

Paeninodin Paenibacillus dendritiformis C454

AGPGTSTPDAFQPDPDEDVHYDS No

Propeptin-1 Microbispora sp. SNA-115

GYPWWDYRDLFGGHTFISP Yes

Propeptin-2 Microbispora sp. SNA-115

GYPWWDYRDLFGGHTFI Yes

Pseudomycoidin Bacillus pseudomycoides DSM 12442

QVFEDEDEQGALHHN Not Tested

RES 701 1 Streptomyces sp. RE-701

GNWHGTAPDWFFNYYW Not Tested

Rhodanodin Rhodanobacter thiooxydans LCS2

GVLPIGNEFMGHAATPG Not Tested

Rubrivinodin Rubrivivax gelatinosus GAPSLINSEDNPAFPQRV Not Tested

Snou-LP S. noursei ATCC 11455 YFGLTGYENLFHFYDKLH Not Tested

Sphaericin Planomonospora

sphaerica GLPIGWWIERPSGWYFPI Yes

Sphingonodin I Sphingobium japonicum GPGGITGDVGLGENNFG Not Tested

Sphingonodin II Sphingobium japonicum GMGSGSTDQNGQPKNLIGG Not Tested

Sphingopyxin I Sphingopyxis alaskensis RB2256

GIEPLGPVDEDQGEHYLFAGG Not Tested

Sphingopyxin II Sphingopyxis alaskensis RB2256

GEALIDQDVGGGRQQFLTG Not Tested

SRO15 2005 Streptomyces

roseosporus GYFVGSYKEYWSRRII Not Tested

Streptomonomicin Streptomonospora alba SLGSSPYNDILGYPALIVIYP Yes

Subteresin Sphingomonas

subterranea GPPGDRIEFGVLAQLPG No

Sungsanpin Streptomyces sp. SNJ013

GFGSKPIDSFGLSWL Not Tested

Syanodin-I Sphingobium yanoikuyae XLDN2-5

GISGGTVDAPAGQGLAG Not Tested

Xanthomonin I Xanthomonas gardneri GGPLAGEEIGGFNVPG No

Page 26: Heterologous expression and characterization of the

17

Xanthomonin II Xanthomonas gardneri GGPLAGEEMGGITT No

Xanthomonin III Xanthomonas gardneri GGAGAGEVNGMSP No

Ulleungdin Streptomyces sp. KCB13F003

GFIGWGKDIFGHYGG Not Tested

Zucinodin Phenylobacterium zucineum HLK1

GGIGGDFEDLNKPFDV Not Tested

9810-LP Streptomyces sp. ADI94-01

GYFVGSYKEYWTRRIV Not Tested

Class III

BI-32169 Streptomyces sp. DSM 14996

GLPWGCPSDIPGWNTPWAC Not Tested

9401-LP1 Streptomyces sp. ADI94-01

AFGPCVENDWFAGTAWIC Not Tested

Class IV

LP2006 Nocardiopsis alba GRPNQGFENDWSCVRVC Yes

1.1.3.2 Biosynthesis of lasso peptides

Biosynthesis of lasso peptides requires a minimum of two enzymes: a lasso peptidase and a lasso

cyclase (Figure 1.1)76. The lasso peptidase is a cysteine protease which cleaves the leader

sequence from the core sequence in the lasso precursor peptide. Some lasso peptides may require

ATP in a pre-folding step prior to leader peptide removal and core cyclization, although this does

not appear to be a universal requirement77,78. Yan et al. have found that in vitro, the lasso

peptidase of microcin J25 requires the cyclase for proteolysis, likely forming a complex with the

cyclase77. Although, leader peptide proteolysis does not require the presence of the cyclase

protein in the case of fusilassin/fuscanodin78.

Page 27: Heterologous expression and characterization of the

18

Figure 1.10. Lasso peptide gene clusters and biosynthesis. A. Generalized lasso peptide gene

cluster. Although all lasso peptide clusters have the genes ACEB, they are not necessarily

arranged in order depicted above. The E and B genes are sometimes fused into a single B gene.

B. General mechanism of lasso peptide biosynthesis.

Ubiquitously found in lasso peptide clusters and widespread in other RiPP gene clusters is the

RiPP recognition element (RRE). The RRE, which is present is roughly 50% of all RiPP clusters,

recognizes the leader sequence of the precursor peptide and enables lasso peptide proteolysis and

cyclization to occur78,79. The RRE forms a winged helix-turn-helix structure and interacts with

the leader peptidase through electrostatic interactions, some of which are under co-evolutionary

pressure78. In some gene clusters, including the microcin J25 gene cluster, the RRE is found

fused to the N-terminus of the lasso peptidase. Recently, Sumida et al. crystallized the

fusilassin/fuscanodin leader peptide with its cognate RRE80. They found that the RRE binds very

tightly to the leader peptide with a dissociation constant of 6 nM, and that conserved residues

play critical roles in the recognition of the leader sequence by the RRE. Specifically, the

conserved YxxP motif and a conserved leucine fit into a hydrophobic cleft formed by the RRE.

Page 28: Heterologous expression and characterization of the

19

There is also evidence of coevolution between residues of the leader sequence and the RRE. The

conserved Tyr of the YxxP motif, as well as the conserved leucine are two of several residues

found to be under strong coevolutionary pressure78.

After the leader sequence is cleaved from the core sequence by the lasso peptidase, the second

lasso peptide biosynthesis enzyme, the lasso cyclase, catalyzes the cyclization of the peptide,

producing the mature lasso peptide. The lasso cyclase, which is homologous the aspartate-

dependent asparagine synthase AsnB, catalyzes the bond formation between the N-terminal

amine of the core peptide with a carboxylate macrolactam acceptor, either glutamate or aspartate,

ultimately producing a macrocycle. The Glu or Asp is located in the 8th or 9th and occasionally

the 7th position in the core sequence of the precursor peptide. The cyclase catalyzes macrolactam

formation presumably through adenylation of the carboxylate, thereby activating the carboxylate

for nucleophilic attack by the free N-terminal amine (Figure 1.11).

Figure 1.11. Proposed mechanism of microcin J25 biosynthesis. Although microcin J25

requires ATP for proteolysis, another lasso peptide, fusilassin/fuscanodin does not. Figure

adapted from Ortega and van der Donk, 201652.

In addition, lasso peptide biosynthetic gene clusters often contain ABC transporters. In microcin

J25, the dedicated ABC transporter, McjD, provides immunity to strains harbouring the microcin

biosynthesis genes81. Other post-translational modifications have been found to occur on some

lasso peptides, including phosphorylation, acetylation and methylation82–84.

The lasso peptide core sequence appears to be highly tolerant of substitutions. In a study

analyzing more than 380 single substitutions in the microcin J25 core sequence, Pavlova and

colleagues found that only three positions in the microcin J25 core sequence compromised

peptide stability85. Substitutions in most positions did not completely compromise RNA

polymerase inhibitory activity, demonstrating that lasso peptides are highly amenable to

engineering.

Page 29: Heterologous expression and characterization of the

20

1.1.3.3 Activity of lasso peptides

The lasso peptides display diverse activities including antibacterial, anti-HIV and inhibition of

the glucagon receptor, endothelin B receptor, and cancer cell invasion86–89. Although diverse

activities have been reported for the lasso peptides, antibacterial activity is the most common. Of

the 69 studied lasso peptides, 21 have reported antibacterial activity and the targets of 8 of these

antibacterial lasso peptides have been investigated. Among these 8 lasso peptides, there are three

common targets: RNA polymerase, ClpC1P1P2 protease and cell wall/lipid II (Figure 1.12).

Figure 1.12. Antibacterial targets of lasso peptides. Figure from Tan et al., 20197.

1.1.3.3.1 Cell wall biosynthesis inhibitors

Siamycin-I, which is a 21-residue, class I lasso peptide produced by certain Streptomyces strains,

has been found to inhibit cell wall biosynthesis by targeting lipid II. It has antimicrobial activity

against many Gram-positive bacteria but not Gram-negatives90,91. Its antimicrobial activity

Page 30: Heterologous expression and characterization of the

21

includes activity against pathogens such as vancomycin-resistant enterococci (VRE) and

methicillin-resistant Staphylococcus aureus, against both of which the minimum inhibitory

concentrations (MIC) of siamycin-I is 7μM. Daniel-Ivad et al. found that siamycin-I activates the

lia (lipid II-interfering antibiotics) promotor, suggesting that the antibiotic acts at the cell wall.

Further work by Tan et al., revealed that mutations in S. aureus that confer resistance to

siamycin-I were found in the walK/R genes92. The WalK/R two-component system is highly

conserved and involved in the regulation of cell metabolism93. Interestingly, the walK/R mutants

exhibited thickened peptidoglycan when compared to the wild type S. aureus. Further in vitro

enzymatic inhibition studies found that siamycin-I interacted with lipid II to prevent the

transglycosylation reaction catalyzed by penicillin-binding protein from occurring. Siamycin-I

displays some distinct differences from other lipid II-interfering antibiotics, including a specific

localization at the division septum of S. aureus and Bacillus subtilis and not resulting in the

accumulation of the of cytoplasmic peptidoglycan precursor, UDP-MurNAc-pentapeptide94.

Siamycin-I and its analogues have also been found to inhibit the entry of HIV into host cells

through binding to CD4, CCR5 and CXCR4 and to block fsr quorum sensing in Enterococcus

faecalis86,95,96.

The only other lasso peptide that has been found to target peptidoglycan synthesis is

streptomonomicin, which is produced by Streptomonospora alba and is a member of the class II

peptides that lack disulfide bridges97. Streptomonomicin has potent antibacterial activity against

various strains of the genus Bacillus including B. anthracis, the causative agent of anthrax,

against which the MIC is 2-4 μM. Similar to siamycin-I-resistant S. aureus, streptomonomicin-

resistant clones of B. subtilis were found to have mutations in the walR gene, including some

mutations that were identical between the resistant strains. This suggests that streptomonomicin

also targets the cell wall, although further experiments are required to determine its specific

molecular target.

1.1.3.3.2 RNA polymerase inhibitors

In contrast to siamycin-I and streptomonomicin, there are several lasso peptides produced by

Proteobacteria which target RNA polymerase in Gram-negative bacteria. These lasso peptides,

Page 31: Heterologous expression and characterization of the

22

which all lack disulfide bonds and therefore all belong to class II, include microcin J25, the first

lasso peptide that was studied in-depth, capistruin, acinetodin, klebsidin and citrocin.

Microcin J25, a 21 amino acid plasmid-encoded peptide, was first discovered in 1992 by

Salomon and Farías73. Microcin J25 is produced by certain strains of Escherichia coli, and is

active against E. coli and Salmonella Newport at concentrations as low as 10 and 5 nM,

respectively. Initially, the mechanism of microcin J25 action was thought to be related to the

function of the cell envelope proteins fhuA, tonB, exbB and sbmA98,99. Later it was found that

mutations in the rpoC gene (which encodes the β’ subunit of RNA polymerase) confers microcin

resistance in E. coli100. RNA synthesis is impaired both in vitro and in vivo in E. coli treated with

microcin J25 and it is thought that the filamentous phenotype of microcin J25-treated E. coli may

result from impaired transcription of genes encoding cell division proteins.

It was then discovered that mutations in rpoB (encoding the β subunit of RNA polymerase) that

confer resistance to streptolydigin result in cross-resistance to microcin J25, suggesting a shared

mechanism of action101. Mutations conferring resistance to microcin J25 in both rpoC and rpoB

map to the secondary channel, which is thought to be important for nucleotide substrate access to

the active site as well as accepting the 3’ end of nascent RNA in backtracked elongation

complexes102. Indeed, microcin J25 binding completely obstructs the secondary channel,

inhibiting both the forward reaction of phosphodiester bond formation, and the reverse reaction

of pyrophosphorolysis103,104. A recent crystal structure has validated this view, demonstrating

that microcin binds deep within the RNA polymerase secondary channel, constricting the solvent

accessible channel to less than 5 Å105. Further, microcin J25 blocks folding of an important

mobile structural element of the β’ subunit called the trigger loop.

Much of the work understanding the mechanism of action of microcin J25 was performed prior

to a published structure that was correct. In fact, the first three published structures of microcin

J25 were later determined to be incorrect106–108. These structures proposed that microcin J25 was

a 21 amino acid peptide cyclized between its N and C terminus, as opposed to the lasso-like

structure known today109.

Another class II lasso peptide that targets RNA polymerase is capistruin. Capistruin is a 19

amino acid peptide, discovered in 2008 that is produced by Burkholderia thailandensis E264110.

Page 32: Heterologous expression and characterization of the

23

It has antibacterial activity against certain strains of Burkholderia and E. coli, with MICs as low

as 12 and 25 μM, respectively. In 2011, Kuznedelov et al. showed that microcin J25-resistant

RNA polymerase is cross-resistant to capistruin, and that capistruin inhibits RNA polymerase-

dependent transcript elongation in vitro111. These results strongly suggested that capistruin and

microcin J25 have a common mechanism of action. Later, Braffman and colleagues determined

the structure of capistruin-bound RNA polymerase and found that the binding sites of capistruin

and microcin J25 largely overlap105. Among the differences, capistruin binds farther (12 Å) from

the RNA polymerase active site than does microcin J25 (6.5 Å) and as a result, capistruin does

not appear to restrict access of nucleotide substrates to the active site. Additionally, while

microcin competitively inhibits binding of nucleotides in the active site, capistruin binds too far

from the active site to be competitive with respect to nucleotide binding.

Acinetodin and klebsidin are two lasso peptide RNA polymerase inhibitors, produced by

Acinetobacter gyllenbergii CIP 110306 and Klebsiella pneumoniae 4541–2, respectively112.

Acinetodin and klebsidin were identified through a cysteine protease-guided genome mining

approach and heterologously expressed in E. coli. Acinetodin does not have antibacterial activity

against Gram-positive or Gram-negative bacteria, while klebsidin has very weak antibacterial

activity against certain strains of the genus Klebsiella.

Although acinetodin and klebsidin do not have strong antibacterial activity, in vitro testing

revealed that they are both inhibitors of RNA polymerase elongation112. Microcin J25, which

was used as a positive control, was the most potent RNA polymerase inhibitor, followed by

klebsidin, then finally acinetodin, which was significantly less potent. Further, a mutation in the

rpoC subunit of RNA polymerase that confers resistance to microcin J25 was found to result in

cross-resistance to acinetodin and klebsidin, indicating that similar to microcin J25 and

capistruin, acinetodin and klebsidin likely target the RNA polymerase secondary channel. A

crystal structure of acinetodin and klebsidin with RNA polymerase would reveal the differences

between the binding modes of acinetodin, klebsidin, capistruin and microcin J25.

Recently, a 19 amino acid class II lasso peptide from Citrobacter pasteurii and Citrobacter

braakii was characterized and also found to be an RNA polymerase inhibitor113. Citrocin has

antimicrobial activity against several Gram-positive strains of bacteria, with MICs as low as 16

Page 33: Heterologous expression and characterization of the

24

μM. In spite of its weaker antimicrobial activity when compared to microcin J25, citrocin is a

more potent inhibitor of RNA polymerase. In fact, 100 μM of microcin J25 was required to

achieve the same level of RNA polymerase inhibition as 1 μM of citrocin. This suggests that

citrocin uptake is the factor limiting its antimicrobial activity.

1.1.3.3.3 ClpC1P1P2 protease inhibitors

Lassomycin, a class II lasso peptide, was identified in a screen of actinomycete extracts for those

that specifically inhibited the growth of Mycobacterium tuberculosis. Lassomycin, which is

produced by a Lentzea kentuckyensis sp, specifically inhibits the C1 subunit of the ClpC1P1P2

protease complex. Lassomycin potently inhibits strains of mycobacteria with MICs lower than 1

μM and is even able to inhibit the growth of multidrug-resistant M. tuberculosis. Lassomycin had

weaker activity against other Actinobacteria such as Propionibacterium (Cutibacterium) acnes,

and no activity against other Gram-positive or Gram-negative bacteria. Lassomycin-resistant

mutants of M. tuberculosis suggested that the target of lassomycin was the ClpC1 ATPase

subunit of the ClpC1P1P2, as all of the mutations mapped to this region. In vitro enzymatic work

revealed that lassomycin increased the ATPase activity of ClpC1, while decreasing the ability of

ClpC1P1P2 to perform proteolysis. This ability of lassomycin to decouple the ATP activity of

ClpC1 from proteolysis, represents a new antibiotic mechanism of action. Interestingly,

lassomycin was highly specific to the ClpC1P1P2 protease, and did not affect ATP hydrolysis in

any other AAA ATPases, including the E. coli ClpC1 homolog, ClpA.

Unlike nearly all other lasso peptides, the C-terminal tail of lassomycin is not threaded through

the N-terminal ring which, in contrast to threaded lasso peptides, allows for the chemical

synthesis of the peptide114. As a result of its potent anti-mycobacterial activity against resistant

strains and the ease with which it is able to be synthesized, lassomycin offers great therapeutic

promise115.

1.1.3.4 Discovery and Heterologous expression of lasso peptides

An advantageous feature of the RiPPs is their suitability for genome mining and heterologous

expression. The lasso peptide characterization in particular has benefitted from genome mining

and heterologous expression tools. Several groups have successfully characterized novel lasso

peptides, which they predicted to have unique features based on genomic data75,110,116. Currently,

Page 34: Heterologous expression and characterization of the

25

there are six genome mining tools that are able to detect lasso peptide biosynthetic gene

clusters46,75,117–120. Perhaps the most commonly used genome mining tool for secondary

metabolite genome mining is AntiSMASH, which has integrated the RODEO lasso peptide

detection algorithm into its search function in its fourth and fifth releases.

Many of the lasso peptides that have been studied through heterologous expression, although the

study of actinobacterial lasso peptides has lagged behind the study of those from Proteobacteria

(Table 1.1). The most popular heterologous host for lasso peptide expression remains E. coli,

although several have been expressed in Streptomyces sp. and one in Sphingomonas

subterranean. To date, only two lasso peptides from Actinobacteria have been expressed in E.

coli: chaxapeptin and fuscanodin/fusilassin, which has been successfully produced in E. coli by

two separate groups. At 0.1 mg/L, the yield of heterologously-expressed chaxapeptin was much

lower than the 0.7 mg/L yield of the natively-expressed peptide121.

Table 1.2. Heterologously produced lasso peptides.

Peptide Name Native Organism Heterologous Host Organism

Class I

Siamycin-3 Streptomyces sp. S. coelicolor M1152

SSV 2083/Sviceucin Streptomyces sviceus S. coelicolor M1146

Class II

Albusnodin Streptomyces albus S. coelicolor M1146 and S. lividans 66

Astexin 1-3 Asticcacaulis excentricus E. Coli BL21

Benenodin-1 Asticcacaulis benevestitus E. Coli BL21

Brevunsin Brevundimonas diminuta Sphingomonas subterranea

Burhizin Burkholderia rhizoxinica E. Coli BL21

Capistruin Burkholderia thailandensis E. Coli BL21

Caulonodulin 1-7 Caulobacter sp. E. Coli BL21

Caulosegnin 1-3 Caulobacter segnis E. Coli BL21

Chaxapeptin Streptomyces leeuwenhoekii C58 E. Coli BL21

Citrocin Citrobacter pasteurii E. Coli BL21

Citrulassin Streptomyces albulus S. lividans 66

Fusilassin/Fuscanodin Thermobifida fusca E. Coli BL21

Page 35: Heterologous expression and characterization of the

26

Klebsidin K. pneumoniae E. Coli BW25113

Microcin J25 E. coli AY25 E. Coli XL-1 Blue

Paeninodin Paenibacillus dendritiformis C454 E. Coli BL21

Rhodanodin Rhodanobacter thiooxydans LCS2 E. Coli BL21

Rubrivinodin Rubrivivax gelatinosus E. Coli BL21

Snou-LP Streptomyces noursei ATCC 11455 S. lividans TK24

Sphingonodin I-II Sphingobium japonicum E. Coli BL21

Sphingopyxin I-II Sphingopyxis alaskensis RB2256 E. Coli BL21

Syanodin-I Sphingobium yanoikuyae XLDN2-5 E. Coli BL21

Xanthomonin I-III Xanthomonas gardneri E. Coli BL21

Zucinodin Phenylobacterium zucineum HLK1 E. Coli BL21

9810-LP Streptomyces sp. ADI94-01 Streptomyces albus J1074

Class III

9401-LP1 Streptomyces sp. ADI94-01 S. albus J1074

1.2 Aim of this work

The aim of this thesis is to expand scientific knowledge about an emerging class of antibiotics,

the lasso peptides. Specifically, the aim of this project is to investigate a unique and

uncharacterized lasso peptide, LP2006. LP2006 is a structurally unique antibacterial lasso

peptide, which may indicate that it has a novel mechanism of action with respect to other

antibacterial lasso peptides. To facilitate the study of LP2006 and other lasso peptides, I intend

develop a system for the production of lasso peptides by heterologous expression. I attempt to

heterologously produce LP2006 in E. coli and several strains of Streptomyces, and successfully

detect the production of LP2006 in a host strain of S. coelicolor M1146.

Page 36: Heterologous expression and characterization of the

27

Chapter 2 Materials and Methods

Materials and Methods

2.1 General experimental procedures

2.1.1 Materials

Unless otherwise stated, all chemicals were purchased from Sigma-Aldrich. Difco Marine Broth

2216 Agar was purchased from BD Biosciences. Bacto Agar, Bacto Malt Extract and Bacto

Yeast Extract were purchased from BD Biosciences.

2.1.2 Strains and plasmids used

Table 2.1. Strains and plasmids used in this study.

Strain/Plasmid Description Source

Environmental strains

Maritime Strains (Table 5.1) Isolates from the maritime provinces of eastern Canada

Nocardiopsis sp. HB141 (=H153) Environmental isolates from the sponge Halichondria panacea (harvested near Kiel, Germany)

(Schneemann et al., 2010)122

Cloning and expression strains

Streptomyces albus J1074 Used commonly for heterologous expression

Streptomyces avermitalis SUKA22 S. avermitalis containing a large, systematic deletions of nonessential genes

(Komatsu et al.,2013)123

Streptomyces coelicolor M1146 S. coelicolor M145, ∆cda ∆red ∆act ∆cpk (Gomes-Escribano and Bibb, 2011)124

Streptomyces coelicolor M1154 S. coelicolor M145, ∆cda ∆red ∆act ∆cpk + rpsL and rpoB mutations

Escherichia coli ET12567(pUZ8002) Methylation-deficient conjugal donor strain (dam-13::Tn9 dcm-6 hsdM Cmr)

Escherichia coli TOP10 General cloning host

Escherichia coli BL21(DE3) Protein expression host

Antimicrobial Testing strains

Bacillus subtilis 168 YB5018 dinC18::Tn917Iac metB5 trpC2 xin-1 SPβ- amyE+

(Jani et al., 2015)125

Bacillus subtilis 168 1A980 Em trpC2 liaI::pMUTIN attSPβ Bacillus Genetic Spore Center

Bacillus subtilis JH642 Common laboratory strain

Enterococcus ATCC 51299 Vancomycin-resistant ATCC

Page 37: Heterologous expression and characterization of the

28

Enterococcus faecalis ATCC 29212 Vancomycin-sensitive ATCC

Escherichia coli BW25113 Common laboratory strain

Escherichia coli BW25113 ∆tolC ∆bamB Hyperpermeable strain of E. coli

Micrococcus luteus Antibacterial test strain

Mycobacterium smegmatis Antibacterial test strain

Saccharomyces cerevisiae Y7092 Anti-yeast test strain

Staphylococcus aureus ATCC BAA-41 Methicillin-resistant ATCC

Staphylococcus aureus ATCC 29213 Methicillin-sensitive ATCC

Staphylococcus epidermis ATCC 14990 Antibacterial test strain ATCC

Plasmids

pACYCDuet-1 Inducible co-expression plasmid Novagen

pRSFDuet-1 Inducible co-expression plasmid Novagen

pSET152-ermE*p Streptomyces overexpression plasmid Bibb et al., 1985126

pGEM T-Easy Cloning vector Promega

2.1.3 Primers used

Table 2.2. Primers used in this study.

Primer

Number Primer Name Sequence (5’ to 3’)

1 27F_16S_universal_primer GAGTTTGATCCTGGCTCA

2 1492R_16S_universal_primer TACGGCTACCTTGTTACGACTT

3 LpeA_For GCGCACACCATGGACGAAGAAAAGATCGGC

4 LpeA_Rev TAATTCAAGCTTTCAGCAGACCCGGACGCAGGA

5 LpeC_For GCGCACCATATGAAGTTCATCGTTCTTCCC

6 LpeC_Rev GACCACGATATCTCAGCCTTCCAGGATCGATAC

7 LpeE_For GCGCACACCATGGAATTCATCGACGACACG

8 LpeE_Rev TAATTCAAGCTTTCATCGCCGCAGCACCCCCAC

9 LpeB_For TAATTCCATATGACGGTACCCGTGGCCCTC

10 LpeB_Rev TAGCTCGATATCTCAGTCATCATCACGGACCAC

11 ACYCDuetUP1 GGATCTCGACGCTCTCCCT

12 DuetDOWN1 GATTATGCGGCCGTGTACAA

13 DuetUP2 TTGTACACGGCCGCATAATC

14 T7_Terminator GCTAGTTATTGCTCAGCGG

15 LpeACEBDD_For GCCGGTTGGTAGGATCAGGAGGATATCATATGGACGAAGAAAAGATCGGC

16 LpeACEBDD_Rev AAGCTTGGGCTGCAGGTCGACTCTAGACGGTCTCGCGACGCAGGTGGT

17 pSET152_ermE*_For GCTCACTCATTAGGCACCCCAGGC

18 pSET152_ermE*_Rev AGGGGGATGTGCTGCAAGGCG

Page 38: Heterologous expression and characterization of the

29

2.1.4 Culture conditions

Nocardiopsis sp. HB141 was grown on either GYM medium (per L Milli-Q water: 4g yeast

extract, 10g malt extract, 4g dextrose), SMMS medium (per L Milli-Q water: 2g Difco casamino

acids, 5.3g TES buffer), ATCC medium 172 (per L Milli-Q water: 10g glucose, 20g soluble

starch, 5g yeast extract, 5g Bacto Peptone, 1g CaCO3), YEME medium (per L Milli-Q water:

10g glucose, 170g sucrose, 3g yeast extract, 5g Bacto Peptone, 3g malt extract) or TSB medium

(per L Milli-Q water: 2.5g glucose, 17g tryptone, 3g soytone, 5g NaCl, 2.5g K2HPO4). 16g per L

of Bacto agar was added to each media if solid growth medium was desired. Nocardiopsis sp.

HB141 was grown for 3 weeks at 30°C and extracted with methanol.

All strains used for antimicrobial testing, DNA manipulation and E. coli used for heterologous

expression were grown using LB medium (per L Milli-Q water: 10g tryptone, 10g NaCl, 5g

yeast extract) and grown at 37°C. All Streptomyces strains used for heterologous expression were

cultured on MYM medium (per L Milli-Q water: 4g yeast extract, 10g malt extract, 4g maltose,

16g agar) at 30°C unless otherwise indicated.

2.1.5 Heterologous expression of LP2006

For the heterologous expression of LP2006 in E. coli, the vectors pACYCDuet-1 and pRSFDuet-

1 were used. The genes lpeACEB were amplified from genomic DNA extracted from

Nocardiopsis sp. HB141 using the primers 3-8 (Table 2.2). A two-step reaction using Q5 DNA

polymerase was used for all PCR reactions unless otherwise specified due to the high GC content

of the template DNA and the high annealing temperature of the primers. The lpeAC gene

fragments were cloned into the pACYCDuet-1 vector, while the lpeEB gene fragments were

cloned into the pRSFDuet-1 vector, resulting in the vectors pACYCDuet-1-lpeAC and

pRSFDuet-1-lpeEB. The correct gene sequences were confirmed by Sanger sequencing with the

primers 11-14.

E. coli BL21(DE3), co-transformed with pACYCDuet-1-lpeAC and pRSFDuet-1-lpeEB was

grown at 37°C in 1L LB medium, induced with IPTG after reaching an OD of 0.4-0.8 and

returned to the shaking incubator for 4-16 hours at 20°C. Several IPTG concentrations were

tested, ranging from 50 μM to 500 μM. Cells were harvested, and the pellet was extracted with

methanol, sonicated and macerated overnight. The supernatant was combined with 20 g/L

Page 39: Heterologous expression and characterization of the

30

XAD16 resin (Sigma Aldrich), stirred overnight, then removed by filtration. The resin was

washed several times with ddH2O, then adsorbed molecules were eluted with 100 mL of

methanol.

For the heterologous expression of LP2006 in S. coelicolor, the integrative vector pSET152-

ermE*p was used. The entire 8-gene LP2006 gene cluster was amplified from the Nocardiopsis

sp. HB141 genomic DNA using the primers 15 and 16 and cloned into pSET152-ermE*p. The

sequence was confirmed using the primers 11-16 and 17-18 (Table 2.2). The vector pSET152-

ermE*p-lpeACEBDD was transformed into E. coli ET12567(pUZ8002) to use as a conjugal

donor. Conjugation to S. coelicolor M1146, M1154, S. avermitalis SUKA22 and S. albus J1074

was performed according to the protocol outlined in Kieser et al., 2000127. Exconjugants were re-

streaked on MYM containing apramycin twice and genomic DNA was extracted to confirm that

the insert was present without mutations. S. coelicolor M1146 pSET152-ermE*p-lpeACEBDD

was grown on MYM and GYM media for 5-7 days and the production of LP2006 was assessed

by LC-MS.

2.2 Isolation and purification of bioactive metabolites

2.2.1 Metabolite extraction

Nocardiopsis sp. HB141 metabolites were extracted from agar or liquid cultures with HPLC-

grade methanol, using a volume of methanol equivalent to the volume of the culture or 200-300

mL, whichever is less. Organic solvent was evaporated from the extracts using a Genevac EZ-2

Elite series evaporator (SP scientific).

2.2.2 Flash chromatography purification

Crude extracts were resuspended in 5% aqueous HPLC-grade methanol to a concentration of 100

mg/mL. After centrifugation and/or filtration, the resuspended crude extracts were further

purified on a Reveleris® X2 Flash chromatography system (Buchi Labortechnik). A 20g C18 40-

60um 100Å cartridge (Aegio Technologies) was used for the separation by flash

chromatography, with a linear gradient from 5 to 100% aqueous HPLC-grade methanol at a flow

rate of 10mL/min. 20mL fractions were collected, dried by Genevac or rotary evaporation and

Page 40: Heterologous expression and characterization of the

31

tested for bioactivity by broth microtiter dilution assay (5.2.1.3). Fractions that were bioactive

were pooled, resuspended and further purified by HPLC.

2.2.3 High-performance liquid chromatography purification

After flash chromatography, the Nocardiopsis sp. HB141 extract was further purified by HPLC

to study the activity of the bioactive metabolite(s). Individual and pooled bioactive flash

chromatography fractions were purified on a Waters Alliance HPLC with a Phenomonex Luna

C18 column (100 Å, 5 μm, 4.6x250mm). The Nocardiopsis sp. HB141 fractions from flash

chromatography were purified by HPLC over two steps. The first step involved the following 30-

minute linear solvent gradient of water/0.1% formic acid (solvent A) and acetonitrile/0.1%

formic acid was used: hold at 5% B for 2 minutes, followed by linear increase until 95% B at 20

minutes, hold at 95% B until 25 minutes, then return to 5% B until 30 minutes. The peak

containing LP2006 was collected and the fraction was dried by Genevac evaporator and

confirmed as LP2006 by mass spectrometry.

The second step involved the following 30-minute linear solvent gradient of water/0.1% formic

acid (solvent A) and acetonitrile/0.1% formic acid was used: hold at 20% B for 2 minutes,

followed by linear increase until 60% B at 20 minutes, hold at 60% B until 25 minutes, then

return to 20% B until 30 minutes. During both purification steps the column temperature was

35°C and the flow rate was 1 mL/min. Again, the LP2006-containing fraction was collected,

dried and confirmed by mass spectrometry.

2.2.4 Liquid chromatography mass spectrometry analysis

A Waters Xevo G2-S QTOF mass spectrometer with Acquity liquid chromatography was used to

analyze extract metabolites. An Acquity UPLC BEH C18 (1.7 μm 2.1*50mm) column was used

to achieve liquid chromatography separation of the sample with a column temperature of 40°C

and a flow rate of 0.2 mL/min. A 20-minute linear solvent gradient of water/0.1% formic acid

(solvent A) and acetonitrile/0.1% formic acid was used: linear increase from 5% to 95% B from

0-12 minutes, hold at 95% B until 17 minutes, return to 5% B at 18 minutes and hold at 5% B

until 20 minutes. Samples were analyzed in the positive mode, using the electrospray ionization

source.

Page 41: Heterologous expression and characterization of the

32

2.3 Susceptibility testing

2.3.1 Disk diffusion assays

The test organism was inoculated from an agar streak plate into a 5mL liquid culture of LB or

YPD media, depending on the organism. The cultures were grown in a shaking incubator

overnight at 30°C or 37°C, depending on the test organism. The next day, the cultures were

diluted 1:100 into fresh media and returned to the shaking incubator. Cultures were grown to an

OD600 of 0.4-0.6, diluted 1:1000 in LB or YPD media and inoculated onto agar plates of the

desired medium to produce a lawn of growth. 6 mm paper filter disks (BS Biosciences) were

placed on the agar plates and 2-10 uL of crude extract resuspended in DMSO were placed onto

the filter disks. The agar plates were incubated overnight at 30°C or 37°C depending on the test

organism and the plates were imaged in the morning and the zone of inhibition surrounding each

filter disk was observed.

2.3.2 Broth microtiter dilution assay

The test organism was inoculated from an agar streak plate into a 5mL liquid culture of YPD

medium for S. cerevisiae or LB medium for all other test organisms. The liquid culture was

grown overnight at 30°C (for S. cerevisiae) or at 37°C for (all other organisms) and a 1:100

subculture was started in the morning. Each culture was grown to an OD of 0.4-0.6 and diluted

1:1000 with fresh media in a 96-well plate. The test compound or extract, resuspended in

DMSO, was added to each well and the plate was incubated overnight. The next morning, the

OD600 of each well was measured. When testing extracts, the extract was considered active if

inhibition of growth was greater than 50%. For the determination of MICs, the same protocol

was followed and antibiotics at twofold increasing concentrations were added to the 96 well plate

liquid cultures.

2.4 Target identification

2.4.1 LacZ reporter assay

Disk diffusion assays were performed using two B. subtilis lacZ reporter strains with 8mg/mL X-

gal added to monitor either cell envelope stress of the induction of the SOS response. The B.

Page 42: Heterologous expression and characterization of the

33

subtilis reporter strain 1A980 (liaI-lacZ fusion) while the strain YB5018 was used to monitor the

DNA damage SOS response (dinC-lacZ fusion).

Results and discussion

3.1 Nocardiopsis sp. HB141 extract testing

3.1.1 Extracts of Nocardiopsis sp. HB141 have antibacterial activity

In a search for novel antimicrobial natural products, nearly 50 strains of marine bacteria were

screened against M. luteus, B. subtilis, E. coli, and S. cerevisiae. Each strain was grown on

several media types for 5 days, extracted with organic solvent and tested for antibacterial or anti-

yeast activity in a broth microtiter dilution assay (for details about the screen see section 5.3.1.2).

The extract was considered a hit if it produced a 50% reduction in the OD600 value of the

indicator organism as compared to the control. Among the hit strains was Nocardiopsis sp.

HB141, which was isolated from homogenates of the marine sponge Halichondria panacea

harvested in the Baltic Sea near Kiel, Germany122. Extracts of Nocardiopsis sp. HB141, grown

on GYM medium, have antibacterial activity against the Gram-positive bacteria M. luteus and B.

subtilis, as demonstrated by the growth inhibitory effect of the extracts in broth microtiter

dilution and disk diffusion assays (Figure 3.1).

Figure 3.1. Disk diffusion assay of Nocardiopsis sp. HB141 extract. DMSO or Nocardiopsis

sp. HB141 extract was added to the filter disk, which was placed on a lawn of bacteria grown on

an agar plate. The HB141 extract inhibits growth of B. subtilis and M. luteus, as indicated by the

zone of growth inhibition surrounding the filter disk demonstrating that has antibacterial activity

against these test organisms.

Page 43: Heterologous expression and characterization of the

34

3.1.2 Nocardiopsis sp. HB141 is a producer of an antibacterial lasso peptide, LP2006

To identify bioactive metabolites that were responsible for the antibacterial activity of the

Nocardiopsis sp. HB141 extract, I analyzed the extract by LC-MS. Metabolite profiling by LC-

MS revealed that the HB141 crude extract contained a mass that was not present in the media

control extract, which eluted at 5 minutes (Figure 3.2). Based on the averaged mass spectra of

the sample, the mass of the identified peak was 1002.9335 m/z and was identified as an M+2H

ion based on the isotope pattern.

Figure 3.2. Extracted ion chromatogram of LP2006 M+2H mass. LC-MS chromatogram of

the ion 1002.93 m/z, the most abundant ion of LP2006. The ion is present in the HB141 crude

extract, but absent in the media control extract.

Next, I used The Dictionary of Natural Products to determine if there is a known natural product

with this mass. I found that the mass and fragmentation pattern of the Nocardiopsis sp. HB141

metabolite matches the reported values of a recently discovered lasso peptide, LP200675 (Figure

3.3 and Table 1.1). LP2006 was identified by a genome-mining tool and it is the establishing and

only member of the class IV lasso peptides.

The error values on all detected ions were all lower than 2 ppm, which is well within the

instrument error of the mass spectrometer. Additionally, there is a close phylogenetic

relationship between Nocardiopsis sp. HB141 and the strain from which LP2006 was first

discovered, Nocardiopsis alba NRRL B-24146. Since there is often a close association between

Page 44: Heterologous expression and characterization of the

35

bacterial chemotype and phylotype, this provides more evidence that Nocardiopsis sp. HB141 is

a producer of LP2006128. Collectively, the phylogenetic data and the mass spectra strongly

suggest that Nocardiopsis sp. HB141 is a producer of LP2006.

Figure 3.3. Mass spectrum of the Nocardiopsis sp. HB141 mass of 1002.9335 m/z. The amino

acid sequence of LP2006 is overlaid on the spectrum and the ions are assigned based on the

values reported for LP2006 by Tietz et al., 201775.

Table 3.1. Ion comparison of HB141 mass and LP2006

Ion Calculated

Mass (Da)

Observed

Mass (Da) Error (ppm)

y6+ 664.2905 664.2905 0

y7+ 850.3698 850.3683 -1.7

[M+2H]2+ 1002.9338 1002.9335 -0.3

b10+ 1155.4966 1155.4951 -1.3

3.1.3 Purification of LP2006

To study the activity of LP2006, I decided to purify the peptide. I started with purification by

flash chromatography, a preparative method for the rapid purification of crude extracts. After

purification by flash chromatography, only two fractions retained antibacterial activity, fractions

6 and 7 (Figure 3.4). I found that fractions 6 and 7 contained the vast majority of the eluted

Page 45: Heterologous expression and characterization of the

36

LP2006, with other fractions containing only trace amounts. These results are consistent with the

hypothesis that LP2006 is responsible for the antibacterial activity of the Nocardiopsis sp.

HB141 crude extract.

Figure 3.4. Disk diffusion assay from flash chromatography fractionated Nocardiopsis sp.

HB141 extract. The flash chromatography fractions of the purified Nocardiopsis sp. HB141

extract are added to a filter disk, which is placed on a lawn of M. luteus or B. subtilis. Fractions 6

and 7 were able to inhibit growth of M. luteus or B. subtilis, while the remaining fractions and

the DMSO control were not able to.

LP2006 was further purified in a two-step HPLC purification to remove co-eluting compounds

(Figure 3.5 and Figure 3.6). The purity of the LP2006 was confirmed by LC-MS and HPLC.

After the three-step purification, the yield of pure LP2006 was less than 100 μg per L of GYM

agar. Nocardiopsis sp. HB141 requires ~3 weeks to reach the sporulation stage of its growth on

GYM agar medium, which is substantially longer than the less than 1 week typical for the

secondary metabolite-rich Streptomyces to sporulate on solid media. As a result, the purification

of LP2006 from Nocardiopsis sp. HB141 is time and labour-intensive. Additionally,

Nocardiopsis sp. HB141 does not grow in the liquid growth media GYM, MYM, or YEME and

only grows on the solid growth media, GYM, MYM or ATCC172 agar. To combat such

challenges, I sought to produce LP2006 through another means. Chemical synthesis has not been

demonstrated for the lasso peptides (with the exception of lassomycin), thus I decided to pursue

the heterologous expression of LP2006. Lasso peptides are generally highly amenable to

heterologous expression, owing to the fact that their biosynthesis requires very few genes

(1.1.3.4).

Page 46: Heterologous expression and characterization of the

37

Figure 3.5. HPLC chromatogram of first-round purification of LP2006. The major peaks

were collected and analyzed by mass spectrometry to determine which peak corresponded to

LP2006. The LP2006 peak was then further purified.

Figure 3.6. HPLC chromatogram of second-round purification of LP2006. The major peaks

were collected and analyzed by mass spectrometry to determine which peak corresponded to

LP2006.

Page 47: Heterologous expression and characterization of the

38

3.2 Heterologous expression of LP2006

3.2.1 Heterologous expression in Escherichia coli

Since Nocardiopsis sp. HB141 is a slow-growing organism and the yields of LP2006 are very

low, I decided to heterologously express the peptide in a host strain. I decided to use E. coli to

heterologously express LP2006 as there has been much success expressing lasso peptides using

E. coli, although, at the time, no other labs had been able to heterologously express lasso

peptides from Actinobacteria in E. coli (Table 1.2).

As LP2006 was originally identified from a strain of Nocardiopsis alba, I searched the NCBI to

see if there were any genome sequences of N. alba available. Indeed, there are three N. alba

published, each of which contains a gene cluster capable of producing LP2006, as verified by

AntiSMASH. The LP2006 biosynthetic gene clusters of each of these three strains have >98%

sequence identity, with no differences in the sequence of the core peptide (Table 3.2). Based on

the high sequence identity between the LP2006 gene clusters and the taxonomic similarity

between N. alba DSM 43377 and Nocardiopsis sp. HB141, I reasoned that the LP2006 gene

cluster of Nocardiopsis sp. HB141 is likely highly similar to those in published genomes

sequences and to be amenable to amplification with primers based on published genome

sequences. Therefore, I designed primers based on the LP2006 cluster of N. alba DSM 43377,

with the primers designed to separately amplify the lpeACEB genes in the cluster (Table 2.2).

The lpeACEB genes are the core genes responsible for biosynthesis, while there are two

transporter genes in the cluster that I did not amplify and two genes flanking the ABC

transporters that are unrelated to the biosynthesis of LP2006 (Figure 3.7). The lpeACEB genes

were cloned into the protein expression vectors pACYCDuet-1 and pRSFDuet-1, with each gene

under the control of its own IPTG-inducible promotor (Figure 3.8).

Page 48: Heterologous expression and characterization of the

39

Table 3.2. Sequence identity of LP2006 biosynthetic proteins of Nocardiopsis alba ATCC

BAA-2165 and Nocardiopsis sp. TP-A0876 compared to Nocardiopsis alba DSM 43377.

Strain LpeA

(Precursor)

LpeC

(Lasso

cyclase)

LpeE (RiPP

recognition

element)

LpeB

(Leader

peptidase)

LpeD1

(ABC

transporter)

LpeD2

(ABC

transporter)

Nocardiopsis alba ATCC BAA-2165

100.0 98.2 100.0 100.0 99.0 99.7

Nocardiopsis sp. TP-A0876

100.0 99.7 100.0 100.0 100.0 98.6

Figure 3.7. LP2006 biosynthetic gene cluster from N. alba DSM 43377. The cluster contains 6

genes important for the biosynthesis of LP2006, lpeACEB, two ABC transporters for export of

LP2006, and two genes flanking the transporters which are unrelated to the biosynthesis of

LP2006.

Page 49: Heterologous expression and characterization of the

40

Figure 3.8. Vectors used for E. coli heterologous expression. The lpeACEB genes were

amplified from Nocardiopsis sp. HB141 using primers designed from the LP2006 gene cluster

from N. alba DSM 43377.

Heterologously expressed LP2006 in E. coli was undetectable by LC-MS under any tested

condition. The conditions manipulated included changing induction time, induction temperature,

inducer (IPTG) concentration. Other groups have been unable to express or have had incredibly

low yields expressing actinobacterial lasso peptides in E. coli, although it is not clear why this is

the case. In fact, only two lasso peptides from Actinobacteria have been expressed E. coli,

chaxapeptin and fusilassin/fuscanodin, although others have tried unsuccessfully83. It is likely

that additional failed heterologous expression attempts have gone unreported in the literature.

Chaxapeptin was the first successful expression of an actinobacterial lasso peptide in E. coli,

although the yield of the expression was 0.1 mg/L, which is lower than the yield of the natively-

produced chaxapeptin at 0.7 mg/L121. Martin-Gomez and colleagues used an approach where the

four chaxapeptin biosynthesis genes, cptACEB, were cloned into a single vector for expression.

cptA was placed under the inducible T7 promotor, while cptCEB were placed under the promotor

from the microcin J25 biosynthesis cluster. Following chaxapeptin, two groups were able to

independently produce the lasso peptide fusilassin/fuscanodin. The first group, who did not

report their final yield, took an approach of expressing the five biosynthesis genes tfuACEBD on

a single plasmid129. The second group took a similar approach used in this study whereby tfuA

was cloned into the pET28 vector and the tfuCEB genes were cloned into the pACYC vector. A

Page 50: Heterologous expression and characterization of the

41

numerical value is not reported for the final yield by this method, although the yield is reported

to be “low”78.

It is possible that the transporter genes may be required for proper biosynthesis of LP2006. In

their heterologous expression of fusilassin/fuscanodin, Koos and Link included the transporter

gene in construct for heterologous expression, although it is not clear if the gene is required for

biosynthesis. Other groups have been able to heterologously express lasso peptides without the

transporter gene(s) present on the expression plasmid. Additionally, the ABC transporters were

not required during in vitro synthesis of the lasso peptides, microcin J25, paeninodin and

fusilassin/fuscanodin76–78,129,130. When recombinantly expressing proteins in E. coli, an approach

often taken to overcome low or no expression is to correct for codon bias in the recombinant

gene131. Koos and Link codon optimized the tfuE and tfuB genes for the in vitro biosynthesis of

fusilassin/fuscanodin129. In other classes of RiPPs, the biosynthetic genes have been codon

optimized for heterologous expression, although the optimization has not always resulted in

higher yields132,133.

I decided to investigate the production of the LP2006 biosynthesis enzymes by SDS-PAGE. As

no antibodies are available against LpeACEB, I relied simply on the observance of induced

bands in crude cell lysates. I compared the cell lysates of E. coli transformed with pACYC-

lpeAC and pRSF-lpeEB and E. coli transformed with the pACYC and pRSF empty vectors

(Figure 3.9). A single, prominent band is observed in +lpeACEB lanes at the expected molecular

weight for LpeC (68 kDa), becoming more prominent in the three hours following induction. It is

likely that this band corresponds to LpeC, although analysis by western blot or mass

spectrometry is required for confirmation. Despite my efforts, the remaining proteins (LpeA: 2

kDa, LpeE: 10 kDa and LpeB: 16 kDa) were not detectable by SDS-PAGE.

Page 51: Heterologous expression and characterization of the

42

Figure 3.9. Protein expression testing of LpeCEB. The crude cell lysates of E. coli either

containing the vector with the lpeACEB genes or containing the empty vectors are compared.

The cells are induced with IPTG and are harvested at induction, or 1, 2 or 3 hours post-induction.

The appearance of a band at the molecular weight of LpeC is visible post-induction in the

lpeACEB-containing cells. Cells were grown at 37°C, and induced with 100 uM IPTG at 20°C.

It is becoming clear that the success of a heterologous expression experiment is dependent on the

lasso peptide and its biosynthetic enzymes as well as the host organism. It is therefore difficult to

predict whether the expression will be successful and what the resulting yield will be. Although

there have bene significant challenges associated with the expression of actinobacterial lasso

peptides in E. coli, the expression of these peptides in an actinobacterial host has been

successful. In light of these findings, I shifted my approach to express LP2006 in a Streptomyces

host strain.

Page 52: Heterologous expression and characterization of the

43

3.2.2 Heterologous expression in Streptomyces coelicolor M1146

To overcome the challenges of expressing an actinobacterial lasso peptide in E. coli, I attempted

to heterologously express LP2006 in a host more closely related to Nocardiopsis sp. HB141. I

chose to try several strains of Streptomyces sp. that have been designed for the heterologous

expression of secondary metabolites, S. coelicolor M1146 and M1154 and S. avermitalis

SUKA22. S. coelicolor M1146 and M1154 are strains that were derived from S. coelicolor M145

through the deletion of four endogenous secondary metabolite gene clusters, allowing for the

strains to allocate resources for the production of heterologous secondary metabolites124. S.

coelicolor M1154 contains two additional point mutations in rpoB and rpsL from S. coelicolor

M1146, which have both been shown to enhance levels of antibiotic production134,135.

Additionally, I tried using S. albus J1074, a strain commonly used as a chassis strain for the

production of secondary metabolites due to its naturally minimized genome and fast growth136.

In contrast to the approach taken the heterologous expression of LP2006 in E. coli, for the

heterologous expression in the Streptomyces host strains, I PCR-amplified and cloned the entire

LP2006 cluster into a cloning vector (Figure 3.10). This approach takes into account the

possibility that the ABC transporters are required for LP2006 biosynthesis. The entire cluster

was cloned into a modified version of the nonreplicative pSET152 plasmid, which contains the

attP site and the integrase φC31 and can integrate in a site-specific manner into the attB site of

the chromosomal φC31 phage. A version of the pSET152 plasmid containing the promotor of the

erythromycin resistance gene (ermE), modified to provide strong, constitutive expression, was

used in this study126.

Page 53: Heterologous expression and characterization of the

44

Figure 3.10. Construct used for S. coelicolor heterologous expression. The entire LP2006

gene cluster was amplified from Nocardiopsis sp. HB141 and cloned into the vector pSET152,

containing the constitutive promotor ermE*p.

Conjugation was attempted using the strains S. coelicolor M1146, M1154, S. avermitalis

SUKA22 and S. albus J1074, but exconjugants were only detected from S. coelicolor M1146 and

M1154. S. coelicolor M1154 exconjugants had a substantial growth defect, thus S. coelicolor

M1146 was chosen for the heterologous expression of the peptide. The integration of pSET152-

ermE*p-lpeACEBDD was confirmed by PCR amplification and Sanger sequencing; no mutations

were present in the cluster.

LP2006 was detected in trace quantities by LC-MS from 5 day MYM cultures, eluting in two

peaks (Figure 3.11).The elution pattern contrasts with that of LP2006 produced by Nocardiopsis

sp. HB141, which elutes in one peak just before 5 minutes. The mass and fragmentation pattern

of the peaks at 5 and 5.5 minutes are identical, suggesting that the molecule eluting at 5.5

minutes exists in a slightly different conformation than then peak at 5 minutes. The peak at 5.5

minutes may correspond to the unthreaded conformation of LP2006, although additional

experiments would be required to test this hypothesis.

Page 54: Heterologous expression and characterization of the

45

Figure 3.11. Detection of heterologously expressed LP2006 by mass spectrometry. Mass

spectrometry chromatogram of the ion 1003.43 m/z, the [M+2H]2+ and most abundant ion of

LP2006. LP2006 is detected in the S. coelicolor M1146 pSET152-ermE*p-lpeACEBDD culture

but not the S. coelicolor M1146 negative control culture.

Additionally, there is a mass discrepancy between LP2006 produced by Streptomyces M1146

and Nocardiopsis sp. HB141 (Table 3.3). Although the y6+ and y7+ ions (corresponding to the

charged C-terminal fragments of the peptide) are at the expected masses for the peptide, the b10+

and b11+ ions are shifted approximately 1000 mDa higher than expected. This suggests that there

is either a mutation in the peptide or a modification on the peptide on the N-terminal half of the

peptide. A single-nucleotide N4D or N9D mutation would result in a mass shift of 0.98401 Da,

consistent with the experimental evidence. However, no mutations were detected in the precursor

peptide when sequenced after conjugation into the S. coelicolor M1146. It is possible that the

strain may have acquired a mutation in the peptide after the conjugation stage, perhaps in

response to peptide toxicity. Another possibility is that the mass shift is a result of a post-

translational modification occurs on the peptide. Analogous to N4D or N9D mutations,

nucleophilic addition of a water in the side chain of either Asn4 or Asn9 would yield the mass

observed by LC-MS. Unfortunately, the low ion intensity of the heterologous LP2006 precludes

the measurement of more accurate ion masses which would be valuable in determining the exact

molecular formulae of the peptide and fragment ions. Additionally, the low yield precludes

structure determination by NMR, which would reveal both modifications and the conformation

of the LP2006.

Page 55: Heterologous expression and characterization of the

46

Table 3.3. Comparison of LP2006 masses from the Nocardiopsis sp. HB141 and those

produced by heterologous expression in S. coelicolor M1146.

Ion Calculated

Mass (m/z)

HB141 Extract M1146 Extract

Observed

Mass (m/z) Error (mDa)

Observed

Mass (m/z) Error (mDa)

y6+ 664.2905 664.2905 0 664.2877 -2.8

y7+ 850.3698 850.3683 -1.7 850.3655 -4.3

[M+2H]2+ 1002.9338 1002.9335 -0.3 1003.4315 497.7

b10+ 1155.4966 1155.4951 -1.3 1156.5081 1011.5

b11+ 1341.5759 1342.5654 989.5

Regardless of the exact nature of the heterologously produced LP2006, the yield of the peptide

was still too low to purify or perform functional studies. The peptide was undetectable by UV-

Vis, and only detectable in trace amounts by LC-MS. Other groups have had similar issues with

peptides expressed in Streptomyces host strains. Zong et al. attempted to heterologously express

albusnodin, encoded in the host strain S. albus DSM 41398, in the strains S. coelicolor M1146, S.

lividans 66 and S. albus J107483. Albusnodin was detected in S. coelicolor M1146 and S. lividans

66 cultures but surprisingly not in S. albus J1074. In all cases, albusnodin was truncated by a

single amino acid at the C-terminus, and it was only detectable by mass spectrometry due to low

yields, precluding NMR studies. The expression of lasso peptides in host Streptomyces strains

has worked for some groups, with some reported yields has high as 6-15 mg/L75,137.

3.3 Bioactivity of LP2006

In the absence of a heterologous producer of LP2006, I shifted efforts to produce the peptide

from the native strain, Nocardiopsis sp. HB141. Large-scale batch cultures of Nocardiopsis sp.

HB141 grown on GYM agar using aluminum baking sheets allowed the purification of nearly 1

mg of LP2006. Pure LP2006 from Nocardiopsis sp. HB141 was tested for antibacterial activity

using the broth microtiter dilution method but no activity was observed (Figure 3.12). The

solubility of the peptide was very likely the reason why no antibacterial activity was observed in

this case as precipitate was visible when resuspending the peptide into the stock solution to

perform the testing.

Page 56: Heterologous expression and characterization of the

47

Figure 3.12. Antibacterial MIC testing of pure LP2006. LP2006 was tested in duplicate

against B. subtilis, E. coli and E. coli ∆tolC ∆bamB using the broth microtiter dilution assay.

Antibacterial activity was retested by disk diffusion assay with the peptide leftover from the first

round of antibacterial testing and with newly purified peptide. LP2006 was resuspended in

DMSO to a lower concentration to prevent precipitation and was tested against the liaI-lacZ

fusion strain, B. subtilis 1A980. LP2006 does not appear to activate the cell wall stress response

gene liaI, indicted by the absence of a blue ring surrounding the zone of inhibition of LP2006

(Figure 3.13). This result contrasts with the lipid II-targeting lasso peptide, siamycin I, which

does activate liaI, suggesting that the antibacterial target of LP2006 may not be the cell wall.

Figure 3.13. LP2006 does not activate the cell wall stress response gene liaI in B. subtilis

1A980. The presence of a blue ring indicates the activation of liaI, resulting in production of

LacZ.

Page 57: Heterologous expression and characterization of the

48

This result may indicate that LP2006 does not target the cell wall, although further experiments

are required to rule out the cell wall as a target. More LP2006 is required to test against the

activation of the SOS response dinC gene using the dinC-lacZ reporter strain B. subtilis YB5018.

Additionally, LP2006-resistance mutants would like prove highly useful in elucidating the target

of LP2006.

Conclusions and future directions

The RiPPs are a diverse and underexplored class of peptides many of which have potent activity.

A subclass of the RiPPs, the lasso peptides, are highly stable to heat and proteolysis owing to

their eponymous structure and disulfide bonds. Many lasso peptides have antimicrobial activity,

targeting RNA polymerase, the ClpC1 protease and lipid II with unique mechanisms. The targets

of 8 lasso peptides have been characterized to date, and many more with antibacterial activity

have uncharacterized targets. LP2006, produced by Nocardiopsis sp. HB141, is a structurally

unique lasso peptide as the only class IV lasso peptide reported to date. It has antibacterial

activity against Gram-positive bacteria including Bacillus anthracis, vancomycin-resistant

Enterococcus faecium and Mycobacterium smegmatis. To investigate the activity of LP2006, the

heterologous expression of the peptide was pursued. First, expression was attempted using an E.

coli host strain, but with LP2006 remaining undetectable by LC-MS and the absence of

comparable expression in the literature, a strategy using a Streptomyces host strain was pursued

instead. LP2006 was successfully heterologously expressed in Streptomyces coelicolor M1146 as

detected by mass spectrometry, but low yield precluded any further experiments with the

heterologously expressed peptide. Additionally, an unexplained shift in the mass and retention

time calls into question the structure and conformation of the heterologously expressed peptide.

Ultimately, the LP2006 was purified from the native strain, Nocardiopsis sp. HB141, and tested

for antimicrobial activity and the activation of the cell wall stress response gene liaI. LP2006

does not appear to activate liaI, unlike the lasso peptide siamycin-I. This suggests that LP2006

differs from siamycin-I in its antibacterial mechanism and considering its unique structure, may

ultimately prove to have a mechanism unique from all other antibacterial lasso peptides.

Exploring the biology and chemistry of novel classes of antimicrobial compounds such as the

lasso peptides is important in the current climate of escalating antibiotic resistance. Identifying

Page 58: Heterologous expression and characterization of the

49

the target of LP2006 will expand our knowledge of both the lasso peptides and the ability of

nature to produce chemically and functionally diverse molecules. In particular, the generation of

LP2006-resistant mutants would likely provide insight into the mechanism of action of the

peptide, at least revealing whether LP2006 shares a mechanism of action with the other

antibacterial lasso peptides. LP2006 could be screened against a library such as the B. subtilis

single gene deletion library to reveal LP2006-resistant and/or hypersensitive strains138.

Dependent upon which genes confer resistance to LP2006 and which B. subtilis single gene

deletion strains have altered susceptibility, molecular target inhibition assays could be performed

to understand the nature of its activity. Ultimately, a crystal or NMR structure would reveal the

precise interactions between LP2006 and its target. I believe this study has helped understand a

promising and understudied group of antibiotics, the lasso peptides, which will enable their

translation into therapeutic and industrial applications.

Page 59: Heterologous expression and characterization of the

50

References

1. Newman, D. J. & Cragg, G. M. Natural Products as Sources of New Drugs from 1981 to

2014. J. Nat. Prod. 79, 629–661 (2016).

2. Arnison, P. G. et al. Ribosomally synthesized and post-translationally modified peptide

natural products: overview and recommendations for a universal nomenclature. Nat. Prod.

Rep. 30, 108–160 (2013).

3. Further Observations on an Inhibitory Substance (Nisin) From Lactic Streptococci. Lancet

250, 5–8 (1947).

4. Highland, J. H. et al. Identification of a Ribosomal Protein Necessary for Thiostrepton

Binding to Escherichia coli Ribosomes. J. Biol. Chem. 250, 1141–1145 (1975).

5. Olivera, B. M. et al. Peptide neurotoxins from fish-hunting cone snails. Science 230,

1338–1343 (1985).

6. Schmidtko, A., Lötsch, J., Freynhagen, R. & Geisslinger, G. Ziconotide for treatment of

severe chronic pain. Lancet 375, 1569–1577 (2010).

7. Tan, S., Moore, G. & Nodwell, J. Put a Bow on It: Knotted Antibiotics Take Center Stage.

Antibiotics 8, 117 (2019).

8. Woodward, S. R., Cruz, L. J., Olivera, B. M. & Hillyard, D. R. Constant and

hypervariable regions in conotoxin propeptides. EMBO J. 9, 1015–20 (1990).

9. Hensens, O. D. & Albers-Schönberg, G. Total structure of the peptide antibiotic

components of thiopeptin by 1H and 13C NMR spectroscopy. Tetrahedron Lett. 19,

3649–3652 (1978).

10. Wieland Brown, L. C., Acker, M. G., Clardy, J., Walsh, C. T. & Fischbach, M. A.

Thirteen posttranslational modifications convert a 14-residue peptide into the antibiotic

thiocillin. Proc. Natl. Acad. Sci. 106, 2549–2553 (2009).

11. Kelly, W. L., Pan, L. & Li, C. Thiostrepton Biosynthesis: Prototype for a New Family of

Bacteriocins. J. Am. Chem. Soc. 131, 4327–4334 (2009).

12. SU, T. L. Micrococcin, an antibacterial substance formed by a strain of Micrococcus. Br.

J. Exp. Pathol. 29, 473–81 (1948).

13. Bagley, M., Dale, J., Merritt, E. & Xiong, X. Thiopeptide Antibiotics. Chem. Rev. 105.

685-714 (2005).

14. Cromwell, G. L., Stahly, T. S., Speer, V. C. & O’Kelly, R. Efficacy of Nosiheptide as a

Page 60: Heterologous expression and characterization of the

51

Growth Promotant for Growing-Finishing Swine-A Cooperative Study. J. Anim. Sci. 59,

1125–1128 (1984).

15. Dennis, S. M., Nagaraja, T. G. & Dayton, A. D. Effect of lasalocid, monensin and

thiopeptin on rumen protozoa. Res. Vet. Sci. 41, 251–256 (1986).

16. Miyairi, N. et al. Thiopeptin, a new feed additive antibiotic: microbiological and chemical

studies. Antimicrob. Agents Chemother. 1, 192–6 (1972).

17. Liao, R. et al. Thiopeptide biosynthesis featuring ribosomally synthesized precursor

peptides and conserved posttranslational modifications. Chem. Biol. 16, 141–7 (2009).

18. Morris, R. P. et al. Ribosomally Synthesized Thiopeptide Antibiotics Targeting

Elongation Factor Tu. J. Am. Chem. Soc. 131, 5946–5955 (2009).

19. Li, Y.-M. et al. From Peptide Precursors to Oxazole and Thiazole-Containing Peptide

Antibiotics: Microcin B17 Synthase. Science (80-. ). 274, 1188–1193 (1996).

20. Lee, S. W. et al. Discovery of a widely distributed toxin biosynthetic gene cluster. Proc.

Natl. Acad. Sci. 105, 5879–5884 (2008).

21. Dunbar, K. L., Melby, J. O. & Mitchell, D. A. YcaO domains use ATP to activate amide

backbones during peptide cyclodehydrations. Nat. Chem. Biol. 8, 569–575 (2012).

22. Zhang, Z. et al. Biosynthetic Timing and Substrate Specificity for the Thiopeptide

Thiomuracin. J. Am. Chem. Soc. 138, 15511–15514 (2016).

23. Bycroft, B. W. & Gowland, M. S. The structures of the highly modified peptide antibiotics

micrococcin P1 and P2. J. Chem. Soc. Chem. Commun.6, 256-258 (1978).

24. Yu, Y. et al. Nosiheptide biosynthesis featuring a unique indole side ring formation on the

characteristic thiopeptide framework. ACS Chem. Biol. 4, 855–64 (2009).

25. Zhang, Q. et al. Radical-mediated enzymatic carbon chain fragmentation-recombination.

Nat. Chem. Biol. 7, 154–60 (2011).

26. Hudson, G. A., Zhang, Z., Tietz, J. I., Mitchell, D. A. & Donk, W. A. van der. In Vitro

Biosynthesis of the Core Scaffold of the Thiopeptide Thiomuracin. J. Am. Chem. Soc.

137, 16012 (2015).

27. Bower, J. et al. Structure-based design of agents targeting the bacterial ribosome. Bioorg.

Med. Chem. Lett. 13, 2455–2458 (2003).

28. Weisblum, B. & Demohn, V. Thiostrepton, an inhibitor of 5OS ribosome subunit function.

J. Bacteriol. 101, 1073–5 (1970).

Page 61: Heterologous expression and characterization of the

52

29. Harms, J. M. et al. Translational Regulation via L11: Molecular Switches on the

Ribosome Turned On and Off by Thiostrepton and Micrococcin. Mol. Cell 30, 26–38

(2008).

30. Cundliffe, E. et al. Ribosomes in thiostrepton-resistant mutants of Bacillus megaterium

lacking a single 50 S subunit protein. J. Mol. Biol. 132, 235–252 (1979).

31. Cameron, D. M., Thompson, J., Gregory, S. T., March, P. E. & Dahlberg, A. E.

Thiostrepton-resistant mutants of Thermus thermophilus. Nucleic Acids Res. 32, 3220–

3227 (2004).

32. Porse, B. T., Leviev, I., Mankin, A. S. & Garrett, R. A. The antibiotic thiostrepton inhibits

a functional transition within protein L11 at the ribosomal GTPase centre. J. Mol. Biol.

276, 391–404 (1998).

33. Wienen, B. et al. Ribosomal protein alterations in thiostrepton- and Micrococcin-resistant

mutants of Bacillus subtilis. J. Biol. Chem. 254, 8031–41 (1979).

34. Hummel, H. & Böck, A. Thiostrepton resistance mutations in the gene for 23S ribosomal

RNA of halobacteria. Biochimie 69, 857–61 (1987).

35. Thompson, J., Schmidt, F. & Cundliffe, E. Site of action of a ribosomal RNA methylase

conferring resistance to thiostrepton. J. Biol. Chem. 257, 7915–7 (1982).

36. Zuurmond, A.-M. et al. GE2270A-resistant Mutations in Elongation Factor Tu Allow

Productive Aminoacyl-tRNA Binding to EF-Tu·GTP·GE2270A Complexes. J. Mol. Biol.

304, 995-1005 (2000).

37. Möhrle, V. G., Tieleman, L. N. & Kraal, B. Elongation Factor Tu1 of the Antibiotic

GE2270A producer planobispora rosea has an unexpected resistance profile against EF-Tu

targeted antibiotics. Biochemical And Biophysical Research Communications 230, (1997).

38. Hashimoto, M. et al. An RNA polymerase inhibitor, cyclothiazomycin B1, and its isomer.

Bioorg. Med. Chem. 14, 8259–8270 (2006).

39. Hayashi, S. et al. Genome Mining Reveals a Minimum Gene Set for the Biosynthesis of

32-Membered Macrocyclic Thiopeptides Lactazoles. Chem. Biol. 21, 679–688 (2014).

40. Rogers, L. A. The inhibiting effect of streptococcus lactis on lactobacillus bulgaricus. J.

Bacteriol. 16, 321–5 (1928).

41. Cleveland, J., Montville, T. J., Nes, I. F. & Chikindas, M. L. Bacteriocins: safe, natural

antimicrobials for food preservation. Int. J. Food Microbiol. 71, 1–20 (2001).

Page 62: Heterologous expression and characterization of the

53

42. Cao, L. T., Wu, J. Q., Xie, F., Hu, S. H. & Mo, Y. Efficacy of Nisin in Treatment of

Clinical Mastitis in Lactating Dairy Cows. J. Dairy Sci. 90, 3980–3985 (2007).

43. Crowther, G. S. et al. Evaluation of NVB302 versus vancomycin activity in an in vitro

human gut model of Clostridium difficile infection. J. Antimicrob. Chemother. 68, 168–

176 (2013).

44. Knerr, P. J. & van der Donk, W. A. Discovery, Biosynthesis, and Engineering of

Lantipeptides. Annu. Rev. Biochem. 81, 479–505 (2012).

45. Kodani, S. et al. From The Cover: The SapB morphogen is a lantibiotic-like peptide

derived from the product of the developmental gene ramS in Streptomyces coelicolor.

Proc. Natl. Acad. Sci. 101, 11448–11453 (2004).

46. Blin, K. et al. antiSMASH 5.0: updates to the secondary metabolite genome mining

pipeline. Nucleic Acids Res. 47, W81–W87 (2019).

47. Price, M. N., Dehal, P. S. & Arkin, A. P. FastTree 2 – Approximately Maximum-

Likelihood Trees for Large Alignments. PLoS One 5, e9490 (2010).

48. Letunic, I. & Bork, P. Interactive Tree Of Life (iTOL) v4: recent updates and new

developments. Nucleic Acids Res. 47, W256–W259 (2019).

49. Jung, G. Lantibiotics—Ribosomally Synthesized Biologically Active Polypeptides

containing Sulfide Bridges and α,β‐Didehydroamino Acids. Angew. Chemie Int. Ed.

English 30, 1051–1068 (1991).

50. Pag, U. & Sahl, H.-G. Multiple Activities in Lantibiotics - Models for the Design of Novel

Antibiotics? Curr. Pharm. Des. 8, 815–833 (2002).

51. Willey, J. M. & van der Donk, W. A. Lantibiotics: Peptides of Diverse Structure and

Function. Annu. Rev. Microbiol. 61, 477–501 (2007).

52. Ortega, M. A. & Van Der Donk, W. A. New Insights into the Biosynthetic Logic of

Ribosomally Synthesized and Post-translationally Modified Peptide Natural Products. Cell

Chem. Biol. 23, 31–44 (2016).

53. Meindl, K. et al. Labyrinthopeptins: A New Class of Carbacyclic Lantibiotics. Angew.

Chemie Int. Ed. 49, 1151–1154 (2010).

54. Ortega, M. A. et al. Structure and mechanism of the tRNA-dependent lantibiotic

dehydratase NisB. Nature 517, 509–512 (2015).

55. Chatterjee, C. et al. Lacticin 481 Synthetase Phosphorylates its Substrate during

Page 63: Heterologous expression and characterization of the

54

Lantibiotic Production. J. Am. Chem. Soc. 127, 15332–15333 (2005).

56. Kupke, T., Kempter, C., Jung, G. & Gotz, F. Oxidative decarboxylation of peptides

catalyzed by flavoprotein EpiD. Determination of substrate specificity using peptide

libraries and neutral loss mass spectrometry. J. Biol. Chem. 270, 11282–11289 (1995).

57. Ökesli, A., Cooper, L. E., Fogle, E. J. & van der Donk, W. A. Nine post-translational

modifications during the biosynthesis of cinnamycin. J. Am. Chem. Soc. 133, 13753–60

(2011).

58. Breukink, E. et al. Use of the cell wall precursor lipid II by a pore-forming peptide

antibiotic. Science 286, 2361–4 (1999).

59. Wiedemann, I. et al. Specific binding of nisin to the peptidoglycan precursor lipid II

combines pore formation and inhibition of cell wall biosynthesis for potent antibiotic

activity. J. Biol. Chem. 276, 1772–9 (2001).

60. Hsu, S.-T. D. et al. The nisin–lipid II complex reveals a pyrophosphate cage that provides

a blueprint for novel antibiotics. Nat. Struct. Mol. Biol. 11, 963–967 (2004).

61. Hasper, H. E. et al. An Alternative Bactericidal Mechanism of Action for Lantibiotic

Peptides That Target Lipid II. Science 313, 1636–1637 (2006).

62. Brotz, H., Bierbaum, G., Reynolds, P. E. & Sahl, H.-G. The Lantibiotic Mersacidin

Inhibits Peptidoglycan Biosynthesis at the Level of Transglycosylation. Eur. J. Biochem.

246, 193–199 (1997).

63. Kodani, S., Lodato, M. A., Durrant, M. C., Picart, F. & Willey, J. M. SapT, a lanthionine-

containing peptide involved in aerial hyphae formation in the streptomycetes. Mol.

Microbiol. 58, 1368–1380 (2005).

64. Wang, H. & van der Donk, W. A. Biosynthesis of the class III lantipeptide catenulipeptin.

ACS Chem. Biol. 7, 1529–35 (2012).

65. Draper, L. A., Cotter, P. D., Hill, C. & Ross, R. P. Lantibiotic resistance. Microbiol. Mol.

Biol. Rev. 79, 171–91 (2015).

66. Hill, C., Draper, L., Ross, R. & Cotter, P. Lantibiotic Immunity. Curr. Protein Pept. Sci.

9, 39–49 (2008).

67. Mazzotta, A. S. & Montville, T. J. Nisin induces changes in membrane fatty acid

composition of Listeria monocytogenes nisin-resistant strains at 10 degrees C and 30

degrees C. J. Appl. Microbiol. 82, 32–8 (1997).

Page 64: Heterologous expression and characterization of the

55

68. Verheul, A., Russell, N. J., Van’T Hof, R., Rombouts, F. M. & Abee, T. Modifications of

membrane phospholipid composition in nisin-resistant Listeria monocytogenes Scott A.

Appl. Environ. Microbiol. 63, 3451–7 (1997).

69. Staubitz, P., Neumann, H., Schneider, T., Wiedemann, I. & Peschel, A. MprF-mediated

biosynthesis of lysylphosphatidylglycerol, an important determinant in staphylococcal

defensin resistance. FEMS Microbiol. Lett. 231, 67–71 (2004).

70. Froseth, B. R. & McKay, L. L. Molecular characterization of the nisin resistance region of

Lactococcus lactis subsp. lactis biovar diacetylactis DRC3. Appl. Environ. Microbiol. 57,

804–11 (1991).

71. Sun, Z. et al. Novel Mechanism for Nisin Resistance via Proteolytic Degradation of Nisin

by the Nisin Resistance Protein NSR. Antimicrob. Agents Chemother. 53, 1964–1973

(2009).

72. Zimmermann, M., Hegemann, J. D., Xie, X. & Marahiel, M. A. Characterization of

caulonodin lasso peptides revealed unprecedented N-terminal residues and a precursor

motif essential for peptide maturation. Chem. Sci. 5, 4032–4043 (2014).

73. Salomón, R. A., Farías, R. N., Salomn, R. A. & Farias, R. N. Microcin 25, a novel

antimicrobial peptide produced by Escherichia coli. J. Bacteriol. 174, 7428–35 (1992).

74. Zimmermann, M., Hegemann, J. D., Xie, X. & Marahiel, M. A. The Astexin-1 Lasso

Peptides: Biosynthesis, Stability, and Structural Studies. Cell Chem. Biol. 20, 558–569

(2013).

75. Tietz, J. I. et al. A new genome-mining tool redefines the lasso peptide biosynthetic

landscape. Nat. Chem. Biol. 13, 470–478 (2017).

76. Duquesne, S. et al. Two Enzymes Catalyze the Maturation of a Lasso Peptide in

Escherichia coli. Chem. Biol. 14, 793–803 (2007).

77. Yan, K.-P. et al. Dissecting the Maturation Steps of the Lasso Peptide Microcin J25 in

vitro. ChemBioChem 13, 1046–1052 (2012).

78. DiCaprio, A. J., Firouzbakht, A., Hudson, G. A. & Mitchell, D. A. Enzymatic

Reconstitution and Biosynthetic Investigation of the Lasso Peptide Fusilassin. J. Am.

Chem. Soc. 141, 290–297 (2019).

79. Burkhart, B. J., Hudson, G. A., Dunbar, K. L. & Mitchell, D. A. A prevalent peptide-

binding domain guides ribosomal natural product biosynthesis. Nat. Chem. Biol. 11, 564–

Page 65: Heterologous expression and characterization of the

56

570 (2015).

80. Sumida, T., Dubiley, S., Wilcox, B., Severinov, K. & Tagami, S. Structural Basis of

Leader Peptide Recognition in Lasso Peptide Biosynthesis Pathway. ACS Chem. Biol. 14,

1619–1627 (2019).

81. Solbiati, J. O., Ciaccio, M., Farías, R. N. & Salomón, R. A. Genetic analysis of plasmid

determinants for microcin J25 production and immunity. J. Bacteriol. 178, 3661–3 (1996).

82. Zhu, S. et al. Insights into the Unique Phosphorylation of the Lasso Peptide Paeninodin. J.

Biol. Chem. 291, 13662–13678 (2016).

83. Zong, C., Cheung-Lee, W. L., Elashal, H. E., Raj, M. & Link, A. J. Albusnodin: An

acetylated lasso peptide from: Streptomyces albus. Chem. Commun. 54, 1339–1342

(2018).

84. Gavrish, E. et al. Lassomycin, a Ribosomally Synthesized Cyclic Peptide, Kills

Mycobacterium tuberculosis by Targeting the ATP-Dependent Protease ClpC1P1P2.

Chem. Biol. 21, 509–518 (2014).

85. Pavlova, O., Mukhopadhyay, J., Sineva, E., Ebright, R. H. & Severinov, K. Systematic

structure-activity analysis of microcin J25. J. Biol. Chem. 283, 25589–25595 (2008).

86. Constantine, K. L. et al. High-resolution solution structure of siamycin II: Novel

amphipathic character of a 21-residue peptide that inhibits HIV fusion. J. Biomol. NMR 5,

271–286 (1995).

87. Katahira, R., Shibata, K., Yamasaki, M., Matsuda, Y. & Yoshida, M. Solution structure of

endothelin B receptor selective antagonist RES-701-1 determined by 1H NMR

spectroscopy. Bioorg. Med. Chem. 3, 1273–1280 (1995).

88. Knappe, T. A., Linne, U., Xie, X. & Marahiel, M. A. The glucagon receptor antagonist

BI-32169 constitutes a new class of lasso peptides. FEBS Lett. 584, 785–789 (2010).

89. Um, S. et al. Sungsanpin, a lasso peptide from a deep-sea streptomycete. J. Nat. Prod. 76,

873–879 (2013).

90. Daniel-Ivad, M. et al. An Engineered Allele of afsQ1 Facilitates the Discovery and

Investigation of Cryptic Natural Products. ACS Chem. Biol. 12, 628–634 (2017).

91. Tsunakawa, M. et al. Siamycins I and II, New Anti-HIV Peptides. I. Fermentation,

Isolation, Biological Activity and Initial Characterization. J. Antibiot. (Tokyo). 48, 433–

434 (1995).

Page 66: Heterologous expression and characterization of the

57

92. Tan, S., Ludwig, K. C., Mueller, A., Schneider, T. & Nodwell, J. R. The Lasso Peptide

Siamycin‐I Targets Lipid II at the Gram-Positive Cell Surface. ACS Chem. Biol. 14, 966–

974 (2019).

93. Dubrac, S., Bisicchia, P., Devine, K. M. & Msadek, T. A matter of life and death: Cell

wall homeostasis and the WalKR (YycGF) essential signal transduction pathway. Mol.

Microbiol. 70, 1307–1322 (2008).

94. Tiyanont, K. et al. Imaging peptidoglycan biosynthesis in Bacillus subtilis with

fluorescent antibiotics. Proc. Natl. Acad. Sci. 103, 11033–11038 (2006).

95. Lin, P. F. et al. Characterization of siamycin I, a human immunodeficiency virus fusion

inhibitor. Antimicrob. Agents Chemother. 40, 133–138 (1996).

96. Nakayama, J. et al. Siamycin Attenuates fsr Quorum Sensing Mediated by a Gelatinase

Biosynthesis-Activating Pheromone in Enterococcus faecalis. J. Bacteriol. 189, 1358–

1365 (2007).

97. Metelev, M. et al. Structure, Bioactivity, and Resistance Mechanism of

Streptomonomicin, an Unusual Lasso Peptide from an Understudied Halophilic

Actinomycete. Chem. Biol. 22, 241–250 (2015).

98. Salomon, R. A. & Farias, R. N. The FhuA Protein Is Involved in Microcin 25 Uptake. J.

Bacteriol. 175, 7741–7742 (1993).

99. Salomon, R. A. & Farias, R. N. The Peptide Antibiotic Microcin 25 Is Imported through

the TonB Pathway and the SbmA Protein. J. Bacteriol. 177, 3323–3325 (1995).

100. Delgado, M. A. A. et al. Escherichia coli RNA polymerase is the target of the

cyclopeptide antibiotic microcin J25. J. Bacteriol. 183, 4543–50 (2001).

101. Yuzenkova, J. et al. Mutations of bacterial RNA polymerase leading to resistance to

microcin j25. J. Biol. Chem. 277, 50867–75 (2002).

102. Korzheva, N. et al. A structural model of transcription elongation. Science 289, 619–25

(2000).

103. Adelman, K. et al. Molecular mechanism of transcription inhibition by peptide antibiotic

Microcin J25. Mol. Cell 14, 753–62 (2004).

104. Mukhopadhyay, J., Sineva, E., Knight, J., Levy, R. M. & Ebright, R. H. Antibacterial

peptide microcin J25 inhibits transcription by binding within and obstructing the RNA

polymerase secondary channel. Mol. Cell 14, 739–51 (2004).

Page 67: Heterologous expression and characterization of the

58

105. Braffman, N. R. et al. Structural mechanism of transcription inhibition by lasso peptides

microcin J25 and capistruin. Proc. Natl. Acad. Sci. 116, 1273–1278 (2019).

106. Blond, A. et al. The cyclic structure of microcin J25, a 21-residue peptide antibiotic from

Escherichia coli. Eur. J. Biochem. 259, 747–756 (2001).

107. Blond, A. et al. Solution structure of microcin J25, the single macrocyclic antimicrobial

peptide from Escherichia coli. Eur. J. Biochem. 268, 2124–2133 (2001).

108. Blond, A. et al. Thermolysin-linearized microcin J25 retains the structured core of the

native macrocyclic peptide and displays antimicrobial activity. Eur. J. Biochem. 269,

6212–6222 (2002).

109. Bayro, M. J. et al. Structure of antibacterial peptide microcin J25: A 21-residue lariat

protoknot. J. Am. Chem. Soc. 125, 12382–12383 (2003).

110. Knappe, T. A. et al. Isolation and Structural Characterization of Capistruin, a Lasso

Peptide Predicted from the Genome Sequence of Burkholderia thailandensis E264. J. Am.

Chem. Soc. 130, 11446–11454 (2008).

111. Kuznedelov, K. et al. The Antibacterial Threaded-lasso Peptide Capistruin Inhibits

Bacterial RNA Polymerase. J. Mol. Biol. 412, 842–848 (2011).

112. Metelev, M. et al. Acinetodin and Klebsidin, RNA Polymerase Targeting Lasso Peptides

Produced by Human Isolates of Acinetobacter gyllenbergii and Klebsiella pneumoniae.

ACS Chem. Biol. 12, 814–824 (2017).

113. Cheung-Lee, W. L., Parry, M. E., Cartagena, A. J., Darst, S. A. & James Link, A.

Discovery and structure of the antimicrobial lasso peptide citrocin. J. Biol. Chem. 294,

6822–6830 (2019).

114. Lear, S. et al. Total chemical synthesis of lassomycin and lassomycin-amide. Org. Biomol.

Chem. 14, 4534–4541 (2016).

115. Lee, H. & Suh, J.-W. Anti-tuberculosis lead molecules from natural products targeting

Mycobacterium tuberculosis ClpC1. J Ind Microbiol Biotechnol 43, 205–212 (2016).

116. Maksimov, M. O., Pelczer, I. & Link, A. J. Precursor-centric genome-mining approach for

lasso peptide discovery. Proc. Natl. Acad. Sci. 109, 15223–8 (2012).

117. Agrawal, P., Khater, S., Gupta, M., Sain, N. & Mohanty, D. RiPPMiner: a bioinformatics

resource for deciphering chemical structures of RiPPs based on prediction of cleavage and

cross-links. Nucleic Acids Res. 45, W80–W88 (2017).

Page 68: Heterologous expression and characterization of the

59

118. van Heel, A. J. et al. BAGEL4: a user-friendly web server to thoroughly mine RiPPs and

bacteriocins. Nucleic Acids Res. (2018). doi:10.1093/nar/gky383

119. Skinnider, M. A., Merwin, N. J., Johnston, C. W. & Magarvey, N. A. PRISM 3: expanded

prediction of natural product chemical structures from microbial genomes. Nucleic Acids

Res. 45, W49–W54 (2017).

120. Santos-Aberturas, J. et al. Uncovering the unexplored diversity of thioamidated ribosomal

peptides in Actinobacteria using the RiPPER genome mining tool. Nucleic Acids Res. 47,

4624–4637 (2019).

121. Martin-Gómez, H., Linne, U., Albericio, F., Tulla-Puche, J. & Hegemann, J. D.

Investigation of the Biosynthesis of the Lasso Peptide Chaxapeptin Using an E. coli-

Based Production System. J. Nat. Prod. 81, 2050-2056 (2018).

122. Schneemann, I. et al. Comprehensive investigation of marine Actinobacteria associated

with the sponge Halichondria panicea. Appl. Environ. Microbiol. 76, 3702–14 (2010).

123. Komatsu, M. et al. Engineered Streptomyces avermitilis host for heterologous expression

of biosynthetic gene cluster for secondary metabolites. ACS Synth. Biol. 2, 384–96 (2013).

124. Pablo Gomez-Escribano, J. & Bibb, M. J. Engineering Streptomyces coelicolor for

heterologous expression of secondary metabolite gene clusters. Microbial Biotechnology

4, 207-215 (2010).

125. Jani, C. et al. Streptomyces: a screening tool for bacterial cell division inhibitors. J.

Biomol. Screen. 20, 275–84 (2015).

126. Bibb, M. J., Janssen, G. R. & Ward, J. M. Cloning and analysis of the promoter region of

the erythromycin resistance gene (ermE) of Streptomyces erythraeus. Gene 38, 215–226

(1985).

127. Kieser T, Bibb MJ, Buttner MJ & Chater KF. Practical Streptomyces Genetics. 2nd ed.

(2000).

128. Jensen, P. R., Williams, P. G., Oh, D.-C., Zeigler, L. & Fenical, W. Species-specific

secondary metabolite production in marine actinomycetes of the genus Salinispora. Appl.

Environ. Microbiol. 73, 1146–52 (2007).

129. Koos, J. D. & Link, A. J. Heterologous and in Vitro Reconstitution of Fuscanodin, a Lasso

Peptide from Thermobifida fusca. J. Am. Chem. Soc. 141, 928–935 (2019).

130. Zhu, S. et al. The B1 Protein Guides the Biosynthesis of a Lasso Peptide. Sci. Rep. 6,

Page 69: Heterologous expression and characterization of the

60

35604 (2016).

131. Rosano, G. L. & Ceccarelli, E. A. Recombinant protein expression in Escherichia coli:

advances and challenges. Front. Microbiol. 5, 172 (2014).

132. Garg, N., Tang, W., Goto, Y., Nair, S. K. & van der Donk, W. A. Lantibiotics from

Geobacillus thermodenitrificans. Proc. Natl. Acad. Sci. U. S. A. 109, 5241–6 (2012).

133. Schmidt, E. W. et al. 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 (2005).

134. Shima, J., Hesketh, A., Okamoto, S., Kawamoto, S. & Ochi, K. Induction of actinorhodin

production by rpsL (encoding ribosomal protein S12) mutations that confer streptomycin

resistance in Streptomyces lividans and Streptomyces coelicolor A3(2). J. Bacteriol. 178,

7276–7284 (1996).

135. Hu, H., Zhang, Q. & Ochi, K. Activation of antibiotic biosynthesis by specified mutations

in the rpoB gene (encoding the RNA polymerase beta subunit) of Streptomyces lividans.

J. Bacteriol. 184, 3984–91 (2002).

136. Zaburannyi, N., Rabyk, M., Ostash, B., Fedorenko, V. & Luzhetskyy, A. Insights into

naturally minimised Streptomyces albus J1074 genome. BMC Genomics 15, 97 (2014).

137. Li, Y. et al. Characterization of Sviceucin from Streptomyces Provides Insight into

Enzyme Exchangeability and Disulfide Bond Formation in Lasso Peptides. ACS Chem.

Biol. 10, 2641–2649 (2015).

138. Koo, B.-M. et al. Construction and Analysis of Two Genome-Scale Deletion Libraries for

Bacillus subtilis. Cell Syst. 4, 291-305.e7 (2017).

139. Ling, L. L. et al. A new antibiotic kills pathogens without detectable resistance. Nature

517, 455–459 (2015).

140. Feling, R. H. et al. Salinosporamide A: A Highly Cytotoxic Proteasome Inhibitor from a

Novel Microbial Source, a Marine Bacterium of the New Genus Salinospora. Angew.

Chemie Int. Ed. 42, 355–357 (2003).

141. Kumar, S., Stecher, G. & Tamura, K. MEGA7: Molecular Evolutionary Genetics Analysis

Version 7.0 for Bigger Datasets. Mol. Biol. Evol. 33, 1870–1874 (2016).

142. Tanizawa, Y., Fujisawa, T. & Nakamura, Y. DFAST: a flexible prokaryotic genome

annotation pipeline for faster genome publication. Bioinformatics 34, 1037–1039 (2018).

Page 70: Heterologous expression and characterization of the

61

143. Aziz, R. K. et al. The RAST Server: Rapid Annotations using Subsystems Technology.

BMC Genomics 9, 75 (2008).

144. Kim, M., Oh, H.-S., Park, S.-C. & Chun, J. Towards a taxonomic coherence between

average nucleotide identity and 16S rRNA gene sequence similarity for species

demarcation of prokaryotes. Int. J. Syst. Evol. Microbiol. 64, 346–351 (2014).

145. Stewart, E. J. Growing unculturable bacteria. J. Bacteriol. 194, 4151–60 (2012).

146. Ritchie, K. B. et al. Survey of Antibiotic-producing Bacteria Associated with the

Epidermal Mucus Layers of Rays and Skates. Front. Microbiol. 8, 1050 (2017).

147. Gärdes, A. et al. Complete genome sequence of Marinobacter adhaerens type strain

(HP15), a diatom-interacting marine microorganism. Stand. Genomic Sci. 3, 97–107

(2010).

148. Hollensteiner, J., Poehlein, A. & Daniel, R. Complete Genome Sequence of Marinobacter

sp. Strain JH2, Isolated from Seawater of the Kiel Fjord. Microbiol. Resour. Announc. 8,

(2019).

149. Martinez, J. S. & Butler, A. Marine amphiphilic siderophores: Marinobactin structure,

uptake, and microbial partitioning. J. Inorg. Biochem. 101, 1692–1698 (2007).

150. Robinson, S. L., Christenson, J. K. & Wackett, L. P. Biosynthesis and chemical diversity

of β-lactone natural products. Nat. Prod. Rep. 36, 458–475 (2019).

Page 71: Heterologous expression and characterization of the

62

Appendix 1 Screen for novel bioactive natural products from marine bacteria

Appendix 1

5.1 Introduction

The discovery of antibiotics has resulted in substantial improvements in the outcome of

treatments for bacterial infections and has enabled certain medical procedures to occur. Natural

products remain extremely relevant in the discovery and development of novel antibiotics, even

in the years following the golden age of antibiotics (~1940-1960). In fact, over the past 35 years,

roughly 60% of approved antibiotics are derivatives of natural products.1

Due to the challenges such as high rediscovery rates of known compounds, interest in natural

product screening is not what it once was. However, since bacterial chemotype roughly follows

phylotype, in theory novel compounds can be discovered by screening uncharacterized

environmental isolates, thereby reducing rates of rediscovery128. An excellent example of this is

the discovery of the antibiotic teixobactin, which was discovered from the previously

uncharacterized microbe, Eleftheria terrae, using a unique culturing method139. A second

example of the discovery of novel natural products by studying uncharacterized microbes is the

discovery of salinosporamide A from Salinispora tropica140. S. tropica is an obligate species of

marine Actinobacteria, which produces the potent 20S proteasome inhibitor, salinosporamide A.

These examples demonstrate that bacteria are capable of producing substantial chemical

diversity, much of which has yet to be discovered. The aim of this project was to investigate the

ability of marine bacteria to produce antibacterial metabolites and to identify and characterize

novel antimicrobial compounds from the Nodwell Maritime Collection.

5.2 Methods

5.2.1 Bioactivity screen

5.2.1.1 Collection and Isolation of maritime strains

There are 42 strains that make up the Nodwell Maritime Collection (NMarC). These strains were

isolated from marine sediments collected at 5 locations in the maritime provinces of Canada

(Hopewell Rocks, New Brunswick; Brackley Beach, PEI; Cavendish Beach, PEI; Fisherman’s

Page 72: Heterologous expression and characterization of the

63

Cove, NS; Crystal Crescent Beach, NS). Marine sediment samples were collected at a depth of 1-

2 metres by Dr. Sheila Elardo. Dr. Sheila Elardo isolated the marine strains by diluting the

sediment samples with sterile artificial seawater and plating them on four types of media: M1,

M2, ISP2 and M2216 medium, each supplemented with 25% artificial seawater (5.2.1.2). The

cultures were incubated at 30°C for up to 12 weeks and individual colonies were picked and re-

streaked on the four different media types.

5.2.1.2 Culture conditions

The growth of all maritime strains was tested on four agar media types: M1 medium (per L: 10g

soluble starch, 4g yeast extract, 2g peptone, 16g agar), M2 medium (per L: 6mL glycerol, 1g

arginine, 1g K2HPO4, 0.5g MgSO4, 16g agar), ISP2 medium (per L: 4g yeast extract, 10g malt

extract, 4g dextrose, 16g agar), and Difco Marine Agar 2216 medium (per L: 55.1g Difco Marine

Agar 2216), all of which were dissolved in artificial seawater (per L: 23.477 g NaCl, 10.64 g

MgCl2 hexahydrate, 3.917 g Na2SO4, 1.102 g CaCl2, 0.664 g KCl, 0.192 g NaHCO3, 0.096 g

KBr, 0.026 g H3BO3, 0.024 g SrCl2, 0.03 g NaF) with the exception of M2216 medium which

was dissolved in Milli-Q water. For metabolite extraction and bioactivity testing, all maritime

strains were grown for 5 days at 30°C prior to methanol extraction.

The test strains of M. luteus, B. subtilis JH642, E. coli BW25113 and E. coli BW25113 ∆tolC

∆bamB were grown in LB medium (per L Milli-Q water: 10g tryptone, 10g NaCl, 5g yeast

extract) at 37°C, while S. cerevisiae Y7092 was grown in YPD medium (per L Milli-Q water:

10g Yeast extract, 20g Peptone, 20g Dextrose). 16 g of agar per 1L of medium was used to make

solid media for disk diffusion assays.

5.2.1.3 Broth microtiter dilution and disk diffusion assay

The test organism was inoculated from an agar streak plate into a 5mL liquid culture of YPD

medium for S. cerevisiae or LB medium for all other test organisms. The liquid culture was

grown overnight at 30°C (for S. cerevisiae) or at 37°C for (all other organisms) and a 1:100

subculture was started in the morning. Each culture was grown to an OD of 0.4-0.6 and diluted

1:1000 with fresh media in a 96-well plate. The test compound or extract, resuspended in

DMSO, was added to each well and the plate was incubated overnight. The next morning, the

Page 73: Heterologous expression and characterization of the

64

OD600 of each well was measured. When testing extracts, the extract was considered active if

inhibition of growth was greater than 50%. For the determination of MICs, the same protocol

was followed and antibiotics at twofold increasing concentrations were added to the 96 well plate

liquid cultures.

To perform the disk diffusion assay, the procedure indicated above was followed until the

subculture reached an OD of 0.4-0.6, at which point the culture was diluted 1:1000 in fresh

media. 100 μL of the diluted culture was spread across an agar plate of LB medium using sterile

glass beads. The plate was dried for 5 minutes and a maximum of five paper filter disks were

distributed across the plate. 2-10 μL of resuspended crude extracts were added to the filter disks

and plates were placed overnight in the incubator at 30°C (for S. cerevisiae) or at 37°C for (all

other organisms). After overnight incubation, the plates were imaged and the zone of inhibition

was measured.

5.2.2 Isolation of the Marinobacter sp. N33 bioactive metabolite(s)

5.2.2.1 Metabolite extraction

Metabolites were extracted from agar or liquid cultures with HPLC-grade methanol, using a

volume of methanol equivalent to the volume of the culture or 200-300 mL, whichever is less.

Samples were sonicated for 15 minutes and macerated overnight. In the morning, extracts were

dried using a Genevac EZ-2 Elite series evaporator (SP scientific) or a Hei-VAP Precision

Rotary Evaporator (Heidolph).

5.2.2.2 Flash chromatography and HPLC purification

Crude extracts were resuspended in 5% aqueous HPLC-grade methanol to a concentration of 100

mg/mL. After centrifugation and/or filtration, the resuspended crude extracts were further

purified on a Reveleris® X2 Flash chromatography system (Buchi Labortechnik). A 20g C18 40-

60um 100Å cartridge (Aegio Technologies) was used for the separation by flash

chromatography, with a linear gradient from 5 to 100% aqueous HPLC-grade methanol at a flow

rate of 10mL/min. 20mL fractions were collected, dried by Genevac or rotary evaporation and

tested for bioactivity by broth microtiter dilution assay (5.2.1.3). Fractions that were bioactive

were pooled, resuspended and further purified by HPLC.

Page 74: Heterologous expression and characterization of the

65

Individual and pooled bioactive flash chromatography fractions were purified on a Waters

Alliance HPLC with a Phenomonex Luna C18 column (100 Å, 5 μm, 4.6x250mm). The

following 30-minute linear solvent gradient of water/0.1% formic acid (solvent A) and

acetonitrile/0.1% formic acid was used: hold at 5% B for 2 minutes, followed by linear increase

until 95% B at 20 minutes, hold at 95% B until 25 minutes, then return to 5% B until 30 minutes.

The collected peak fractions were dried by Genevac evaporator and tested for bioactivity.

5.2.3 Genomic studies

5.2.3.1 Genomic DNA extraction

For taxonomic identification, genomic DNA was extracted using the DNeasy Blood and Tissue

Kit (Qiagen). For genomic DNA extraction, each strain was grown in an overnight 5mL liquid

culture using the medium from which the strain was first originally isolated. Following the

overnight incubation, the cells were harvested, lysed and the genomic DNA was purified

according to the DNeasy Blood and Tissue Kit protocol for DNA extraction from Gram-positive

bacteria.

For Pacific Biosciences sequencing of Marinobacter sp. N33, genomic DNA was extracted using

the Genomic-tip 20/G Kit (Qiagen). 10mL liquid overnight cultures of Difco Marine Agar 2216

medium (5.2.1.2) were used for genomic DNA extraction. Following the overnight liquid culture

incubation, the cells were harvested and the DNA was extracted and purified according to the

Genomic-tip 20/G protocol for the extraction of DNA from Gram-negative bacteria. In order to

minimize DNA shearing, the sample was handled with care, the DNA vortexing steps were

skipped, and only wide-bore pipette tips were used. The genomic DNA was resuspended to a

concentration of in 10 mM Tris·Cl, pH 8.5 and was shipped on ice to Genome Quebec where the

sample was sequenced using a Pacific Biosciences RSII SMRT sequencer.

5.2.3.2 Phylogenetic analysis

Using the extracted genomic DNA, the 16S rRNA gene of each member of the maritime strain

collection was amplified by PCR using 27F and 1492R universal primers. The amplified

sequence was cloned into the pGEM-T Easy vector (Promega) and transformed in E. coli TOP10

for plasmid propagation. Plasmids were harvested and purified using the QIAprep Spin Miniprep

Page 75: Heterologous expression and characterization of the

66

Kit (Qiagen) and were sequenced by Sanger sequencing. The 16S rRNA gene sequences were

then aligned using the ClustalW multiple sequence alignment. Phylogenetic trees were

constructed by maximum likelihood method using the MEGA 7 program141.

5.2.3.3 Whole genome sequencing of Marinobacter sp. N33

Whole genome sequencing of Marinobacter sp. N33 was performed using the long-read

sequencing technology PacBio SMRT sequencing using the RSII sequencer with one SMRT cell.

The Marinobacter sp. N33 genome was assembled by Genome Quebec using the Hierarchical

Genome Assembly Process 2.2.0. Genome annotation was performed using two annotation

pipelines: DDBJ Fast Annotation and Submission Tool (DFAST) and Rapid Annotation using

Subsystem Technology (RAST)142,143.

5.3 Results and discussion

5.3.1 Screen of marine bacteria

5.3.1.1 Phylogenetic analysis

Phylogenetic analysis was used a means to prioritize strains for further study for the discovery of

novel antibacterial metabolites. As Actinomycetes tend to be more prolific producers of

secondary metabolites, and bacterial chemotype tends to follow phylotype, I reasoned that

studying rare Actinomycetes would result in a higher discovery rate of novel antibacterial

metabolites128. Thus, the taxonomic identity of each Nodwell Maritime Collection strain was

determined using the PCR-amplified 16S gene sequences of each strain (Table 5.1). BLAST was

used to determine the top-hit taxon and strain, and the percent similarity of the 16S gene to that

of the top-hit taxon was recorded. The phylogenetic identity of fungal strains was not pursued.

Considering that 98.65% 16S similarity is commonly used as the cutoff for the definition of a

new species, among the Maritime Collection there were several strains in the collection that may

be potential novel species144. Of note, N31 had a 16S percent similarity of 96.2% to the top-hit

taxon of Pontibacter ummariensis, indicating that this strain is almost certainly a novel taxon.

Several other strains had 16S percent similarity values very close to the cut off and therefore

further chemical characterization is required to determine if the strain is a novel species.

Page 76: Heterologous expression and characterization of the

67

Table 5.1. Identity and characteristics of maritime strains

NMarC Sediment Source Top-hit taxon name Top-hit strain % Identity

1 Brackley Beach, PEI Erythrobacter seohaensis SW-135(T) 100

2 Brackley Beach, PEI Celeribacter halophilus ZXM137(T) 99.93

3 Brackley Beach, PEI Streptomyces coelescens DSM 40421(T) 99.86

4 Brackley Beach, PEI Bacillus algicola KMM 3737(T) 99.66

5 Fisherman's Cove, NS Paracoccus seriniphilus DSM 14827(T) 98.05

6 Fisherman's Cove, NS Streptomyces venezuelae ATCC 10712(T) 99.86

7 Fisherman's Cove, NS Streptomyces venezuelae ATCC 10712(T) 99.93

8 Fisherman's Cove, NS Sphingomonas paucimobilis NBRC 13935(T) 99.79

9 Fisherman's Cove, NS Streptomyces venezuelae ATCC 10712(T) 99.86

10 Fisherman's Cove, NS Bacillus hwajinpoensis SW-72(T) 99.46

11 Fisherman's Cove, NS Cobetia marina DSM 4741(T) 99.93

12 Fisherman's Cove, NS Streptomyces gancidicus NBRC 15412(T) 99.86

13 Fisherman's Cove, NS Cobetia marina DSM 4741(T) 99.93

14 Crystal Crescent Beach,

NS Streptomyces coelescens DSM 40421(T) 99.65

15 Crystal Crescent Beach,

NS Bacillus tequilensis KCTC 13622(T) 99.73

16 Hopewell Rocks, NB Pseudoalteromonas rubra ATCC 29570(T) 98.66

17 Hopewell Rocks, NB Bacillus algicola KMM 3737(T) 100

18 Hopewell Rocks, NB Pseudoalteromonas rubra ATCC 29570(T) 98.73

19 Hopewell Rocks, NB Streptomyces venezuelae ATCC 10712(T) 99.65

20 Hopewell Rocks, NB Fungal

21 Hopewell Rocks, NB Altererythrobacter sp.*

22 Hopewell Rocks, NB Bacillus zhangzhouensis DW5-4(T) 99.66

23 Fisherman's Cove, NS Streptomyces venezuelae ATCC 10712(T) 99.86

24 Cavendish Beach, PEI Pseudoalteromonas rubra ATCC 29570(T) 98.8

25 Cavendish Beach, PEI Kribbella hippodromi S1.4(T) 98.78

26 Cavendish Beach, PEI Streptomyces venezuelae ATCC 10712(T) 99.93

27 Cavendish Beach, PEI Streptomyces gibsonii NRRL B-1335(T)

99.72

28 Cavendish Beach, PEI Microbacterium aquimaris JS54-2(T) 99.79

29 Hopewell Rocks, NB Altererythrobacter ishigakiensis ATCC BAA-

2084(T) 98.72

30 Brackley Beach, PEI Streptomyces gancidicus NBRC 15412(T) 99.58

31 Brackley Beach, PEI Pontibacter ummariensis NKM1(T) 96.19

Page 77: Heterologous expression and characterization of the

68

32 Fisherman's Cove, NS Streptomyces venezuelae ATCC

10712(T) 99.72

33 Fisherman's Cove, NS Marinobacter litoralis SW-45(T) 99.86

34 Fisherman's Cove, NS Streptomyces venezuelae ATCC 10712(T) 99.93

35 Cavendish Beach, PEI Bacillus mesophilus SA4(T) 98.15

36 Cavendish Beach, PEI Streptomyces venezuelae ATCC 10712(T) 99.72

37 Brackley Beach, PEI Labrenzia alba CECT 5094(T) 98.93

38 Crystal Crescent Beach,

NS Fungal

39 Hopewell Rocks, NB Vibrio diabolicus HE800(T) 99.54

40 Crystal Crescent Beach,

NS Bacillus tequilensis KCTC 13622(T) 99.86

41 Brackley Beach, PEI Fungal

42 Fisherman's Cove, NS Streptomyces venezuelae ATCC 10712(T) 99.93

*Incomplete coverage of 16S rRNA gene sequence

To understand the composition of the Nodwell Maritime Collection, a phylogenetic tree was

constructed using the 16S rRNA gene sequences (Figure 5.1). Most NMarC strains were found

to be Actinobacteria or Proteobacteria (17 and 13 strains out of 42, respectively). This bias

towards Actinobacteria or Proteobacteria is likely reflective of the fact that many bacterial taxa

are unculturable or very difficult to culture under standard laboratory conditions which may lack

specific growth factors required for growth145.

Many of the strains were very closely related strains from the Streptomyces, which are not easily

differentiated based solely on the 16S rRNA gene sequence. Differentiation between strains of

Streptomyces often requires multi-locus sequence typing to provide sufficient resolution.

Page 78: Heterologous expression and characterization of the

69

Figure 5.1. Phylogenetic tree of Nodwell Maritime Collection strains.

Page 79: Heterologous expression and characterization of the

70

5.3.1.2 Bioactivity screening

The extracts of each Maritime strain, cultured on a four different media types, were screened for

antimicrobial activity against M. luteus, B. subtilis, E. coli and S. cerevisiae (Figure 5.3). This

screen was performed with the help of 5 other students: Jan Falguera, Jethro Prinston, Victoria

Riccio, Bilyana Ivanova and Brian Hicks. Nearly 500 conditions were screened, with each

condition representing a different NMarC strain, culture condition, or test strain. Conditions that

were considered hits were those that had OD values 50% lower than the culture controls. The

majority of conditions tested resulted in less than 50% increase or reduction in OD values with

respect to the controls (Figure 5.3). Some extracts appear to have enhanced growth of the test

organism, perhaps providing additional nutrients for growth. Many of the extracts that improved

growth of the test organism were tested against S. cerevisiae, suggesting that no antifungal

compounds were present in those extracts. In total, there were 69 hits among the conditions

screened, many of which were against M. luteus or B. subtilis.

Figure 5.2. Screen of Nodwell Maritime Collection strains. Hit conditions, highlighted in

blue, are those with Y values lower than -1, indicating a 50% reduction in optical density with

respect to the culture control.

Page 80: Heterologous expression and characterization of the

71

Figure 5.3. Distribution of growth inhibition values.

When the hits are analyzed based on phylotype of the NMarC organism, as expected, most of the

hits are from actinobacterial extracts (Figure 5.4). There was one reproducible hit from a non-

actinobacterial strain: strain N33, for which the top-hit taxon is Marinobacter litoralis (Table

5.1). Marinobacter sp. N33 had antibacterial activity against M. luteus and B. subtilis when

grown on either ISP2 medium or M1 medium. Although some species of Marinobacter are

reported to have antibacterial activity, no antibacterial metabolites have thus far been isolated

from a Marinobacter species146. Therefore, I chose to purify and characterize the metabolite(s)

produced by Marinobacter sp. N33 that are responsible for the antibacterial activity of the strain,

in hopes that the metabolite(s) have not yet been reported.

Page 81: Heterologous expression and characterization of the

72

Figure 5.4. Phylogenetic tree and antibacterial activity of Nodwell Maritime Collection

strains tested against B. subtilis. A 50% reduction of OD is considered a hit, depicted in red.

Extracts were also tested against M. luteus, E. coli, and S. cerevisiae (data included in Figure 5.2

and Figure 5.3).

5.3.2 Marinobacter sp. N33 extract testing

5.3.2.1 Bioactivity testing

The Marinobacter sp. N33 extracts was retested for antibacterial activity against B. subtilis and

M. luteus by disk diffusion assay (Figure 5.5). The antibacterial activity of Marinobacter sp.

N33 is substantially more pronounced when the organism is grown on ISP2 medium compared to

M1 medium. The extract also had antibacterial activity against M. luteus but not against E. coli,

consistent with the findings of the screen by broth microtiter dilution assay (5.3.2.1).

Page 82: Heterologous expression and characterization of the

73

Figure 5.5. Disk diffusion assay of Marinobacter sp. N33 crude extract using B. subtilis

JH642. The extract of Marinobacter sp. N33 grown on ISP2 medium has a larger zone of

inhibition than when grown on M1 medium.

5.3.2.2 Genome sequencing and genome mining

To investigate the antibacterial activity of Marinobacter sp. N33, the genome was sequenced

using Pacific Biosciences Single-Molecule Real-Time sequencing (SMRT). SMRT sequencing

was chosen because this sequencing technology produces long reads which facilitates the

assembly of a complete genome. Since secondary metabolite gene clusters can be large or

contain repetitive DNA sequences, long-read sequencing aid the complete and accurate assembly

of gene clusters for secondary metabolite biosynthesis.

The Marinobacter sp. N33 DNA reads were assembled into a single 3.4 Mbp contig of 54% GC

content with an average read length of 12,600 bases and 204X coverage. The contig had

overlapping ends, resulting in a closed circular genome, which is in accordance with the

genomes of other species of Marinobacter147,148 (Figure 5.6). Although other contigs were

assembled from the sequence reads, it is likely that these remaining contigs are due to minor

DNA contamination as the read coverage was low and the contigs had low sequence similarity to

Marinobacter sequences. As a result, no plasmids were identified from the Pacific Biosciences

sequencing run.

N33 Extract

ISP2 Media

N33 Extract

M1 Media

DMSO

Neg Control

Page 83: Heterologous expression and characterization of the

74

Figure 5.6. Circular representation of the Marinobacter sp. N33 genome. The circular tracks

from the outside inward are: Tracks 1 and 2 depict protein-coding genes on the forward and

reverse sequences, respectively; Track 3 depicts tRNA genes; Track 4 depicts rRNA genes;

Track 5 depicts GC content; Track 6 depicts GC skew [(G−C)/(G+C)].

The results from genome annotation using the RAST program were in highly similar to those

obtained using DFAST. Overall, the DFAST annotation program predicted 3184 genes, with a

mean length of 981 bp, similar to the genome statistics of close relatives (Table 5.2).

Page 84: Heterologous expression and characterization of the

75

Table 5.2. Comparison of Marinobacter sp. N33 genome statistics to close relatives.

Marinobacter sp. N33 M. excellens HL-55 M. vinifirmus FB1

Length (Mbp) 3.4 4.0 3.8

GC Content (%) 54.1 56.3 58.0

# Genes 3184 3670 3522

Mean Gene Length

(bp) 981 995 986

BGCs

(AntiSMASH) 2 1 1

To investigate the ability of Marinobacter sp. N33 to produce antibacterial secondary

metabolites, I analyzed the complete genome sequence using several genome mining tools for

secondary metabolite detection (). Using AntiSMASH 4.0 and PRISM, only two secondary

metabolite gene clusters were identified in the Marinobacter sp. N33 genome: the gene cluster

for the synthesis of ectoine, an osmoprotectant and the cluster for the synthesis of a siderophore,

likely a marinobactin, a class of siderophores isolated from members of the Marinobacter

genus149. Neither ectoine nor a siderophore are expected to have antibacterial activity, suggesting

that AntiSMASH 4.0 and PRISM were not detecting the gene cluster of the antibacterial

metabolite. Recently, AntiSMASH 5.0 has been released and is able to detect more secondary

metabolite gene clusters46. Analyzing the Marinobacter sp. N33 genome sequence by

AntiSMASH 5.0 reveals the presence of two additional biosynthetic gene clusters not detected

by AntiSMASH 4.0. Both of these predicted clusters belong to the beta-lactones, a diverse class

of natural products which includes members with antibacterial activity150. It is possible that the

antibacterial metabolite(s) produced by Marinobacter sp. N33 are beta-lactones and went

undetected due to the detection limits of AntiSMASH 4.0. This would also explain the difficulty

in purifying the active metabolite(s) using formic acid (5.3.2.3), as beta-lactones are highly prone

to acid hydrolysis and thermal degradation150.

5.3.2.3 Bioactivity guided fractionation and purification

Unable to initially glean any information about the class of antibacterial secondary metabolite

from genome mining tools, I focused on purification and characterization of the bioactive

Page 85: Heterologous expression and characterization of the

76

metabolite. Following initial preparative purification by flash chromatography, the bioactivity of

each collected fraction was assessed and the bioactive fractions 17, 18 and 19 were further

purified via HPLC (Figure 5.7). Several peaks were collected from each HPLC run, although

none of the collected peaks had antibacterial activity. This may be because the antibacterial

metabolite is unstable at high temperatures or in the presence of acid (which was added to each

solvent during the HPLC purification but not during the flash chromatography purification).

Figure 5.7. UV chromatogram of the purification of fraction 19 from flash chromatography

by HPLC. The baseline drift is due to the percent composition of acetonitrile increasing over

the course of the run. The displayed UV wavelength is 220nm.

Page 86: Heterologous expression and characterization of the

77

Copyright Acknowledgements

Explicit copyright permission is not required for reproduction of the following articles in this

thesis:

Tan, S., Moore, G. & Nodwell, J. Put a Bow on It: Knotted Antibiotics Take Center Stage.

Antibiotics 8, 117 (2019).