© 2013 neha garg

187
© 2013 Neha Garg

Upload: vuongdieu

Post on 04-Jan-2017

219 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: © 2013 Neha Garg

© 2013 Neha Garg

Page 2: © 2013 Neha Garg

EXPLORING AND UNDERSTANDING LANTIBIOTIC BIOSYNTHESIS

BY

NEHA GARG

DISSERTATION

Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Biochemistry

in the Graduate College of the University of Illinois at Urbana-Champaign, 2013

Urbana, Illinois

Doctoral Committee: Professor Wilfred A. van der Donk, Chair, Director of Research Professor Satish K. Nair, Co-director Professor Susan A. Martinis Professor Huimin Zhao

Page 3: © 2013 Neha Garg

 

ii  

Abstract1,2

The ribosome translates genomic information into structural and catalytic protein

molecules. In addition to its role in protein synthesis, the ribosome also produces small non-

catalytic peptides that exert antimicrobial properties. The recent surge in the availability of

genome sequences from a wide variety of bacterial sources has led to the rapid discovery of

ribosomally-synthesized and post-translationally modified peptides (RiPPs). The biosynthetic

machinery of RiPPs offers a new platform for the discovery and engineering of peptidic natural

products. Although rapid advances have been made in elucidating biosynthetic pathways

generating various RiPPs, investigations with respect to enzyme chemistry and action are still in

their nascent stage. The novel chemistry via which post translational modifications (PTMs) are

introduced into RiPPs emphasizes the synthetic potential found in nature. One such class of

RiPPs are the lantibiotics, lanthipeptides with antimicrobial activity. The unifying PTMs found

in lanthipeptides are thioether crosslinks, called lanthionine and methyl lanthionine. The

crosslinks are formed via the dehydration of serine and threonine residues to dehydroalanine

(Dha) and dehydrobutyrine (Dhb) residues, respectively, followed by Michael-type addition of

cysteines to the dehydrated residues. Nisin, a class I lantibiotic produced by Lactococcus lactis,

is one of the oldest known antibiotics. In this study, we have developed a production strategy for

class I lantibiotics in Escherichia coli and show successful production of nisin as a proof of

concept.

1Parts of this chapter have been adapted with permission from:

1. Garg N, Tang W, Goto Y, Nair SK, van der Donk WA,. Proc Natl Acad Sci U S A. 2012 Apr 3; 109(14):5241-6. Reproduced with permission of the publisher and available from http://www.pnas.org/ (DOI: 10.1073/pnas.1116815109).

2. Garg N, Salazar-Ocampo LMA, van der Donk WA. Proc Natl Acad Sci U S A. 2013 Apr 30;110(18):7258-63. Reproduced with permission of the publisher and available from http://www.pnas.org/ (DOI: 10.1073/pnas.1222488110).

Page 4: © 2013 Neha Garg

iii  

The dehydration reaction in class I lantibiotics, thiopeptides, and goadsporin is catalyzed

by LanB or LanB-like proteins. Although LanB proteins have been studied since 1992, in vitro

reconstitution of their dehydration activity has been elusive. Herein, we demonstrate the in vitro

activity of NisB, the dehydratase involved in the biosynthesis of nisin. NisB requires glutamate,

adenosine-5’-triphosphate, Mg2+, and the ribosomal/membrane fraction of bacterial cell extract

to dehydrate its substrate peptide NisA. Mutation of 23 highly conserved residues of NisB

identified a number of amino acids that are essential for dehydration activity. In addition, these

mutagenesis studies identified three mutants, R786A, R826A, and H961A that result in multiple

glutamylations of the NisA substrate. Glutamylation was observed both during E. coli co-

expression of NisA with these mutants and during in vitro assays. Treatment of the glutamylated

substrate with wild type NisB resulted in dehydrated NisA suggesting that the glutamylated

peptide is an intermediate in dehydration. Collectively, these studies suggest that dehydration

involves glutamylation of the side chains of Ser and Thr followed by elimination. The proposed

reaction mechanism is unprecedented for lanthipeptide dehydration and these findings facilitate

mechanistic investigations for other LanB proteins that are involved in the biosynthesis of

lantibiotics, thiopeptides, and goadsporin.

Nisin is used extensively for food preservation and is also FDA-approved for the

treatment of bovine mastitis. Despite its commercial use for more than 55 years, no evidence of

significant resistance to nisin has emerged. Unfortunately, the stability of nisin is limited at pH

7.0 and elevated temperatures, which hinders its broader application. Conceivably, lanthipeptides

from thermophilic bacteria, or bacteria which live in harsher ecological niches than L. lactis,

might not suffer from the limitations of nisin. In this study, using genome mining, we discovered

two novel lantibiotics, geobacillin I and geobacillin II, from a thermophilic bacterium

Page 5: © 2013 Neha Garg

iv  

Geobacillus thermodenitrificans NG80-2 that has an optimal growth temperature of 65 °C. The

heterologous production system that was developed for nisin in E. coli enabled their production,

and biochemical and structural characterization. Geobacillin I is a class I lantibiotic and

resembles nisin in structure, but it contains seven thioether crosslinks, two more than nisin and

the most crosslinks found in any lantibiotic to date. Compared to nisin, geobacillin I displayed

increased activity against Streptococcus dysgalactiae, one of the causative agents of bovine

mastitis. Geobacillin I also demonstrated increased stability compared to nisin A at temperatures

of 37 °C and 60 °C at pH 7.0. Despite the lack of a hinge region that has been reported as

important for the pore-formation activity of nisin, geobacillin I was found to form pores in the

membranes of Gram-positive bacteria. Geobacillin II is a class II lantibiotic, and was found to

have a ring topology unlike any known lantibiotic, as determined by tandem mass spectrometry.

It contains unusual stereochemistry at the first lanthionine ring. Interestingly, geobacillin II has a

narrow spectrum of antibacterial action as it only demonstrates antimicrobial activity against

Bacillus strains amongst the bacteria tested. The activity of the class II lanthionine synthetase

catalyzing the PTMs in geobacillin II was reconstituted in vitro and the enzyme was found to be

active at temperatures up to 80 °C. Seven Geobacillus strains were screened for production of the

geobacillins using whole cell MALDI-MS and five were shown to produce geobacillin I; none

produced geobacillin II.

In summary, a strategy was developed for the production of class I lantibiotics in E. coli.

Using this methodology, two novel lantibiotics, geobacillin I and geobacillin II were produced

and characterized. Further, NisB-catalyzed dehydration and GeoM-catalyzed

dehydration/cyclization was investigated in vitro and provided insights into the mechanism of

catalysis.

Page 6: © 2013 Neha Garg

v  

To my father Sh. Shiv Raj Garg

Page 7: © 2013 Neha Garg

vi  

Acknowledgements

I would like to express my deepest appreciation to all those who provided me with the

possibility of completing my doctoral degree in the Biochemistry department at the University of

Illinois at Urbana-Champaign. First, I would like to acknowledge with much gratitude the crucial

and important role of my mentor, Professor Wilfred A. van der Donk. Without his scientific

acumen, his farsightedness, and persistent help, this dissertation would not have been possible.

Apart from constantly improving the budding scientist in me, he has been my English teacher

and has been very generous in sharing his rich and valuable wisdom. I can only hope to acquire a

small percentage of his patience, his sophisticated manner and his boundless dedication to

mentoring. It has been a great honor to work under his guidance.

I am grateful to my co-advisor Professor Satish K. Nair for his useful comments,

unreserved help, and engagement through the learning process of graduate school. I am

extremely fortunate to have had access to the knowledge and expertise of the Nair laboratory

which has gone a long way toward the successful completion of my work. I would also like to

thank my committee members Professor Susan A. Martinis and Professor Huimin Zhao for their

support and encouragement throughout my degree process.

A good support system is very important to survive and stay sane in graduate school. In

this regard, special thanks to Martha Freeland and Nancy Holda for administrative assistance and

life lessons. My most sincere thanks to Jeff Goldberg, for guiding me through the graduate

school requirements and procedures. I appreciate the hard work and professionalism of all the

staff members of the Biochemistry department and the university in taking care of the needs of

the students.

Page 8: © 2013 Neha Garg

vii  

I am grateful to all the past and present members of van der land. Special recognition and

thanks to Dr. Yuki Goto, Dr. Juan Esteban Veláquez, and Dr. Trent Oman for getting me started

on the wonderful journey in the van der land and for their support in my projects. The experience

of graduate school would not have been so full of pleasure without the love, support, and

friendship of my girl friends Nancy, Ayşe, Xiao, and Min. Thank you Pat, John, and Noah for

your friendship and helpful discussions about anything and everything. Thank you Lindsey for

helping me stay calm during my qualifiers. I appreciate and thank Weixin for the endless supply

of food, chocolate and for contributing to my research. Thank you Despina for making me little

Ms. Marie Curie. I am going to miss your humor. I am very thankful to my undergrad Ting Chen

for keeping me sane in my final semester and for her very cheerful and positive attitude. Thanks

to Chantal and Manny for making the lab a very happy place and spreading the good cheer.

Thank you Subho for helping me with chemistry questions. Thank you Xiling for helping me out

with late-night experiments.

Thank you Gabby for your friendship, love, and for putting up with me for the last three

years. Subgroups would not have been so much fun if it was not for you. I am indebted to Becca

for her help and expertise with endnote, for proofreading and improving my writing. I have a

learnt a lot from your very accepting, relaxed, and positive outlook towards life.

I want to thank Furong Sun, Haijun Yao, Kevin Ryan Tucker from mass spectrometry

facility, Alexander Vladimirovich Ulanov from metabolomics facility, and Dean Olson from

NMR facility for training me and keeping all instrumentation in order.

I would also like to thank my college friends Nikhil, KDR, Kshitij, Livee, Shruti, Shweta,

Sudeepa, and Shruti for loving me so much, for being there for me always, for showing me the

light when needed and for your continued friendship. A special mention to Anika Jain and Ankur

Page 9: © 2013 Neha Garg

viii  

Jain for loving me, supporting me, feeding me and making my time in Urbana-Champaign truly

wonderful.

I am eternally indebted to my mother Veena Garg for her love and for all the sacrifices

she made for me and for my sister ensuring our success in all endeavors of life. Last but not

least, my deepest gratitude to my friend since college and my better half, Vinayak. We have

grown together as individuals and as scientists. There is so much to learn from you that I believe

one life time is not enough. Thank you for your infinite love, care and for giving meaning to my

life.

Page 10: © 2013 Neha Garg

ix  

Table of contents

Chapter 1: Introduction……………………………………………………………………………1

Chapter 2: In vitro reconstitution, biochemical, and mechanistic characterization of the nisin

dehydratase NisB………………………………………………………………………………...26

Chapter 3: Discovery, heterologous production, and characterization of a novel type I lantibiotic

geobacillin I……………………………………………………………………………………...77

Chapter 4: Discovery, heterologous production, and characterization of geobacillin II, and in

vitro reconstitution of its synthetase enzyme GeoM……………………..……………………..115

Appendix: Structural characterization of lantibiotic dehydrogenase ElxO…………………….167

Page 11: © 2013 Neha Garg

1  

Chapter 1: Introduction

1.1 Ribosomally-synthesized and post-translationally modified peptides

Natural products have a diverse variety of applications, such as their use as antibiotics

(1), anticancer agents (2), sunscreens (3), and are produced by all domains of life. They often

possess a biological activity that can be exploited for pharmaceutical drug design and discovery.

Natural products usually have more hetero-atoms, more stereogenic centers, and more fused

rings than synthetic small molecules. In some cases, they also satisfy Lipinski’s “rule of five” (4-

5). Thus, natural products possess interesting properties from a pharmaceutical standpoint. The

golden era of antibiotic natural product discovery was the 1940’s and 1950’s. Statistical analysis

performed by Watve et al. in 2001 predicted that the exclusive screening of Streptomyces would

result in the discovery of 15-20 new antibiotics each year for the next 50 years (6). However,

slow discovery rates because of the potential for rediscovery were projected to be economically

unviable and hence, pharmaceutical companies lost interest in natural product discovery.

Recently, the explosion in availability of genome sequences, development of high-throughput

screening methods, novel culture methods, heterologous DNA-based methods, and strategies

such as metagenomics have led to renewed interest in antibiotic discovery (7). Fanning the fire is

the growing emergence of antibiotic-resistant infectious bacterial strains. The emergence of

vancomycin-resistant Enterococci (VRE) and methicillin-resistant Staphylococcus aureus

(MRSA) (8-9), have highlighted the importance of developing novel pharmacophores towards

which the odds of developing resistance would be very low. One approach towards this goal

involves identification of antibiotics with multiple modes of actions such that it would be

difficult for bacteria to reprogram multiple biological targets to develop resistance. This

approach would require designing new screens and possibly exploring untouched habitats and

Page 12: © 2013 Neha Garg

2  

classes of microorganisms. Efforts directed towards developing new scaffolds rather than re-

engineering existing ones may provide a long term solution to the problem of emerging

resistance against antibiotics. One such promising opportunity is provided by a relatively new

class of natural peptide antibiotics termed ribosomally-synthesized and post-translationally

modified peptides (RiPPs). In order to survive in their niche and obtain nutrients, bacteria have

to constantly fight against competing bacteria. To achieve this, bacteria produce a number of

small molecules including RiPPs. The in silico studies made possible by the recent surge in

availability of genome sequences has revealed that RiPPs are much more widespread than

previously suggested (10-13). The post-translational modifications expand the chemical space of

ribosomal peptides, which is typically restricted by the use of only twenty naturally occurring

amino acids, and provide these molecules with restricted conformations for tighter target

binding, and enhanced biological and chemical stability.

Interestingly, RiPPs are produced by all domains of life including humans (14) and

possess vast structural and functional diversity. The various biological activities possessed by

these molecules include antimicrobial, antitumor, antiviral, anti-inflammatory, and

morphological activities. Despite the structural and functional diversity and the difference in

sources of origin, the biosynthetic logic leading to generation of RiPPs has been conserved

among all classes of these molecules. RiPPs are smaller than 10 kDa in size and range in amino

acid length between 20 and 110 amino acid residues. The ribosomally-synthesized precursor

peptide can be subdivided into various segments. Each segment has its own unique functions but

not all segments are present in the biosynthesis of a given molecule. Most RiPPs contain an N-

terminal leader sequence, which guides the enzymatic tailoring of the C-terminal core region

and/or export of the natural product (11). Typically, it is the core region that finally becomes the

Page 13: © 2013 Neha Garg

3  

bioactive molecule. Bottromycins are the only known class of RiPPs that contains a C-terminal

rather than an N-terminal leader region and hence this region is called a “follower peptide”.

Some RiPPs also contain an N-terminal signal sequence for guiding subcellular localization and

a C-terminal recognition sequence, which can be removed in processes such as macrocyclization.

The annotation of RiPPs with the associated biosynthetic pathways in recent years has

significantly enhanced the knowledge of diversity and breadth of post-translational modifications

found in these products. The biosynthetic pathways for RiPPs are amenable to in vivo or in vitro

engineering. Some of the tailoring enzymes are shared between pathways of different classes of

RiPPs, for instance, lanthipeptide dehydratases, radical-SAM methyl transferases, and

cyclodehydratases (11). Most tailoring enzymes are promiscuous and can modify core regions

with varied amino acid sequences. Thus, a non-silent change at the nucleotide level of the

precursor peptide directly translates into a change in the final RiPP produced. Such engineering

is a relatively facile process since these pathways are relatively short and a single enzyme can

carry out a number of modifications.

1.2 Lanthipeptides

Amongst the pool of known RiPPs, lanthipeptides are peptide antibiotics (15) that contain

two alanine residues crosslinked via thioether bridges called lanthionine, or thioether crosslinks

between alanine and ethylglycine termed methyllanthionine. Lanthipeptides derive their name

from the presence of the universal lanthionine motif. Lanthipeptides that possess antimicrobial

activity are termed lantibiotics (11). The first lantibiotic, nisin, was discovered in 1928, isolated

in 1947, and used as food preservative since 1951 (16-20). Although the first lanthipeptide gene

cluster was sequenced in 1988 (21), the protein-like ribosomal synthesis of nisin was first

Page 14: © 2013 Neha Garg

4  

suggested as early as 1966 (22), and its structure was elucidated in 1971 (17). Lanthipeptides are

small peptides typically 19-38 amino acids long and contain extensive post-translational

modifications (Figure 1.1). The defining modifications include 2,3-dehydroalanine (Dha), (Z)-

2,3-didehydrobutyrine (Dhb), and the thioether containing residues lanthionine (Lan) and 3-

methyllanthionine (MeLan). Apart from the lanthionine modifications, some lantibiotics also

contain other modifications such as chlorination, hydroxylation, and the formation of labionin,

D-alanine, lysinoalanine, 2-oxo-propionyl, and S-aminovinyl-D-cysteine (Figure 1.1) (23).

1.3 Biosynthesis of lanthipeptides

The thioether bridges in lanthipeptides are enzymatically installed either by the action of

two separate enzymes termed dehydratases (LanB) and cyclases (LanC) or by the action of

bifunctional enzymes (LanM, LanL, or LabKC/RamC), which carry out both dehydration and

cyclization activities. Each lanthipeptide gene cluster contains a precursor peptide LanA and at

least one of these tailoring enzymes; lanthipeptides are classified depending on which of these

tailoring enzymes is present in the biosynthetic gene cluster. Their domain architecture is

depicted in Figure 1.2. A few examples of each lanthipeptide class are shown in Figure 1.3.

Page 15: © 2013 Neha Garg

5  

Figure 1.1. The post-translational modifications found in lanthipeptides. The color scheme used for these drawings is also followed in Figure 1.3.

Page 16: © 2013 Neha Garg

6  

Figure 1.2. Schematic representation of the domain architecture of tailoring enzymes present in different classes of lanthipeptides. Individual domains are labeled. The conserved residues in pSer/pThr lyase and kinase domains are shown as light green bars. The conserved zinc ligands present in the LanC cyclase of class I, and in the LanC-like domains of class II and IV proteins are shown with dark green bars. Adapted from reference (24).

Page 17: © 2013 Neha Garg

7  

Figure 1.3. Structures of select examples of class I (nisin, and epilancin 15X), class II (actagardine, and cinnamycin), class III (labyrinthopeptin A2), and class IV (venezuelin) lanthipeptides.

The class I lanthipeptides are defined by the presence of a dedicated dehydratase

generically termed LanB and a cyclase termed LanC. A more specific designation is used for

each lanthipeptide, e.g. GeoB and GeoC for geobacillin I. In the biosynthesis of class I

lanthipeptides, leader peptide removal and transport of the final molecule is performed by a

protease LanP and an adenosine-5’-triphosphate (ATP)-binding cassette (ABC) transporter

LanT, respectively (25). The leader peptide can be proteolytically removed either intracellularly

or extracellularly. For example, the leader peptide of nisin is removed extracellularly whereas the

leader peptide of epilancin 15 X is removed intracellularly (26). Some gene clusters, such as the

gene cluster of subtilin, lack a LanP, and the leader peptide is suggested to be removed by a non-

specific protease found elsewhere in the genome (27-28). Recently, gene clusters with multiple

LanAs and LanBs have been reported (10).

Page 18: © 2013 Neha Garg

8  

Formation of the thioether bridges in class II, III and IV lanthipeptides is catalyzed by

bifunctional lanthionine synthetases. The bifunctional LanM enzymes involved in the

biosynthesis of class II lanthipeptides have an N-terminal dehydration domain that does not

display sequence homology with known enzymes in the database and a C-terminal cyclization

domain that displays homology with the LanC cyclases of the class I machinery. A bifunctional

transporter/protease, which is involved in leader peptide removal and transport, is found in most

class II lanthipeptide gene clusters. In certain cases, a second proteolysis event is required to

completely remove the leader peptide (29). Biosynthetic gene clusters with two LanAs and two

LanMs have been extensively characterized, and such lantibiotics are termed two-component

lantibiotics. Recently, gene clusters with more than two LanAs and LanMs have also been

reported (10).

The individual domains of the lanthipeptide synthetases involved in the biosynthesis of

class III (30) and class IV (24, 31) lanthipeptides are more familiar and have known counterparts

in the protein database. The dehydration reaction is achieved by phosphorylation of

serines/threonines by a central kinase domain, followed by elimination of phosphate by an N-

terminal phosphoSer/phosphoThr lyase domain (24, 31). The class III and IV lanthipeptide

synthetases differ in their C-terminal cyclase domains. Although the cyclization domains of all

four classes show significant sequence homology, the cyclization domains of class III enzymes

lack the three conserved metal binding residues that are fully conserved in class I, II and IV

enzymes (32-35). A unique cyclization reaction catalyzed by some of the class III enzymes also

has been characterized recently. It is proposed that these enzymes catalyze the formation of a

carbon-carbon crosslink by the attack of an intermediate enolate onto a second dehydroalanine.

The genes responsible for secretion and leader peptide removal are not yet characterized for class

Page 19: © 2013 Neha Garg

9  

III and class IV lanthipeptides, but genes with sequence homology to ATP transporters do occur

in the biosynthetic gene clusters (30).

1.4 Lanthipeptide leader peptides

Although the precursor peptides contain dehydratable residues in both the leader and core

region, only residues present in the core region undergo dehydration. The leader peptide varies in

length and is usually between 23-59 residues (36). Various roles have been proposed for the

leader peptides of RiPPs. The leader peptide of nisin has been demonstrated to play important

roles in maturation, transport, and proteolysis (37). A recent study demonstrated that the NisA

leader was not sufficient to pull down the NisB/C complex suggesting that the tailoring enzymes

NisB and NisC recognize both leader and core (38). The binding studies between NisB and

different segments of modified and unmodified precursor peptide demonstrated that NisB had

greater affinity for dehydrated prenisin but did not bind modified nisin lacking the leader region

(39).

1.5 Mechanism of dehydration and cyclization in lanthipeptide biosynthesis

The lanthipeptide synthetases catalyzing the biosynthesis of all four classes of

lanthipeptides described above involve dehydration and cyclization steps to install lanthionine

and methyllanthionine bridges. The LanB enzymes do not have sequence homology with known

enzymes in protein database, their structures have not been solved, in vitro activity has not been

reconstituted, and no reaction intermediates have been isolated from the in vivo studies. Hence,

the mechanism of action of these enzymes is still unknown. In contrast, the first study reporting

the in vitro reconstitution of a LanM enzyme (40-41) in 2004 shed light on the catalytic

mechanism of the dehydration domain of LanM enzymes. The synthetase LctM involved in the

Page 20: © 2013 Neha Garg

10  

biosynthesis of lacticin 481 phosphorylates the hydroxyl group of Ser/Thr in the core peptide

followed by elimination of phosphate resulting in dehydro amino acids. Extensive site directed

mutagenesis of the conserved residues of LctM was performed to gain insights into the

mechanism of dehydration (42). A base catalyzed deprotonation of the hydroxyl group of

Ser/Thr followed by attack on the γ-phosphate of ATP would produce a phosphorylated peptide

and adenosine-5’-diphosphate (ADP) (Figure 1.4). In contrast to hydrolysis of the phosphate

ester, which is a common biological reaction, base-catalyzed elimination of phosphate is

proposed to yield a dehydrated Ser/Thr. In the case of LctM, either Asp242 or Asp259 is

proposed to be the first base involved in deprotonation of Ser/Thr. Arg399 or Asp 259 could act

as second base and deprotonate the α-carbon of pSer/pThr, generating an enolate, which

subsequently eliminates the phosphate group. Lys159 and/or Mg2+ may possibly aid by acting as

general acid in the reaction.

The dehydration reaction catalyzed by LanL enzymes that belong to class IV

lanthipeptides was recently characterized (24, 31). The domain architecture of these enzymes as

described above also suggests a phosphorylation and elimination mechanism. The cyclization

domain of VenL, the synthetase involved in venezuelin production in Steptomyces venezuelae,

was deleted and the N-terminal bifunctional kinase/lyase domain was isolated (31). In vitro assay

with this domain resulted in dehydrations of Ser/Thr in the substrate peptide. Mutagenesis

studies in this domain suggested that Lys80 could deprotonate the α-carbon of pSer/pThr aided

by Lys51, which coordinates to the carbonyl group of the pSer/pThr based on the

crystallographic structure of a pThr lyase involved in virulence (43). Further, a role of His53 in

protonating the phosphate leaving group was proposed. Interestingly, the synthetase AciKC

involved in the biosynthesis of the class III lanthipeptide catenulipeptin in Catenulispora

Page 21: © 2013 Neha Garg

11  

acidiphila DSM 44928, was shown to utilize all four NTPs as co-substrates (44). Based on

sequence homology and domain architecture of this class of synthetases, the dehydration was

proposed to take place by a similar phosphorylation and elimination mechanism (44-45).

Figure 1.4. Proposed mechanism of catalysis of lanthipeptide synthetases. A) Mechanism for the dehydration of Ser by LctM (42). B) Mechanism of dehydration of Ser by VenL (24).

Page 22: © 2013 Neha Garg

12  

The lanthipeptide cyclase enzymes LanC, and the C-terminal cyclase domain of LanM,

LanL, and LanKC enzymes catalyze the regio- and stereoselective Michael-type addition of

cysteine side chains onto dehydroamino acids. The in vitro reconstitution of the cyclization

reaction, and the first and only X-ray crystal structure of a LanC enzyme involved in the

biosynthesis of a lanthipeptide was first described in 2006 for nisin (33). NisC was shown to be a

zinc-containing metalloprotein. The zinc ion was proposed to lower the pKa of thiol groups of

cysteine allowing the nucleophilic addition of activated thiols onto unsaturated amino acids and

hence was proposed to play a catalytic role. Zinc ligands such as Cys284, Cys330, and His331 in

NisC and Cys781, Cys836, and His837 in LctM are very well conserved in lanthipeptide

cyclases. Interestingly, the cyclization domain of ProcM, a lanthipeptide synthetase involved in

the cyclization of twenty nine different substrate peptides, contains three cysteines (Cys924,

Cys970, and Cys971) as zinc ligands rather than two cysteines and one histidine as found in all

other LanC enzymes and cyclase domains of LanM enzymes (46). This difference may be the

reason for the highly promiscuous cyclization activity of ProcM. With respect to catalysis, a

conserved His residue (His212 in NisC/His725 in LctM) was proposed to be the catalytic base

that deprotonates the cysteine thiol of the substrate or the catalytic acid that protonates the

enolate (35). The mechanism of the cyclization reaction of LanKC enzymes is unknown and

may be different from the mechanism of LanC, LanM, and LanL enzymes since LanKC proteins

lack the conserved zinc ligands found in these enzymes.

1.6 Mechanism of action of lanthipeptides

Lanthipeptides are known to function as antibacterials (25), signaling molecules (47),

enzyme inhibitors (48), and morphogenic peptides (49). The most studied biological function of

lanthipeptides is their antibacterial activity. Lantibiotics such as nisin, microbisporicin, epilancin

Page 23: © 2013 Neha Garg

13  

15X, and haloduracin are known to display nanomolar minimal inhibitory activities against

pathogens of clinical interest (50-52). Originally, the capability of lantibiotics to inhibit bacterial

growth was attributed to binding of these cationic peptides with the negatively charged

phospholipids in the cell wall of bacterial cells. A large amount of experimental evidence

collected in the past 15 years has shed more light onto the mechanism of action of select

lantibiotics. Nisin was the first lantibiotic whose specific target was found to be lipid II (53).

Lipid II is a membrane-anchored cell wall precursor that is essential for peptidoglycan

biosynthesis (Figure 1.5). The peptidoglycan of all bacteria consists of a polymer of alternating

amino sugars, N-acetylglucosamine (GlcNAc) and N-acetylmuramic acid (MurNAc). The glycan

polymers of GlcNAc and MurNAc chains are cross-linked by a pentapeptide typically with the

sequence L-alanyl-γ-D-glutamyl-diaminopimelyl-(or L-lysyl)-D-alanyl-D-alanine that is attached

to the MurNAc sugar (54). This crosslinking provides the cell with the required structural

rigidity and mechanical strength. The precursor, lipid II, is comprised of an undecaprenyl

membrane anchor, a pyrophosphate group, the sugars GlcNAc and MurNAc, and a pentapeptide

chain (Figure 1.5). Nisin has been used as food preservative since 1951 but significant resistance

to nisin outside of laboratory conditions has not been observed. Lipid II is biosynthesized in ten

enzymatic steps, which are highly conserved between different bacterial species. This

complexity may explain why development of resistance against nisin is more challenging. By

binding to the pyrophosphate moiety of lipid II nisin blocks the transglycosylation step wherein

peptidoglycan glycosyltransferases catalyze the polymerization of lipid II to the growing

peptidoglycan (55).

Page 24: © 2013 Neha Garg

14  

Figure 1.5. Proposed mechanism of action of the lantibiotic nisin (15, 53). The pyrophosphate of lipid II is shown with circles labeled P. The hexagons represent the sugars GlcNAc and MurNAc. The MurNAc carries a pentapeptide that is represented by circles.

Nisin also sequesters lipid II and removes it from its functional location (53).

Sequestering of lipid II by nisin also allows nisin to form transient holes in the bacterial

membrane (Figure 1.5) (56). Specifically the A and B rings of nisin bind to the pyrophosphate of

lipid II as determined by the solution NMR structure of nisin bound to a truncated lipid II variant

with only three isoprene units (57). Furthermore, the hinge region of nisin, comprised of amino

acid residues Asn20, Met21, and Lys22 (Figure 1.3), allows it the flexibility to adopt a

conformation required for the C-terminus to insert in the bacterial membrane and hence form

pores 2 nm in diameter (56, 58-59). A study using liposomes demonstrated that the presence of

lipid II in a membrane increases the pore-forming efficiency of nisin 1000-fold (53). The

significant increase of pore forming activity of nisin was demonstrated when only two lipid II

molecules per 105 phospholipid molecules greatly enhanced the release of dyes from vesicles.

Lipid II-mediated change in the orientation of nisin from parallel to perpendicular with respect to

the membrane surface (60) is believed to result in formation of a stable pore structure.

Page 25: © 2013 Neha Garg

15  

The importance of lipid II as target for antibacterials is emphasized by the fact that lipid

II is targeted by four different classes of antibacterials including glycopeptides (61). Nisin does

not show cross-reactivity with these peptides since it binds to the pyrophosphate of lipid II,

which is not recognized by other classes of lipid II-targeting antibiotics. Some lipid II targeting

lantibiotics such as actagardine and mersacidin lack pore formation activity (62-63). An

investigation of the mechanism of action of mersacidin revealed that the binding site on lipid II is

the GlcNAc sugar instead of the pyrophosphate moiety (64). An interesting mechanism of action

has been suggested for two-component lantibiotics. Studies with haloduracin have demonstrated

that the alpha component of haloduracin binds to lipid II and the beta component binds the alpha

component in complex with lipid II and results in pore formation (65). Studies have also

uncovered examples of non-lipid II-mediated biological activity of lantibiotics. One such

example is the inhibition of phospholipase A2 by cinnamycin (48). Cinnamycin binds to

phosphatidylethanolamine with 1:1 stoichiometry and hence prevents phospholipase A2-

catalyzed release of arachidonic acids. Other lanthipeptides such as SapB and SapT lack

antibacterial activity but have morphogenic activity and result in the formation of aerial hyphae

in Streptomycetes (49, 66).

1.7 Regulation and Immunity

Unlike Trojan horse antibiotics (67-68), lanthipeptides are processed to active forms by

the producer organism, which necessitates the presence of biosynthetic regulation and self-

immunity mechanisms within the producer cells to prevent self death. The gene clusters

encoding lanthipeptides code for a variety of regulatory proteins, unique to each cluster.

Biosynthesis of some lantibiotics such as nisin (69) and subtilin (70) is regulated by a two-

component regulatory system. This system comprises of a transmembrane receptor histidine

Page 26: © 2013 Neha Garg

16  

kinase (LanK) and a transcriptional response regulator (LanR). The extracellular lantibiotic can

be detected by NisK, which initiates a signal transduction cascade by autophosphorylation of a

histidine residue. Further, transfer of the phosphoryl group by LanK to an aspartate residue of

LanR promotes the binding of LanR to the promoter elements of the biosynthetic gene cluster.

The gene encoding LanK is absent from some gene clusters, for instance, from the gene cluster

of the lantibiotic epidermin. The biosynthesis of epidermin is regulated by binding of EpiQ to the

EpiA promoter (71). The gene cluster of mersacidin encodes for two LanR type proteins (MrsR1

and MrsR2) and one LanK sensor protein (MrsK) (72). The biosynthesis of lacticin 3147 and

cytolysin is regulated by repressor proteins rather than activators, termed LtnR (73) and

CylR1/R2 (74), respectively. The activation of promoters in the biosynthesis of mutacin II is

regulated by the MutR protein and currently unknown components of the growth medium (75-

76). MutR is homologous to the family of Rgg (regulator gene of glucosyltransferase)

transcription regulators. Conversely, the production of lacticin 481 is independent of regulatory

proteins and is rather activated by the acidic pH of the environment (77). Thus, regulation of the

biosynthesis of lanthipeptides is controlled by a wide variety of different regulatory mechanisms.

The study of self-immunity conferring mechanisms for lanthipeptides is in its nascent

stage. Similar to regulatory mechanisms, self-resistance mechanisms vary significantly among

lanthipeptide gene clusters. Self-immunity of some lantibiotics such as nisin, subtilin, epidermin,

mersacidin, lacticin 481, and mutacin II is provided by the expression of an ATP-dependent

transporter protein LanFEG (73, 78-81). A second immunity protein LanI, either by itself or

along with LanFEG proteins, also is encoded in certain gene clusters such as NisI in nisin, SpaI

in subtilin, and PepI in the Pep5 biosynthetic cluster (25). A more complex immunity mechanism

Page 27: © 2013 Neha Garg

17  

was recently reported for protection against epilancin 15X and may involve the expression of

three genes, namely elx1, elx2, and elx3 (26).

1.8 In vivo engineering of nisin

The biosynthetic pathways of lanthipeptides are highly evolvable and have immense

potential for engineering in laboratory settings (82). This promise is in part due to leader peptide-

guided biosynthesis. The tailoring enzymes recognize the leader peptide and by doing so can

modify core regions containing variable amino acid substitutions. Various analogues of nisin

containing mutations in the amino acid residues of the A-ring, the C-ring, and the hinge region

between the C- and D-rings have been generated, and display enhanced biological activity. All

ring B analogues generated to date have poor biological activity as compared to wild type nisin

(83). The substitution of Dha5 by Dhb in nisin Z resulted in 2-10-fold lower biological activity

against a number of strains (84), whereas replacement of Dhb2 for Dha resulted in 2-fold

enhanced biological activity (85). The substitution of amino acids at position 3, 4 and 5 in ring A

generated three mutants, KSL, KFI and VFG, that resulted in loss of immunity in the producer

organisms (83). These analogues also had enhanced biological activity against some strains of

Staphylococcus and Lactococcus. The double mutation M17Q/G18T (84) resulted in generation

of two nisin Z analogues, one with Dhb at position 18 and another with Thr at position 18. The

variant with Dhb at position 18 had biological activity comparable to wild type nisin Z. Thus, an

additional dehydroresidue at position 18 does not enhance or reduce biological activity. The

mutant with threonine at this position had 3-fold higher activity against Micrococcus flavus. The

most interesting structure-activity relationships were observed for analogues with substitutions in

the hinge region. Site-saturated mutagenesis was performed in the hinge region at positions 20,

21, and 22. Noteworthy mutations are M21V (patented as Nisin V) and K22T (patented as nisin

Page 28: © 2013 Neha Garg

18  

T) (86-87). The K22T mutant showed enhanced activity against the human and bovine pathogen

Streptococcus agalactiae. The M21V mutation resulted in enhanced activity against a wide

variety of pathogenic bacteria including MRSA, VRE, heterogeneous vancomycin-intermediate

S. aureus (hVISA), Clostridium difficile, Listeria monocytogenes, and S. agalactiae. The

mutations N20K and M21K resulted in enhanced biological activity against Gram-negative

bacteria including Shigella, Pseudomonas and Salmonella spp (88). The mutation N20P

displayed enhanced bioactivity against MRSA, S. aureus ST528, and DPC5425, but reduced

activity against S. agalactiae (56).

1.9 Summary of investigations presented in this thesis

Through the studies described herein, we have attempted to build an appreciation for the

immense untapped potential for the discovery of enzymatic chemistry involved in lanthipeptide

biosynthesis. This thesis will describe efforts towards development of technologies for the in

vivo and in vitro production of class I (chapter 2 and 3) and class II lanthipeptides (chapter 4),

and utilization of this technology to produce novel lanthipeptides, including structural and

functional characterization and investigation of structure-activity relationships. In addition, this

thesis will describe new insights into the dehydration reaction catalyzed by LanB enzymes and

demonstrate the power of genome mining for new lanthipeptides. Sequencing of microbial

genomes to discover silent gene clusters, together with metagenomic approaches to discover

lanthipeptide gene clusters within the microbial metabolome, will further enrich this field.

Page 29: © 2013 Neha Garg

19  

1.10 References

1. Newman, D. J., Cragg, G. M., and Snader, K. M. (2003) Natural products as sources of new drugs over the period 1981-2002, J Nat Prod 66, 1022-1037.

2. Kinghorn, A. D., Chin, Y. W., and Swanson, S. M. (2009) Discovery of natural product anticancer agents from biodiverse organisms, Curr Opin Drug Discov Devel 12, 189-196.

3. Balskus, E. P., and Walsh, C. T. (2008) Investigating the initial steps in the biosynthesis of cyanobacterial sunscreen scytonemin, J Am Chem Soc 130, 15260-15261.

4. Quinn, R. J., Carroll, A. R., Pham, N. B., Baron, P., Palframan, M. E., Suraweera, L., Pierens, G. K., and Muresan, S. (2008) Developing a drug-like natural product library, J Nat Prod 71, 464-468.

5. Lipinski, C. A., Lombardo, F., Dominy, B. W., and Feeney, P. J. (2001) Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings, Adv Drug Deliv Rev 46, 3-26.

6. Watve, M. G., Tickoo, R., Jog, M. M., and Bhole, B. D. (2001) How many antibiotics are produced by the genus Streptomyces?, Arch Microbiol 176, 386-390.

7. Bode, H. B., and Müller, R. (2005) The impact of bacterial genomics on natural product research, Angew Chem Int Ed Engl 44, 6828-6846.

8. Johnson, A. P., and Woodford, N. (2002) Glycopeptide-resistant Staphylococcus aureus, J Antimicrob Chemother 50, 621-623.

9. Appelbaum, P. C. (2006) The emergence of vancomycin-intermediate and vancomycin-resistant Staphylococcus aureus, Clin Microbiol Infect 12 Suppl 1, 16-23.

10. Marsh, A. J., O'Sullivan, O., Ross, R. P., Cotter, P. D., and Hill, C. (2010) In silico analysis highlights the frequency and diversity of type 1 lantibiotic gene clusters in genome sequenced bacteria, BMC Genomics 11, 679.

11. Arnison, P. G., Bibb, M. J., Bierbaum, G., Bowers, A. A., Bugni, T. S., Bulaj, G., Camarero, J. A., Campopiano, D. J., Challis, G. L., Clardy, J., Cotter, P. D., Craik, D. J., Dawson, M., Dittmann, E., Donadio, S., Dorrestein, P. C., Entian, K. D., Fischbach, M. A., Garavelli, J. S., Goransson, U., Gruber, C. W., Haft, D. H., Hemscheidt, T. K., Hertweck, C., Hill, C., Horswill, A. R., Jaspars, M., Kelly, W. L., Klinman, J. P., Kuipers, O. P., Link, A. J., Liu, W., Marahiel, M. A., Mitchell, D. A., Moll, G. N., Moore, B. S., Müller, R., Nair, S. K., Nes, I. F., Norris, G. E., Olivera, B. M., Onaka, H., Patchett, M. L., Piel, J., Reaney, M. J., Rebuffat, S., Ross, R. P., Sahl, H. G., Schmidt, E. W., Selsted, M. E., Severinov, K., Shen, B., Sivonen, K., Smith, L., Stein, T., Süssmuth, R. D., Tagg, J. R., Tang, G. L., Truman, A. W., Vederas, J. C., Walsh, C. T., Walton, J. D., Wenzel, S. C., Willey, J. M., and van der Donk, W. A. (2013) Ribosomally synthesized and post-translationally modified peptide natural products: overview and recommendations for a universal nomenclature, Nat Prod Rep 30, 108-160.

Page 30: © 2013 Neha Garg

20  

12. Wieland Brown, L. C., Acker, M. G., Clardy, J., Walsh, C. T., and Fischbach, M. A. (2009) Thirteen posttranslational modifications convert a 14-residue peptide into the antibiotic thiocillin, Proc Natl Acad Sci U S A 106, 2549-2553.

13. Melby, J. O., Nard, N. J., and Mitchell, D. A. (2011) Thiazole/oxazole-modified microcins: complex natural products from ribosomal templates, Curr Opin Chem Biol 15, 369-378.

14. Schneider, J. J., Unholzer, A., Schaller, M., Schafer-Korting, M., and Korting, H. C. (2005) Human defensins, J Mol Med (Berl) 83, 587-595.

15. Willey, J. M., and van der Donk, W. A. (2007) Lantibiotics: Peptides of Diverse Structure and Function, Annu Rev Microbiol 61, 477-501.

16. Rogers, L. A. (1928) The inhibiting effect of Streptococcus lactis on Lactobacillus bulgaricus, J Bacteriol 16, 321-325.

17. Gross, E., and Morell, J. L. (1971) The Structure of Nisin, J Am Chem Soc 93, 4634-4635.

18. Rogers, L. A., and Whittier, E. O. (1928) Limiting factors in lactic fermentation, J Bacteriol 16, 211-229.

19. Mattick, A. T., and Hirsch, A. (1947) Further observations on an inhibitory substance (nisin) from lactic streptococci, Lancet 2, 5-8.

20. Delves-Broughton, J., Blackburn, P., Evans, R. J., and Hugenholtz, J. (1996) Applications of the bacteriocin, nisin, Antonie Van Leeuwenhoek 69, 193-202.

21. Schnell, N., Entian, K. D., Schneider, U., Götz, F., Zahner, H., Kellner, R., and Jung, G. (1988) Prepeptide sequence of epidermin, a ribosomally synthesized antibiotic with four sulphide-rings, Nature 333, 276-278.

22. Hurst, A. (1966) Biosynthesis of the antibiotic nisin by whole Streptococcus lactus organisms, J Gen Microbiol 44, 209-220.

23. Knerr, P. J., and van der Donk, W. A. (2012) Discovery, biosynthesis, and engineering of lantipeptides, Annu Rev Biochem 81, 479-505.

24. Goto, Y., Ökesli, A., and van der Donk, W. A. (2011) Mechanistic studies of Ser/Thr dehydration catalyzed by a member of the LanL lanthionine synthetase family, Biochemistry 50, 891-898.

25. Chatterjee, C., Paul, M., Xie, L., and van der Donk, W. A. (2005) Biosynthesis and Mode of Action of Lantibiotics, Chem Rev 105, 633-684.

26. Velásquez, J. E., Zhang, X., and van der Donk, W. A. (2011) Biosynthesis of the antimicrobial peptide epilancin 15X and its N-terminal lactate, Chem Biol 18, 857-867.

Page 31: © 2013 Neha Garg

21  

27. Corvey, C., Stein, T., Dusterhus, S., Karas, M., and Entian, K. D. (2003) Activation of subtilin precursors by Bacillus subtilis extracellular serine proteases subtilisin (AprE), WprA, and Vpr, Biochem Biophys Res Commun 304, 48-54.

28. Chakicherla, A., and Hansen, J. N. (1995) Role of the leader and structural regions of prelantibiotic peptides as assessed by expressing nisin-subtilin chimeras in Bacillus subtilis 168, and characterization of their physical, chemical, and antimicrobial properties, J Biol Chem 270, 23533-23539.

29. McClerren, A. L., Cooper, L. E., Quan, C., Thomas, P. M., Kelleher, N. L., and van der Donk, W. A. (2006) Discovery and in vitro biosynthesis of haloduracin, a new two-component lantibiotic, Proc Natl Acad Sci USA 103, 17243-17248.

30. Müller, W. M., Schmiederer, T., Ensle, P., and Süssmuth, R. D. (2010) In vitro biosynthesis of the prepeptide of type-III lantibiotic labyrinthopeptin A2 including formation of a C-C bond as a post-translational modification, Angew Chem Int Ed Engl 49, 2436-2440.

31. Goto, Y., Li, B., Claesen, J., Shi, Y., Bibb, M. J., and van der Donk, W. A. (2010) Discovery of unique lanthionine synthetases reveals new mechanistic and evolutionary insights, PLoS Biol 8, e1000339.

32. Kodani, S., Hudson, M. E., Durrant, M. C., Buttner, M. J., Nodwell, J. R., and Willey, J. M. (2004) The SapB morphogen is a lantibiotic-like peptide derived from the product of the developmental gene ramS in Streptomyces coelicolor, Proc Natl Acad Sci U S A 101, 11448-11453.

33. Li, B., Yu, J. P., Brunzelle, J. S., Moll, G. N., van der Donk, W. A., and Nair, S. K. (2006) Structure and mechanism of the lantibiotic cyclase involved in nisin biosynthesis, Science 311, 1464-1467.

34. Okeley, N. M., Paul, M., Stasser, J. P., Blackburn, N., and van der Donk, W. A. (2003) SpaC and NisC, the cyclases involved in subtilin and nisin biosynthesis, are zinc proteins, Biochemistry 42, 13613-13624.

35. Li, B., and van der Donk, W. A. (2007) Identification of essential catalytic residues of the cyclase NisC involved in the biosynthesis of nisin, J Biol Chem 282, 21169-21175.

36. Oman, T. J., and van der Donk, W. A. (2010) Follow the leader: the use of leader peptides to guide natural product biosynthesis, Nat Chem Biol 6, 9-18.

37. Plat, A., Kluskens, L. D., Kuipers, A., Rink, R., and Moll, G. N. (2011) Requirements of the engineered leader peptide of nisin for inducing modification, export, and cleavage, Appl Environ Microbiol 77, 604-611.

38. Khusainov, R., Heils, R., Lubelski, J., Moll, G. N., and Kuipers, O. P. (2011) Determining sites of interaction between prenisin and its modification enzymes NisB and NisC, Mol Microbiol 82, 706-718.

Page 32: © 2013 Neha Garg

22  

39. Mavaro, A., Abts, A., Bakkes, P. J., Moll, G. N., Driessen, A. J., Smits, S. H., and Schmitt, L. (2011) Substrate recognition and specificity of the NisB protein, the lantibiotic dehydratase involved in nisin biosynthesis, J Biol Chem 286, 30552-30560.

40. Xie, L., Miller, L. M., Chatterjee, C., Averin, O., Kelleher, N. L., and van der Donk, W. A. (2004) Lacticin 481: in vitro reconstitution of lantibiotic synthetase activity, Science 303, 679-681.

41. Chatterjee, C., Miller, L. M., Leung, Y. L., Xie, L., Yi, M., Kelleher, N. L., and van der Donk, W. A. (2005) Lacticin 481 Synthetase Phosphorylates its Substrate during Lantibiotic Production, J Am Chem Soc 127, 15332-15333.

42. You, Y. O., and van der Donk, W. A. (2007) Mechanistic investigations of the dehydration reaction of lacticin 481 synthetase using site-directed mutagenesis, Biochemistry 46, 5991-6000.

43. Zhu, Y., Li, H., Long, C., Hu, L., Xu, H., Liu, L., Chen, S., Wang, D. C., and Shao, F. (2007) Structural insights into the enzymatic mechanism of the pathogenic MAPK phosphothreonine lyase, Mol Cell 28, 899-913.

44. Wang, H., and van der Donk, W. A. (2012) Biosynthesis of the class III lantipeptide catenulipeptin, ACS Chem Biol 7, 1529-1535.

45. Müller, W. M., Ensle, P., Krawczyk, B., and Süssmuth, R. D. (2011) Leader peptide-directed processing of labyrinthopeptin A2 precursor peptide by the modifying enzyme LabKC, Biochemistry 50, 8362-8373.

46. Tang, W., and van der Donk, W. A. (2012) Structural characterization of four prochlorosins: a novel class of lantipeptides produced by planktonic marine cyanobacteria, Biochemistry 51, 4271-4279.

47. Kuipers, O. P., Beerthuyzen, M. M., de Ruyter, P. G., Luesink, E. J., and de Vos, W. M. (1995) Autoregulation of nisin biosynthesis in Lactococcus lactis by signal transduction, J Biol Chem 270, 27299-27304.

48. Märki, F., Hanni, E., Fredenhagen, A., and van Oostrum, J. (1991) Mode of action of the lanthionine-containing peptide antibiotics duramycin, duramycin B and C, and cinnamycin as indirect inhibitors of phospholipase A2, Biochem Pharmacol 42, 2027-2035.

49. Willey, J. M., Willems, A., Kodani, S., and Nodwell, J. R. (2006) Morphogenetic surfactants and their role in the formation of aerial hyphae in Streptomyces coelicolor, Mol Microbiol 59, 731-742.

50. Ekkelenkamp, M. B., Hanssen, M., Danny Hsu, S. T., de Jong, A., Milatovic, D., Verhoef, J., and van Nuland, N. A. (2005) Isolation and structural characterization of epilancin 15X, a novel lantibiotic from a clinical strain of Staphylococcus epidermidis, FEBS Lett 579, 1917-1922.

Page 33: © 2013 Neha Garg

23  

51. Castiglione, F., Lazzarini, A., Carrano, L., Corti, E., Ciciliato, I., Gastaldo, L., Candiani, P., Losi, D., Marinelli, F., Selva, E., and Parenti, F. (2008) Determining the structure and mode of action of microbisporicin, a potent lantibiotic active against multiresistant pathogens, Chem Biol 15, 22-31.

52. Oman, T. J., and van der Donk, W. A. (2009) Insights into the Mode of Action of the Two-Peptide Lantibiotic Haloduracin, ACS Chem Biol 4, 865-874.

53. Breukink, E., Wiedemann, I., van Kraaij, C., Kuipers, O. P., Sahl, H., and de Kruijff, B. (1999) Use of the cell wall precursor lipid II by a pore-forming peptide antibiotic, Science 286, 2361-2364.

54. Schneider, T., and Sahl, H. G. (2010) Lipid II and other bactoprenol-bound cell wall precursors as drug targets, Curr Opin Investig Drugs 11, 157-164.

55. Hasper, H. E., Kramer, N. E., Smith, J. L., Hillman, J. D., Zachariah, C., Kuipers, O. P., de Kruijff, B., and Breukink, E. (2006) An alternative bactericidal mechanism of action for lantibiotic peptides that target lipid II, Science 313, 1636-1637.

56. Wiedemann, I., Breukink, E., van Kraaij, C., Kuipers, O. P., Bierbaum, G., de Kruijff, B., and Sahl, H. G. (2001) 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, 1772-1779.

57. Hsu, S. T., Breukink, E., Tischenko, E., Lutters, M. A., de Kruijff, B., Kaptein, R., Bonvin, A. M., and van Nuland, N. A. (2004) The nisin-lipid II complex reveals a pyrophosphate cage that provides a blueprint for novel antibiotics, Nat Struct Mol Biol 11, 963-967.

58. Sahl, H. G., Kordel, M., and Benz, R. (1987) Voltage-dependent depolarization of bacterial membranes and artificial lipid bilayers by the peptide antibiotic nisin, Arch Microbiol 149, 120-124.

59. Kordel, M., Schuller, F., and Sahl, H. G. (1989) Interaction of the pore forming-peptide antibiotics Pep 5, nisin and subtilin with non-energized liposomes, FEBS Lett 244, 99-102.

60. van Heusden, H. E., de Kruijff, B., and Breukink, E. (2002) Lipid II induces a transmembrane orientation of the pore-forming peptide lantibiotic nisin, Biochemistry 41, 12171-12178.

61. Izaki, K., Matsuhashi, M., and Strominger, J. L. (1968) Biosynthesis of the peptidoglycan of bacterial cell walls. 8. Peptidoglycan transpeptidase and D-alanine carboxypeptidase: penicillin-sensitive enzymatic reaction in strains of Escherichia coli, J Biol Chem 243, 3180-3192.

62. Brötz, H., Josten, M., Wiedemann, I., Schneider, U., Götz, F., Bierbaum, G., and Sahl, H.-G. (1998) Role of lipid-bound peptidoglycan precursors in the formation of pores by nisin, epidermin and other lantibiotics, Mol Microbiol 30, 317-327.

Page 34: © 2013 Neha Garg

24  

63. Brötz, H., and Sahl, H. G. (2000) New insights into the mechanism of action of lantibiotics--diverse biological effects by binding to the same molecular target, J Antimicrob Chemother 46, 1-6.

64. Brötz, H., Bierbaum, G., Leopold, K., Reynolds, P. E., and Sahl, H. G. (1998) The lantibiotic mersacidin inhibits peptidoglycan synthesis by targeting lipid II, Antimicrob Agents Chemother 42, 154-160.

65. Oman, T. J., Lupoli, T. J., Wang, T. S., Kahne, D., Walker, S., and van der Donk, W. A. (2011) Haloduracin alpha binds the peptidoglycan precursor lipid II with 2:1 stoichiometry, J Am Chem Soc 133, 17544-17547.

66. Kodani, S., Lodato, M. A., Durrant, M. C., Picart, F., and Willey, J. M. (2005) SapT, a lanthionine-containing peptide involved in aerial hyphae formation in the streptomycetes, Mol Microbiol 58, 1368-1380.

67. Stefanska, A. L., Fulston, M., Houge-Frydrych, C. S., Jones, J. J., and Warr, S. R. (2000) A potent seryl tRNA synthetase inhibitor SB-217452 isolated from a Streptomyces species, J Antibiot (Tokyo) 53, 1346-1353.

68. Reader, J. S., Ordoukhanian, P. T., Kim, J. G., de Crecy-Lagard, V., Hwang, I., Farrand, S., and Schimmel, P. (2005) Major biocontrol of plant tumors targets tRNA synthetase, Science 309, 1533.

69. Engelke, G., Gutowski-Eckel, Z., Kiesau, P., Siegers, K., Hammelmann, M., and Entian, K.-D. (1994) Regulation of nisin biosynthesis and immunity in Lactococcus lactis 6F3, Appl Environ Microbiol 60, 814-825.

70. Klein, C., Kaletta, C., and Entian, K. D. (1993) Biosynthesis of the lantibiotic subtilin is regulated by a histidine kinase/response regulator system, Appl Environ Microbiol 59, 296-303.

71. Peschel, A., Augustin, J., Kupke, T., Stevanovic, S., and Götz, F. (1993) Regulation of epidermin biosynthetic genes by EpiQ, Mol Microbiol 9, 31-39.

72. Guder, A., Schmitter, T., Wiedemann, I., Sahl, H. G., and Bierbaum, G. (2002) Role of the single regulator MrsR1 and the two-component system MrsR2/K2 in the regulation of mersacidin production and immunity, Appl Environ Microbiol 68, 106-113.

73. McAuliffe, O., O'Keeffe, T., Hill, C., and Ross, R. P. (2001) Regulation of immunity to the two-component lantibiotic, lacticin 3147, by the transcriptional repressor LtnR, Molecular Microbiology 39, 982-993.

74. Haas, W., Shepard, B. D., and Gilmore, M. S. (2002) Two-component regulator of Enterococcus faecalis cytolysin responds to quorum-sensing autoinduction, Nature 415, 84-87.

75. Sulavik, M. C., and Clewell, D. B. (1996) Rgg is a positive transcriptional regulator of the Streptococcus gordonii gtfG gene, J Bacteriol 178, 5826-5830.

Page 35: © 2013 Neha Garg

25  

76. Sulavik, M. C., Tardif, G., and Clewell, D. B. (1992) Identification of a gene, rgg, which regulates expression of glucosyltransferase and influences the Spp phenotype of Streptococcus gordonii Challis, J Bacteriol 174, 3577-3586.

77. Hindré, T., Le Pennec, J. P., Haras, D., and Dufour, A. (2004) Regulation of lantibiotic lacticin 481 production at the transcriptional level by acid pH, FEMS Microbiol Lett 231, 291-298.

78. Siegers, K., and Entian, K. D. (1995) Genes involved in immunity to the lantibiotic nisin produced by Lactococcus lactis 6F3, Appl Environ Microbiol 61, 1082-1089.

79. Rincé, A., Dufour, A., Uguen, P., Le Pennec, J. P., and Haras, D. (1997) Characterization of the lacticin 481 operon: the Lactococcus lactis genes lctF, lctE, and lctG encode a putative ABC transporter involved in bacteriocin immunity, Appl Environ Microbiol 63, 4252-4260.

80. Peschel, A., and Götz, F. (1996) Analysis of the Staphylococcus epidermidis genes epiF, -E, and -G involved in epidermin immunity, J Bacteriol 178, 531-536.

81. Saris, P. E., Immonen, T., Reis, M., and Sahl, H. G. (1996) Immunity to lantibiotics, Antonie Van Leeuwenhoek 69, 151-159.

82. Ross, A. C., and Vederas, J. C. (2011) Fundamental functionality: recent developments in understanding the structure-activity relationships of lantibiotic peptides, J Antibiot (Tokyo) 64, 27-34.

83. Rink, R., Wierenga, J., Kuipers, A., Kluskens, L. D., Driessen, A. J., Kuipers, O. P., and Moll, G. N. (2007) Dissection and modulation of the four distinct activities of nisin by mutagenesis of rings A and B and by C-terminal truncation, Appl Environ Microbiol 73, 5809-5816.

84. Kuipers, O. P., Rollema, H. S., Yap, W. M., Boot, H. J., Siezen, R. J., and de Vos, W. M. (1992) Engineering dehydrated amino acid residues in the antimicrobial peptide nisin, J Biol Chem 267, 24340-24346.

85. Kuipers, O. P., Bierbaum, G., Ottenwalder, B., Dodd, H. M., Horn, N., Metzger, J., Kupke, T., Gnau, V., Bongers, R., van den Bogaard, P., Kosters, H., Rollema, H. S., de Vos, W. M., Siezen, R. J., Jung, G., Götz, F., Sahl, H. G., and Gasson, M. J. (1996) Protein engineering of lantibiotics, Antonie Van Leeuwenhoek 69, 161-169.

86. Field, D., Connor, P. M., Cotter, P. D., Hill, C., and Ross, R. P. (2008) The generation of nisin variants with enhanced activity against specific gram-positive pathogens, Mol Microbiol 69, 218-230.

87. Cotter, P. D., Hill, C. (2009) Nisin derivatives and the use thereof. Patent application WO 2009135945A1.

88. Yuan, J., Zhang, Z. Z., Chen, X. Z., Yang, W., and Huan, L. D. (2004) Site-directed mutagenesis of the hinge region of nisinZ and properties of nisinZ mutants, Appl Microbiol Biotechnol 64, 806-815.

Page 36: © 2013 Neha Garg

26  

Chapter 2: In vitro reconstitution, biochemical, and mechanistic characterization of the

nisin dehydratase NisB 1

2.1 Introduction

Nisin is the best characterized member of the lantibiotics family and has been used in the

food industry for over 40 years to combat food-borne pathogens (1). Nisin is biosynthesized by

NisB-catalyzed dehydration of Ser and Thr residues in the core peptide of the precursor peptide

yielding dehydroalanine (Dha) and dehydrobutyrine (Dhb) residues, respectively (Figure 2.1).

Figure 2.1. Biosynthesis of nisin. (A) NisB dehydrates Ser and Thr residues in the core peptide of NisA, and NisC catalyzes the addition of Cys to Dha and Dhb. Then NisP removes the leader peptide. Although the figure is drawn with all dehydrations being completed before cyclization commences, it is likely that dehydration and cyclization are alternating activities in lantibiotic biosynthesis (2-4). (B) Ser and Thr residues are first dehydrated to generate Dha and Dhb, respectively. Subsequent intramolecular addition of the thiols of Cys residues to the Dha and Dhb results in the formation of lanthionine (Lan) and methyllanthionine (MeLan) structures, respectively. Abu stands for 2-aminobutyric acid.

1Parts of this chapter have been adapted with permission from:

1. Shi Y, Yang X, Garg N, & van der Donk WA. J Am Chem Soc. 2011 Mar 2;133(8):2338-41. Reproduced with permission of the publisher and available from http://pubs.acs.org/journal/jacsat (DOI: 10.1021/ja109044r).

2. Neha G, Salazar-Ocampo LMA, van der Donk WA. Proc Natl Acad Sci U S A. 2013 Apr 30;110(18):7258-63. Reproduced with permission of the publisher and available from http://www.pnas.org/ (DOI: 10.1073/pnas.1222488110).

Page 37: © 2013 Neha Garg

27  

Subsequent Michael-type addition of cysteine side chains to the dehydrated residues by the

action of NisC results in the formation of lanthionine and methyllanthionine (Figure 2.1).

Engineering of nisin has been very fruitful in generating analogues with advantageous

properties and has been accomplished by manipulation of the sequence of the NisA precursor

peptide in native host strains and heterologous expression systems (5-8). To be able to take full

advantage of in vivo or in vitro introduction of non-proteinogenic amino acids, production of

nisin in Escherichia coli or reconstitution of the nisin biosynthetic pathway in vitro would be

highly desirable. As described in chapter 1, lanthipeptides have been divided into four classes

based on distinct biosynthetic enzymes that carry out the dehydration and cyclization reactions

(9). The biosyntheses of class II, III, and IV lanthipeptides have all been reconstituted in vitro

(10-12), but in vitro reconstitution of the biosynthesis of class I lanthipeptides that include nisin

has been elusive. In vivo production of lanthipeptides in E. coli has been reported for a truncated

analogue of the class II lantibiotic nukacin ISK-1(13), for bovicin HJ 50 (14), and for

lichenicidin (15). Genetic and bioengineering studies have demonstrated that NisB carries out the

dehydration reaction and does not need any other lanthipeptide biosynthetic enzymes for this

activity (16-17). The cyclization reaction catalyzed by NisC, the LanC involved in nisin

biosynthesis, has been reconstituted in vitro (18) (Figure 2.1A). However, despite considerable

effort in multiple laboratories, reconstitution of the activity of LanB dehydratases has been

unsuccessful (19-23), and hence no information is available about their mechanism of catalysis.

The LanB dehydratases are 115-120 kDa proteins that do not share homology with the

dehydratases from class II-IV lanthipeptides nor with other proteins in the databases. The

sequence identity across the LanB family is only about 30%. Genes encoding homologous LanB-

like enzymes are also present in the gene clusters of non-lantibiotic peptide natural products,

Page 38: © 2013 Neha Garg

28  

such as the thiopeptides (24-27) and goadsporin (28). The presence of dehydro-amino acids in

ribosomally synthesized natural products render the peptides proteolytically more stable and

enhance the biological activity (29). Furthermore, dehydroamino acids can be used as promising

targets for synthesis of non-natural amino acids by serving as potential sites for modification.

The organic synthesis of dehydro-amino acids is not only cumbersome but also expensive, and

involves a multistep procedure for which use of protecting groups is essential. Also, the chemical

methods are usually limited to synthesis of smaller peptides. A potential alternative route to

synthesis of dehydro-amino acids is through the biosynthetic machinery of lantibiotics. Various

medically relevant nonlantibiotic peptides fused to the nisin leader peptide have been shown to

be effectively dehydrated thus demonstrating the possible exploitation of LanB’s in production

of dehydroamino acid-containing peptides (16). Thus, in this study we undertook the task of

production of nisin in E. coli and in vitro reconstitution of the dehydratase NisB.

2.2 Experimental procedures

2.2.a Construction of pRSFDuet-1/His6-NisA+NisB and pACYCDuet-1/NisC

PCR amplification of nisA, nisB and nisC was performed using the following PCR cycle

conditions: denaturation (98 °C for 20 s), annealing (60 °C for 20 s), and extension (72 °C for 30

s for nisA, 120 s for nisB, and 60 s for nisC) using the primers listed in Table 2.1 at the end of

this chapter and genomic DNA of Lactococcus lactis 6F3 as DNA template. The PCR reaction

included 1× Phusion GC buffer (Finnzymes), DMSO (2%), Phusion hot start high fidelity DNA

polymerase (0.05 U/µL) and primers (1 µM each). The amplification products were run on a 1%

agarose gel and the products were purified using a QIAquick Gel Extraction Kit (QIAGEN). The

amplified nisA DNA fragment and the pRSFDuet-1 vector were digested with BamHI and

Page 39: © 2013 Neha Garg

29  

HindIII; PCR-amplified nisC and pACYCDuet-1 vector were digested with BglII and XhoI; and

PCR amplified nisB was digested with NdeI and XhoI at 37 °C for 12 h. The digests were run on

an agarose gel and the desired products extracted using a QIAquick Gel Extraction Kit. The

resulting DNA inserts were ligated with the respective digested vector at 4 °C for 15 h using T4

DNA ligase (Invitrogen). The ligation reaction mixtures were diluted 6-fold with water prior to

transformation. E. coli DH5α cells were transformed with the ligation product using heat shock

and were plated on Luria Bertani (LB)-kanamycin agar plates and grown at 37 °C for 15 h. Five

colonies were picked and incubated in 5 mL of LB-kanamycin medium at 37 °C for 10 h,

followed by isolation of the plasmids using a QIAprep Spin Miniprep Kit (QIAGEN). The

sequences of the resulting plasmids were confirmed by DNA sequencing using the appropriate

primers by the Biotechnology Center of the University of Illinois at Urbana-Champaign. The

plasmid vector pRSFDuet-1 containing the nisA gene was digested with NdeI and XhoI and

ligated with PCR-amplified nisB digested with NdeI and XhoI. The ligation and transformation

of E. coli DH5α cells was carried out as described above. The colonies were analyzed for insert

by colony PCR and plasmid was isolated from 5 colonies and sequenced (also see notebook 3,

page 33, 38, 45).

2.2.b Overexpression and purification of modified nisin

Chemically competent E. coli BL21(DE3) cells were co-transformed with pRSFDuet-1

containing the nisA and nisB genes and pACYCDuet-1 containing nisC. Overnight culture grown

from a single colony transformant was used as inoculum to grow 1.5 L of terrific broth (TB)

medium containing 50 mg/L kanamycin and 25 mg/L chloramphenicol at 37 °C until the OD600

reached about 0.6. The incubation temperature was then changed to 18 °C and the culture was

induced with 0.5 mM IPTG. The induced cells were shaken continually at 18 °C for an

Page 40: © 2013 Neha Garg

30  

additional 15 h. The cells were harvested by centrifugation (11,900 × g for 10 min, Beckman

JLA-10.500 rotor). The cell pellet was resuspended in 45 mL of start buffer (20 mM Tris, pH

8.0, 500 mM NaCl, 10% glycerol, containing a protease inhibitor cocktail from Roche Applied

Science) and lysed by MultiFlex C3 homogenizer (Avestin). The sample was centrifuged at

23,700 × g for 30 min at 4 °C. The supernatant was loaded onto a HiTrap HP nickel affinity

column (GE Healthcare) pre-equilibrated with start buffer. Following loading, the column was

washed with wash buffer (start buffer + 30 mM imidazole). The peptide was eluted from the

column using elution buffer (start buffer + 1 M imidazole). The pellet was homogenized using a

sonicator (35% amplitude, 4.4 s pulse, 9.9 s pause for total 20 min) in start buffer to remove any

soluble proteins. The pellet from this step was resuspended in 30 mL of denaturing buffer (6 M

guanidine hydrochloride, 20 mM NaH2PO4, 500 mM NaCl, pH 7.5). The insoluble portion was

removed by centrifugation and the supernatant was loaded onto a HiTrap HP nickel affinity

column. The column was washed with denaturing buffer containing 30 mM imidazole and eluted

with 15 mL of denaturing buffer containing 1 M imidazole. Desalting of the eluent from both

supernatant and pellet fractions was performed using preparative scale reversed phase high

performance liquid chromatography (RP-HPLC) using a Waters Delta-pak C4, 15 µm, 300 Å,

25×100 mm, PrepPak Cartridge. A gradient of 2-100% of solvent B (0.086% TFA in 80%

acetonitrile/20% water) was used (solvent A: 0.1 % TFA in water). Modified prenisin began to

elute at 52% B. The fractions containing modified prenisin were lyophilized and analyzed by

matrix-assisted laser desorption time of flight mass spectrometry (MALDI-TOF MS) (also see

notebook 5 page 3).

Page 41: © 2013 Neha Garg

31  

2.2.c Cleavage of modified prenisin with trypsin and antimicrobial activity assay

An aliquot of 100 µL of 500 µM modified NisA was incubated with 3 µL of 30 µM

trypsin (Worthington Biochemicals) at room temperature for 3 h. The resulting mixture was

checked by matrix-assisted laser desorption/ionization mass spectrometry (MALDI-TOF MS)

and the proteolytic fragment corresponding to mature nisin was observed calcd: 3354 (nisin);

found 3356 (M+H). Antimicrobial activity assays were carried out using solid agar diffusion.

GM17 agar was melted in a microwave, cooled to 45 °C and mixed with 150 µL of overnight

grown Lactococcus lactis HP (108-109 cfu). After the agar was solidified, 10 µL of 50 µM and

500 µM of heterologously expressed and trypsin cleaved NisA was spotted on the surface. As a

control, nisin purified from commercial Nisaplin (Danisco) using RP-HPLC was also spotted

(also see notebook 5 page 8).

2.2.d Tandem mass spectrometry analysis to confirm the correct ring topology of nisin

Commercial, HPLC-purified nisin and in vivo produced (E. coli) nisin were subjected to

ESI-MSMS analysis using collision induced dissociation. To carry out ESI-MSMS, 10 μL of

each nisin preparation was injected onto a BEH C8 column (1.7 μm, 1.0 × 100 mm), and

separated by UPLC using a gradient of 3% mobile phase A (0.1% formic acid in water) to 97%

mobile phase B (0.1% formic acid in methanol) over 12 min. The instrument settings used

included capillary voltage and cane voltage of 3500 V and 40 V, respectively; 120 °C as source

temperature; 300 °C as desolvation temperature; cone gas flow of 150 L/h; and desolvation gas

flow of 600 L/h. A transfer collision energy of 4 V was used for both MS and tandem MS, while

the trap collision energy was set to 6 V for MS and a fixed voltage 28.4 V was used for MSMS

(also see notebook 5 page 9).

Page 42: © 2013 Neha Garg

32  

2.2.e Construction of pRSFDuet-1/His6-NisB+ NisA expression plasmid

The PCR amplification of nisA and nisB was performed by PCR using the conditions as

described above with genomic DNA of Lactococcus lactis 6F3 as DNA template. The primer

sequences are given in Table 2.1. The amplifications were analyzed on a 1% agarose gel and the

products were purified using a QIAquick Gel Extraction Kit. The amplified nisA DNA fragment

and the pRSFDuet-1 vector were digested with NdeI and XhoI in multiple cloning site (MCS) -2

at 37 °C for 12 h. The reaction mixtures were run on agarose gel and the desired bands extracted

using a QIAquick Gel Extraction Kit. The resulting double-digested nisA DNA insert was ligated

with the digested vector at 4 °C for 15 h using T4 DNA ligase (Invitrogen). E. coli DH5α cells

were transformed with the ligation product using heat shock and were plated on LB-kanamycin

agar plates and grown at 37 °C for 15 h. Colonies were picked up and incubated in 5 mL of LB-

kanamycin medium at 37 °C for 10 h, followed by isolation of the plasmids using a QIAprep

Spin Miniprep Kit (QIAGEN). The desired sequences of the resulting plasmids (pRSFDuet-1-

NisA) were confirmed by DNA sequencing using the appropriate primers by the Biotechnology

Center of the University of Illinois at Urbana-Champaign. pRSFDuet-1/NisA containing the nisA

gene was digested with BamHI and NotI (MCS-1) and ligated with the nisB PCR product

digested with BamHI and NotI. The ligation and transformation of E. coli DH5α cells was

performed as described above. Colonies were analyzed for the presence of insert and plasmid

was isolated and sequenced (also see notebook 5 page 5).

Page 43: © 2013 Neha Garg

33  

2.2.f Purification of unmodified His6-NisA and His6-NisA modified by wild type and mutant

NisB

The plasmid pET15 His6-NisA encoding unmodified NisA was obtained from former lab

member Dr. Trent J. Oman. The peptide NisA was expressed and purified using methodology as

described above for expression and purification of modified nisin. The plasmid pRSFDuet-His6-

NisA/His6-NisB was constructed using the same procedure as described above except that nisA

was inserted in MCS1 and nisB in MCS2. In order to introduce a hexahistidine tag at the N-

terminus of NisB, an oligonucleotide sequence encoding hexahistidine was incorporated in the 5’

end of the primer immediately before the start codon of nisB. Site directed mutagenesis was

carried out on pRSFDuet-His6-NisA/His6-NisB to replace conserved polar residues in NisB with

alanine using Phusion DNA polymerase and the primers listed in Table 2.1. Chemically

competent BL21(DE3) cells were transformed with pRSFDuet-His6-NisA/His6-NisB containing

the nisB variant genes. An overnight culture grown from a single colony transformant was used

as inoculum for 1.5 L of TB medium containing 50 µg/L kanamycin and 25 µg/L

chloramphenicol, and the culture was grown at 37 °C until the OD600 reached about 0.6. The

incubation temperature was then changed to 18 °C and the culture was induced with 0.5 mM

IPTG. The induced cells were shaken continually at 18 °C for an additional 15 h. The cells were

harvested by centrifugation (11,900 × g for 10 min, Beckman JLA-10.500 rotor). The cell pellet

was resuspended in 45 mL of start buffer and dehydrated non-cyclized prenisin was purified as

described above. The fractions containing modified NisA were lyophilized and analyzed by

MALDI-TOF MS (also see notebook 5 page 2).

Page 44: © 2013 Neha Garg

34  

2.2.g Overexpression and purification of His6-NisB and mutants R83A, R87A, T89A,

D121A, D299A, R464A, R505A, E751A, R786A, H961A, R966A, and R972A 

Single colony transformants of E. coli BL21 (DE3) containing pRSFDuetHis6-NisB/His6-

NisA or the corresponding NisB mutants were grown in a 37 oC shaker for 12-15 h in 30 mL of

LB medium supplemented with 50 μg/mL kanamycin and 25 μg/mL chloramphenicol. These

cultures were used to inoculate 4 L of TB media and grown at 37 oC until the OD600 was ~ 0.7.

IPTG was added to a final concentration of 0.5 mM and the cultures were grown at 18 oC for an

additional 20 h. Cells were harvested by centrifugation at 11,900 × g for 25 min at 4 oC. The cell

paste was resuspended in 20 mM Tris-HCl, 500 mM NaCl, and 10% glycerol, pH 8.0. The cells

were lysed using a C3 Homogenizer (Avestin Inc.). The lysed cells were pelleted by

centrifugation at 23,700 x g for 25 min. The purification by nickel column chromatography was

accomplished using a 5 mL HiTrap chelating HP nickel affinity column (GE Healthcare) on an

AKTA FPLC system (Amersham Pharmacia Biosystems). The supernatant from the

centrifugation step was applied to the column using a 50 mL superloop. The protein was detected

by absorbance at 280 nm. The column was washed with buffer A containing 1 M NaCl, 20 mM

Tris-HCl, pH 8.0, and 30 mM imidazole to remove any non-specifically bound proteins until a

flat baseline was observed. A gradient of 0-100% B over 30 min was used to elute the bound

proteins where buffer B contained 1 M NaCl, 20 mM Tris-HCl, pH 8.0, and 200 mM imidazole.

The most concentrated fraction (4 mL) was loaded onto an S-200 gel filtration column

(Amersham Biosciences) equilibrated in 20 mM HEPES pH 7.5, 300 mM NaCl. The eluted

protein was stored at −80 ˚C until use. The mutants were purified using the procedure described

above (also see notebook 5 page 6 and notebook 7 page 39 and 42).

Page 45: © 2013 Neha Garg

35  

2.2.h In vitro dehydration activity assay

The activity assay contained 200 µM His6-NisA, 5 µM His6-NisB, 50 mM HEPES pH

7.5, 10 µL E. coli cell extract, 1 mM TCEP, 10 mM MgCl2, 1.5 mM spermidine, 175 mM

glutamate, and 5 mM ATP. The activity assays were incubated at 30 °C for 4-6 h with gentle

shaking. The cell extract of E. coli BL21 codon-plus RIL was prepared as reported previously

(30) (see notebook 7 page 35, 52, and 63).

2.2.i Fractionation of cell extract into small molecule and macromolecule fractions

The cell extract (300 µL) was passed through a 3 kDa amicon filter. The retentate on the

filter was brought to a final volume of 300 L with 10 mM Tris-acetate buffer (pH 8.2)

containing 14 mM Mg(OAc)2, 60 mM KOAc, and 5 mM DTT. In a second experiment, the cell

extract was also passed through a Sephadex G-25 PD10 column (bed volume 8.3 mL; GE

healthcare) pre-equilibrated with 10 mM Tris-acetate buffer (pH 8.2) containing 14 mM

Mg(OAc)2, 60 mM KOAc, and 5 mM DTT. The excluded volume fraction collected was 3.5 mL

and included volume fraction collected was 7 mL. The different fractions thus obtained were

used as the cell extract substitutes in the activity assays, which were carried out as described

above (also see notebook 7 page 37).

2.2.j Fractionation of cell extract into cytoplasmic and membrane fractions

The cell extract (200 µL) was fractionated using a Beckman ultracentrifuge with a TLA

100 rotor. The extract was centrifuged for 2 h at 75,000 rpm (~ 200,000 × g). The supernatant

was carefully removed. The pellet, which presumably contains inner and outer membrane

components and ribosomes, was dissolved in 100 µL of 1 M Trisethanol amine acetate buffer,

Page 46: © 2013 Neha Garg

36  

pH 7.5. The different fractions thus obtained were used as substitutes for untreated cell extract in

the activity assays, which were carried out as described above (also see notebook 7 page 38).

2.2.k Glutamate elimination assay

His6-NisA carrying several +129 Da adducts was produced in E. coli by coexpression of

His6-NisB-R786A with His6-NisA. The peptide was purified by RP-HPLC and incubated with

His6-NisB or its Ala-mutants in HEPES buffer, pH 7.5, for 1.5 h. The assay mixture was desalted

using a µC18 ZipTip and analyzed by MALDI-MS (also see notebook 7 page 58).

2.3 Results

2.3.a Heterologous production of nisin in E. coli

The genes encoding His6-NisA and NisB were inserted into the pRSFDUET-1 vector and the

nisC gene was inserted into a pACYCDUET-1 plasmid. E. coli BL21 (DE3) cells were

transformed with both plasmids, grown on LB medium, and induced with IPTG. After harvesting

the cells, lysis, and IMAC purification, 24 mg of fully modified NisA was obtained per liter of

culture. The production strategy is summarized in Figure 2.2. The leader peptide was then

removed using trypsin (18), and the final product was analyzed by tandem MS and bioassays.

Similar fragmentation patterns in tandem MS and similar zones of inhibition in antimicrobial

assays were observed for authentic nisin and for nisin produced in E. coli demonstrating the

formation of authentic nisin (Figure 2.3 and 2.4). These experiments also demonstrate that active

NisB can be expressed in E. coli opening up in vitro reconstitution and mechanistic studies.

Page 47: © 2013 Neha Garg

37  

Figure 2.2. Strategy for nisin production in E. coli.

Page 48: © 2013 Neha Garg

38  

Figure 2.3. Production and characterization of nisin produced in E. coli. A) MALDI-MS spectrum of NisA modified in E. coli and treated with trypsin to remove the leader peptide. B) Antimicrobial assays against L. lactis HP of authentic nisin (positive control, PC-40 µM), nisin produced in E. coli (50 µM, 500 µM), and NC (negative control, reaction buffer containing trypsin). C) HPLC trace showing the purity of nisin produced in E. coli. The peak at 33 min is fully modified nisin.

Page 49: © 2013 Neha Garg

39  

Figure 2.4. Tandem MS analysis of nisin produced in E. coli and commercial nisin. ESI-MSMS analysis of commercial nisin (top) and nisin produced in E. coli (bottom). The b3 ion seen in both samples and the y30 ion seen in the bottom sample are uncommon for lanthionine rings but are consistently observed for intact nisin. Dha5 may be involved in this unusual cleavage of the A-ring in nisin.

Page 50: © 2013 Neha Garg

40  

2.3.b In vitro reconstitution of dehydration activity of NisB

The successful production of nisin in E. coli was achieved by co-expression of

hexahistidine tagged NisA (His6-NisA) and untagged NisB from a pRSFDuet-1 vector, and

untagged NisC from a pACYC vector as described above (31). For the purpose of in vitro studies

on NisB, these constructs were altered such that the gene for a hexahistidine tagged NisB (His6-

NisB) was cloned into the multiple cloning site 1 (MCS-1) of pRSFDUET-1 and the gene for

His6-tagged precursor peptide NisA was inserted into MCS-2. Using this plasmid, His6-NisA and

His6-NisB were co-expressed in E. coli and analysis of the cell extracts indicated full

dehydration of the His6-NisA peptide indicating that the His6-tag on NisB did not interfere with

in vivo NisB activity (Figure 2.5). The dehydration efficiency was slightly lower when omitting

co-expression with NisC (compare Figure 2.3A and 2.5). His6-NisB was then purified from the

cytoplasmic fraction of the E. coli cells using nickel affinity chromatography followed by gel

filtration chromatography. Curiously, incubation of the purified enzyme with His6-NisA in the

presence of ATP and Mg2+ did not result in any dehydration activity. Omitting the gel filtration

step also did not provide active enzyme, nor could any dehydrated peptide be detected after

addition of His6-NisA to purified His6-NisB in the presence of crude cell extracts of the E. coli

cells that produced His6-NisB and fully dehydrated His6-NisA. These results are similar to the

previously reported unsuccessful attempts to achieve in vitro activity of LanB enzymes (19-23).

Page 51: © 2013 Neha Garg

41  

Figure 2.5. Nisin produced in E. coli. MALDI-MS spectrum of NisA dehydrated in E. coli by NisB and treated in vitro with trypsin to remove the leader peptide. The efficiency of dehydration is slightly lower as compared to dehydrations in the presence of NisC. Also, partial dehydration of Ser29 is also observed leading to nine-fold dehydrated peptide.

Because the His6-tagged NisB did dehydrate His6-NisA in E. coli, we reasoned that some

property of the intracellular environment of the bacteria must contribute to activity. After

evaluating a host of different possibilities such as addition of ribosomes and crowding agents

without success, conditions used to optimize in vitro protein biosynthesis (32) were assessed.

When glutamate (175 mM), the most abundant metabolite in E. coli (33), and spermidine (1.5

mM), a polyamine involved in modulating various cellular functions, were added to His6-NisB (5

M), His6-NisA (200 M), ATP (5 mM), Mg2+ (10 mM), and E. coli cell extract prepared

according to a previously reported procedure that stabilizes cellular components (30), NisA was

dehydrated eight and nine times as determined by matrix-assisted laser desorption ionization

(MALDI) mass spectrometry (MS) (Figure 2.6). Nisin produced in the producing organism

Page 52: © 2013 Neha Garg

42  

Figure 2.6. In vitro reconstitution of dehydration activity of NisB. MALDI mass spectrum of His6-NisA incubated with His6-NisB, ATP, Mg2+, glutamate, and E. coli cell extract.

Lactococcus lactis and in E. coli is dehydrated eight times, with Ser29 in the core peptide

escaping dehydration (Figure 2.1). The observation of a partial ninth dehydration in vitro may be

caused by the absence of rings D and E, which may prevent dehydration of Ser29. Omitting ATP

resulted in unmodified His6-NisA as the only observed peptide (Figure 2.7A ). Loss of activity

also occurred when cell extract was replaced with water (Figure 2.7B), but using cell extract

from E. coli cells that did not contain any exogenous lantibiotic genes did result in full

dehydration. Complete loss in activity was also observed on omitting NisB from activity assays

(Figure 2.7D). Furthermore, spermidine was not essential for NisB activity as it could be omitted

from the assay without affecting the dehydration activity. Because the cell extract contained

Page 53: © 2013 Neha Garg

43  

significant amounts of glutamate, its importance could not be determined at this stage, but it was

later shown to be required (vide infra).

Figure 2.7. Determination of the required ingredients in the generic NisB activity assay. A) In vitro NisB activity assay in the absence of ATP. B) In vitro NisB activity assay in the absence of cell extract prepared as described in the reference (34). C) In vitro enzyme activity assay of NisB in the absence of glutamate. The cell extract was passed through a PD10 column in order to remove residual glutamate present in the extract. D) In vitro activity assay in the absence of NisB (but including all other ingredients). The peak at m/z 7273 in panels A and D originates from the E. coli cell extract (35).

Page 54: © 2013 Neha Garg

44  

2.3.c Fractionation of cell extract to investigate the cellular component required for

dehydration activity

The cell extract was fractionated into small molecule and macromolecule fractions using

two methods. First, cell extract was passed through a 3 kDa amicon ultracentrifugation filter.

Full dehydration activity was observed when the retentate was used in the assay mixture instead

of cell extract (Figure 2.8A), whereas no activity was seen when the flow-through was used

(Figure 2.8B). The cell extract was also passed through a Sephadex G-25 PD10 column, and the

macromolecule fraction was collected separately from the small molecule fraction. Again,

dehydration activity was observed by adding the macromolecule fraction to the NisB activity

assay, whereas no activity was observed with the small molecule fraction. The observation that

the macromolecule fraction contained the ingredient(s) needed for activity also allowed a test of

the requirement for glutamate. Addition of the macromolecular fraction of the gel filtration

experiment to the activity assay in the absence of glutamate did not result in dehydration of NisA

(Figure 2.7C), whereas inclusion of Glu (175 mM) again resulted in full dehydration activity.

In an attempt to identify the ingredient in the macromolecular fraction that is necessary

for activity, the cell extract was fractionated using ultracentrifugation at 200,000 × g. Addition of

the fraction that contains the membrane components as well as ribosomes to the activity assay

afforded dehydrated NisA peptides (Figure 2.8C), but activity was not observed when the

cytoplasmic fraction was added to the assay mixture (Figure 2.8D). Further fractionation of the

membrane fraction resulted in loss of activity. Moreover, addition of purified ribosomes to the

activity assay instead of cell extract did not result in dehydration activity and neither could cell

extract be replaced by exogenously added commercial polar lipids consisting of 67%

phosphatidylethanolamine, 23.2% phosphatidylglycerol, 9.8% cardiolipin (100600 P). Hence, at

Page 55: © 2013 Neha Garg

45  

present the ingredient present in the crude membrane fraction that is required for activity is not

known.

Figure 2.8. Investigation of the active ingredient(s) in cell extract. A) Full dehydration activity was observed when using the retentate of an ultrafiltration experiment with a membrane with a 3 kDa molecular weight cut-off for the in vitro NisB assay. B) MALDI-MS spectrum of His6-NisA incubated with His6-NisB, ATP, Mg2+, Glu and the flow through fraction of the ultrafiltration experiment. C) MALDI-mass spectrum of His6-NisA incubated with His6-NisB, ATP, Mg2+, Glu and the crude membrane fraction of an ultracentrifugation experiment at 200,000 x g. D) MALDI-mass spectrum of His6-NisA incubated with His6-NisB, ATP, Mg2+, Glu and the cytoplasmic fraction of the ultracentrifugation experiment.

Page 56: © 2013 Neha Garg

46  

2.3.d Alanine substitution of highly conserved residues in NisB

A previous study on the effects of mutagenesis of five conserved amino acids in NisB on

the dehydration reaction was performed in a Lactococcus lactis heterologous host (36). In the

current study, we interrogated the role of 23 additional highly conserved residues with side

chains that could be involved in the dehydration activity of NisB (Figure 2.9), including several

residues identified in a previous bioinformatics study (37). Because of ease of manipulation, the

previously described E. coli expression system (31) was used as an initial test. Mutation of

residues Arg575, Ser581, and Asp648 of His6-NisB to Ala did not affect the dehydration activity

as the product was eight-fold dehydrated His6-NisA. Mutation of Glu44, Lys68, Lys116, Glu133,

Arg505, Lys741, Glu823, Ser958, and Glu975 to Ala resulted in partially dehydrated products,

suggesting these residues are important but not essential for dehydration (Figure 2.10).

Conversely, individual replacement of residues Arg14, Arg83, Arg87, Thr89, Asp121, Asp299,

Arg464, and Arg966 with Ala resulted in abolishment of dehydration (Figure 2.11), suggesting

they are important for catalysis. The most interesting observations were made with the NisB

mutants R786A, R826A, and H961A. MALDI-MS analysis of the His6-NisA peptides that were

purified after co-expression with these mutants demonstrated multiple adducts of 129 Da (Figure

2.12). The inactive mutants and the mutants that led to adducts of 129 Da were selected for

further in vitro investigation. They were expressed with an N-terminal hexahistidine tag and

purified using nickel affinity column chromatography and gel filtration chromatography,

resulting in similar yields as wild type protein. Thus, the observed loss of activity or generation

of adducts was not a consequence of poor expression or solubility.

Page 57: © 2013 Neha Garg

47  

Figure 2.9. Continued on page 48.

Page 58: © 2013 Neha Garg

48  

Figure 2.9. Continued on page 49.

Page 59: © 2013 Neha Garg

49  

Figure 2.9. Sequence alignment of NisB with select other class I lantibiotic dehydratase enzymes. The residues mutated in this study are boxed with NisB residue numbers on top in blue. Mutations at positions resulting in adducts corresponding to 129 Da are denoted by a star under the residues.

Page 60: © 2013 Neha Garg

50  

Figure 2.10. Continued on page 51.

Page 61: © 2013 Neha Garg

51  

Figure 2.10. Modification of His6-NisA by NisB containing mutations at conserved residues in E. coli. MALDI-MS analysis of Ala mutants of NisB at positions (A) Glu44, (B) Lys68, (C) Lys116, (D) Glu133, (E) Arg505, (F) Lys741, (G) Glu823, (H) Ser958, and (I) Glu975.

Page 62: © 2013 Neha Garg

52  

Figure 2.11. Continued on page 53

Page 63: © 2013 Neha Garg

53  

Figure 2.11. Modification of His6-NisA by NisB containing mutations at conserved residues in E. coli. MALDI-MS analysis of Ala mutants of NisB at positions (A) Arg14, (B) Arg83, (C) Arg87, (D) Thr89, (E) Asp121, (F) Asp299, (G) Arg464, and (H) Arg966.

Page 64: © 2013 Neha Garg

54  

Figure 2.12. Modification of His6-NisA by NisB-R786A, R826A, and H961A in E. coli. MALDI mass spectra of His6-NisA co-expressed with (A) NisB-R786A, (B) NisB-R826A, and (C) NisB-H961A.

Page 65: © 2013 Neha Garg

55  

2.3.e Mechanistic investigation of NisB

To investigate the identity of the 129 Da adducts, the His6-NisA peptide that was

modified by the mutant NisB-R786A in E. coli was purified using nickel affinity

chromatography and RP-HPLC. The purified peptide was digested with trypsin and analysis of

the resulting fragments indicated that all the +129 Da adducts were present in the core peptide.

The observed masses are reminiscent of polyglutamylation of eukaryotic proteins, which

involves consecutive additions of glutamate to the side chain of Glu residues in proteins via

amide linkages between the -carboxy and the -amino groups of Glu (38). But unlike such

polyglutamylation, the 129 Da adducts on NisA could be readily removed by treatment with

base, suggesting that if the 129 Da adducts are associated with glutamylation, the linkage is more

likely to involve an ester, thioester, or anhydride. To confirm that Glu was the source of the

adducts, in vitro assays were performed. Incubation of His6-NisB-R786A with His6-NisA, ATP,

Mg2+, cell extract, and [2,3,3,4,4-d5]-Glu and subsequent MALDI-MS analysis showed multiple

+134 adducts on NisA (Figure 2.13). Thus, glutamate was confirmed as the source of the adducts

that are produced during co-expression in E. coli and in vitro assays.

Page 66: © 2013 Neha Garg

56  

Figure 2.13. In vitro activity assays of NisB-R786A with unlabeled and deuterium-labeled glutamate. MALDI mass spectra of His6-NisA incubated in vitro with NisB-R786A, ATP, Mg2+, cell extract and (A) glutamate or (B) 2,3,3,4,4-d5 glutamate.

Page 67: © 2013 Neha Garg

57  

To test whether the adducts could be intermediates in the dehydration process, His6-NisA

obtained from coexpression with NisB-R786A was again purified (Figure 2.14A), and the

modified peptide containing the multiple 129 Da adducts was incubated with purified wild type

His6-NisB. MALDI-MS analysis demonstrated conversion of NisA containing the adducts to

dehydrated peptide (Figure 2.14B). None of the other components of the in vitro NisB assay

mixture that are important for dehydration (glutamate, cell extract, Mg2+, or ATP) was required

for glutamate elimination. We also tested the in vitro glutamate elimination activity of the His6-

NisB mutants, which were all expressed in E. coli and purified. The alanine mutants of Arg83,

Arg87, Thr89, Asp121, Asp299, Arg 464, and Arg966 were confirmed to be inactive with

respect to dehydration of His6-NisA, but they had similar glutamate elimination activity as the

wild type protein on glutamylated His6-NisA (Figure 2.15). The alanine mutants of Lys116, and

Glu97 displayed partial dehydration activity but they again had similar glutamate elimination

activity as wild type NisB. The only mutants that exhibited only very low glutamate elimination

activity were the same mutants that in E. coli resulted in observation of glutamylated NisA:

Arg786Ala, Arg826Ala, and His961Ala (Figure 2.14).

Page 68: © 2013 Neha Garg

58  

Figure 2.14. Investigation of the in vitro elimination activity of wild type and alanine mutants of His6NisB. (A) RP-HPLC purified glutamylated peptide. MALDI-MS spectra showing the product from incubation of His6-NisA carrying several +129 Da adducts produced by His6-NisB-R786A with (B) wild type His6-NisB, (C) His6-NisB-His961Ala, (D) His6-NisB-Arg786Ala, and (E) His6-NisB-Arg826Ala.

Page 69: © 2013 Neha Garg

59  

Figure 2.15. Continued on page 60.

Page 70: © 2013 Neha Garg

60  

Figure 2.15. Investigation of the in vitro elimination activity of alanine mutants of His6NisB. MALDI-MS spectra showing the products from incubation of His6-NisA carrying several +129 Da adducts produced by His6-NisB-Arg786Ala mutant with (A) His6-NisB-Arg83Ala, (B) His6-NisB-Arg87Ala, (C) His6-NisB-Thr89Ala, (D) His6-NisB-Lys116Ala, (E) His6-NisB-Asp121Ala, (F) His6-NisB-Asp299Ala, (G) His6-NisB-Arg464Ala, (H) His6-NisB-Arg966Ala, (I) His6-NisB-Arg972Ala, and (J) His6-NisB-Glu975Ala.

Page 71: © 2013 Neha Garg

61  

When the products of the glutamate elimination assay containing glutamylated His6-NisA

and wild type NisB were incubated with L-glutamic acid dehydrogenase and NAD+, a clear

increase in absorbance at 340 nm was observed (Figure 2.16) suggesting oxidation of glutamate

that had been released from the NisA peptide by wild type NisB. Attempts to determine the sites

of glutamylation within the core peptide were unsuccessful as the Glu adducts were removed in

tandem MS experiments before any fragmentation of the peptide bonds. This observation once

again suggests that the linkage of Glu to NisA is not by an amide bond. Moreover, the observed

production of dehydrated peptide upon treatment with wild type NisB is most consistent with an

ester linkage to Ser/Thr. Further, various small metabolites were tested for dehydration activity

and the results are summarized in Table 2.1. Replacement of glutamate with the citric acid cycle

intermediates α-ketoglutarate, malate, and succinate did not result in dehydration activity.

Incubation with amino acids glutamine and iso-glutamine in place of glutamate resulted in partial

dehydration activity whereas incubation with aspartate resulted in up to 3 dehydrations (Figure

2.17).

Page 72: © 2013 Neha Garg

62  

Figure 2.16. Glutamic acid dehydrogenase assay to confirm identity of 129 Da adducts. Increase in absorbance at 340 nM upon incubation of the reaction mixture from Figure 2.13B with L-glutamic acid dehydrogenase and NAD+ (black curve). The increase in absorbance at 340 nm was not observed when the incubation was performed in the absence of His6-NisA carrying adducts (blue curve).

Page 73: © 2013 Neha Garg

63  

Figure 2.17. Continued on page 64.

Page 74: © 2013 Neha Garg

64  

Figure 2.17. Investigation of the activity of small metabolites as potential co-substrates for dehydration activity of NisB. MALDI-MS spectra showing the product from incubation of His6-NisA with His6-NisB, ATP, Mg2+, macromolecule fraction of E. coli cell extract, and (A) glutamine, (B) aspartate, (C) isoglutamine, (D) glutarate, (E) α-ketoglutarate, (F) malate and, (G) succinate. The cell extract was passed through a PD10 column in order to remove residual glutamate present in the extract. The peaks corresponding to M-14 and M-28 in penel E and G are MALDI-MS artifact peaks and are not observed with ESI-MS.

Page 75: © 2013 Neha Garg

65  

2.4 Discussion

NisB was initially isolated from membrane vesicles of the nisin producer L. lactis in 1992

(19) and was recently also purified in soluble form (23), but reconstitution of its activity in vitro

has remained elusive. In this study, we demonstrate the successful production of authentic nisin

in E. coli using coexpression of the precursor peptide NisA with tailoring enzymes NisB and

NisC. This methodology may be useful for cryptic type I lanthipeptides that cannot be isolated

from the producer strain, for compounds that are produced in very low abundancies, or for

preparing these compounds in isotopically labeled forms. Furthermore, NisB activity was

successfully reconstituted in vitro and required the membrane fraction of cell extract, ATP,

Mg2+, and glutamate. A direct role of glutamate in the physiological dehydration process is

suggested by several experimental observations. First, multiple glutamate adducts to the NisA

substrate are observed both in vitro and in E. coli with the NisB mutants R786A, R826A, and

H961A. Furthermore, small amounts of +129 Da adducts are also observed with wild type NisB

(Figure 2.6). Adducts of 130/131 Da were also present in the MALDI-MS spectra of NisA

products generated by the truncated NisB mutants Δ887-892 and Δ838-851 in a previous study

in L. lactis that showed that these truncated proteins were unable to dehydrate (36). The authors

suggested that the M+130/131 Da peaks were associated with NisA that still contained the N-

terminal Met, but no explanation was offered for the presence of ions of M+(2×130/131) Da,

which cannot be explained by incomplete removal of Met. Given that these truncations were in

the same C-terminal region of NisB as the single mutants that generated glutamylated products in

this study, we suggest that the observed peaks in L. lactis were glutamylated peptides. The

difference between the observed mass adducts of 130/131 Da and the calculated 129 Da adducts

Page 76: © 2013 Neha Garg

66  

for glutamylation could be a consequence of the experimental error in MALDI spectra of

peptides of this size.

A key observation in this study was that wild type NisB converted glutamylated NisA to

dehydrated NisA, demonstrating glutamylation generates a chemically competent intermediate.

Because none of the required ingredients for the dehydration reaction (ATP, Mg2+, Glu and cell

extract) were needed for glutamate elimination, a process that involves hydrolysis of the ester

followed by dehydration of the regenerated Ser/Thr residues can be ruled out. Collectively, we

believe that these data provide support for glutamylation of Ser/Thr as a required step for

dehydration. His961, Arg786, and Arg826 are clearly important for elimination of glutamate,

since alanine substitutions at these positions resulted in greatly diminished elimination activity in

vitro as well as in vivo. The His961Ala mutant of His6-NisB was also the least active in

glutamylation of His6-NisA (Figure 2.14) suggesting the mutation may have either a general

deleterious effect on protein structure or that this residue is important for both glutamylation and

elimination.

On the basis of our results, we propose that one of the carboxylic acid groups of

glutamate is activated by ATP and subsequently reacts with the side chains of serine and

threonine residues that are targeted for dehydration, resulting in glutamate esters (Figure 2.18A).

Upon deprotonation of the alpha proton of the glutamylated Ser or Thr residues, elimination of

glutamate results in dehydroalanine/dehydrobutyrine residues. It is possible that glutamylation

and elimination occur in different domains as is also observed for class III and IV lanthipeptides

that contain separate kinase and lyase domains (12, 39). If so, then the C-terminus of NisB

contains the likely lyase domain. The leader peptide binding site has been previously shown to

not be located in this C-terminal domain (36). The proposed activation mechanism for LanB

Page 77: © 2013 Neha Garg

67  

Figure 2.18. General mechanism for the elimination reaction. (A) The proposed mechanism for NisB catalyzed elimination of glutamate resulting in Dha/Dhb residue. Following addition of glutamate to serine and threonine in the precursor peptide, the abstraction of alpha proton leads to the formation of Dha and Dhb residues respectively via an enolate intermediate. (B) Reaction catalyzed by virginiamycin B lyase, the resistance protein to virginiamycin B. The elimination of the carboxylate is initiated by abstraction of the alpha proton and results in Z-dehydrobutyrine. Although not explicitly shown, the overall reaction may occur with an enolate intermediate (40).

dehydratases stands in contrast to the activation of the hydroxyl groups of Ser and Thr by

phosphorylation during the dehydration process of class II, III, and IV lanthipeptides (11-12, 41).

Why glutamylation of Ser/Thr would evolve as an alternative means of Ser/Thr activation is

unclear. One possibility is that it is a remnant of the use of activated Glu for other chemistry by

the ancestors of LanB proteins. Although unexpected in the context of Ser and Thr dehydration,

the elimination of a carboxylate from an ester linkage to the side chain of Thr in peptides has

precedent in the resistance protein to the cyclic hexadepsipeptide virginiamycin (42-43) as well

as related streptogramin antibiotics (44-45). The virginiamycin B lyase (Vgb) from

Staphylococcus aureus catalyzes an elimination reaction of the macrolactone that is formed by

an ester bond between the hydroxy group of a Thr and the C-terminal carboylic acid (Figure

Page 78: © 2013 Neha Garg

68  

2.18B). This anti elimination reaction generates a Z-dehydrobutyrine (42), the same

stereochemistry that is also generated by NisB in nisin (46-47). Vgb requires a Mg2+ for its

elimination reaction (42), whereas NisB catalyzes the elimination of Glu in the absence of Mg2+.

The Vgb crystal structure illustrates that the Mg2+ binds a 2-hydroxypicolinic acid group in

virginiamycin B (43) that is not present in the substrates for NisB, providing an explanation as to

why Mg2+ is not needed for the elimination step in NisB.

Whether NisB activates Glu itself or whether this is carried out by an enzyme in the

membrane fraction of the E. coli cell extract is currently not known because attempts to detect

Glu activation in the absence of cell extract have thus far been unsuccessful. Because of the

required presence of cell extract, I also have not been able to determine which of the two

carboxylates of Glu is activated and whether this activation involves adenylation or

phosphorylation. Nevertheless, this study provides the long-sought in vitro activity of a LanB

protein and a unique mode of Ser/Thr activation that provides the basis for further investigation.

By providing the conditions needed for in vitro activity, this study will also aid research on

several classes of natural products that utilize LanB proteins (48).

Future directions

In this study, I have demonstrated the in vitro dehydration activity of NisB, and have

shown that NisB catalyzed dehydration required the addition of a membrane-containing fraction

of E. coli cell extract, glutamate, ATP and Mg2+. Future studies should focus on determining the

role of the membrane-containing fraction, the mode by which ATP is utilized to activate

glutamate i.e. via adenylation or phosphorylation, and which carboxyl group of glutamate is

attached to the hydroxyl of serines/threonines undergoing dehydration.

Page 79: © 2013 Neha Garg

69  

2.5. Tables Table 2.1. Sequences of oligonucleotide primers used for site-directed mutagenesis.

Primer name Primer sequence

NisB_BamHI_F GTA GAC AGG GAT CCG ATG ATA AAA AGT TCA TTT AAA GCT CAA C

NisB_NotI_R AAT CAT AAG CGG CCG CTC ATT TCA TGT ATT CTT CCG AAA CAA AC

NisA_NdeI_F AAG CAG CCG CAT ATG AGT ACA AAA GAT TTT AAC TTG G

NisA_XhoI_R CTA GCT CGA GTT ATT TGC TTA CGT GAA TAC TAC

NisA_BamHI_F CTA GAT GGA TCC GAT GAG TAC AAA AGA TTT TAA CTT GG

NisA_HindIII_R CTA GAA GCT TTT ATT TGC TTA CGT GAA TAC TAC AAT G

NisB_NdeI_F AAG CAG CCG CAT ATG CAT CAT CAT CAT CAT CAT ATA AAA AGT TCA TTT AAA GCT CAA C

NisB_XhoI_R CTA GCT CGA GTC ATT TCA TGT ATT CTT CCG AAA CAA ACA ACC

NisC_BglII_F CTA GGG AAG ATC TGA TGA ATA AAA AAA ATA TAA AAA GAA ATG TTG

NisC_XhoI_R CTA GCT CGA GTC ATT TCC TCT TCC CTC CTT TCA AAA AAT CGT C

NisB_R14A_F GCT CAA CCG TTT TTA GTA GCG AAT ACA ATT TTA TCT CC

NisB_R14A_R TAC TAA AAA CGG TTG AGC TTT AAA TGA AC

NisB_E44A_F GTA AAA ATA AAG TTT TTT TGG CGC AGT TAC TAC TAG CTA ATC C

NisB_E44A_R CAA AAA AAC TTT ATT TTT ACT TAC AGT CTC

NisB_K68A_F GCT GGT CTG TTA AAG AAG GCG AGG GTT AAA AAA TTA TTT G

NisB_K68A_R CAA ATA ATT TTT TAA CCC TCG CCT TCT TTA ACA GAC CAG C

NisB_R83A_F CTA TTT ACA AGT ATT ATA AGG CGA GTT ATT TAC GAT CAA CTC

NisB_R83A_R CTT ATA ATA CTT GTA AAT AGA TTC AAA TAA TTT TTT AAC

NisB_R87A_F GTA TTA TAA GAG AAG TTA TTT AGC GTC AAC TCC ATT TGG ATT ATT TAG

NisB_R87A_R CTA AAT AAT CCA AAT GGA GTT GAC GCT AAA TAA CTT CTC TTA TAA TAC

NisB_T89A_F GAG AAG TTA TTT ACG ATC AGC GCC ATT TGG ATT ATT TAG TG

NisB_T89A_R CAC TAA ATA ATC CAA ATG GCG CTG ATC GTA AAT AAC TTC TC

NisB_K116A_F GTT AAT GGG AAA GAC TAC AGC GGG TAT AAG ATT GGA TAC TC

NisB_K116A_R GAG TAT CCA ATC TTA TAC CCG CTG TAG TCT TTC CCA TTA AC

NisB_D121A_F CAA AGG GTA TAA GAT TGG CGA CTC AGT GGT TGA TTC GCC

NisB_D121A_F GGC GAA TCA ACC ACT GAG TCG CCA ATC TTA TAC CCT TTG

Page 80: © 2013 Neha Garg

70  

Table 2.1 (cont.)

NisB_E133A_F CGC CTA GTT CAT AAA ATG GCG GTA GAT TTC TCA AAA AAG

NisB_E133A_R CTT TTT TGAG AAA TCT ACC GCC ATT TTA TGA ACT AGG CG

NisB_R505A_F GTA TTT CTT CCA GAA AAT ATC GCG CAT GCT AAC GTT ATG CAT AC

NisB_R505A_R GTATGCATAACGTTAGCATGCGCGATATTTTCTGGAAGAAATAC

NisB_R575A_F CGT TAT TCA GTA ATG AGC TAG CGT TTC TTT ACG AAA TTT C

NisB_R575A_R TAG CTC ATT ACT GAA TAA CGT TTT ATT GTA C

NisB_R581A_F CTA AGA TTT CTT TAC GAA ATT GCG TTA GAT GAC AAG TTT GG

NisB_R581A_R CCA AAC TTG TCA TCT AAC GCA ATT TCG TAA AGA AAT CTT AG

NisB_D648A_F GAG TTT TAT ATT GTC AAT GGA GCG AAT AAA GTT TAT TTA TCA C

NisB_D648A_R GTG ATA AAT AAA CTT TAT TCG CTC CAT TGA CAA TAT AAA ACT C

NisB_K741A_F CCC TTT AAC GAG TGG CTT TAT CTA GCG TTG TAC ATT TCT ATA AAT CG

NisB_K741A_R CGA TTT ATA GAA ATG TAC AAC GCT AGA TAA AGC CAC TCG TTA AAG GG

NisB_R786A_F CCA CAT ATT AGA TTG GCG ATA AAA TGT TCA GAT TTA TTT TTA G

NisB_R786A_R CTA AAA ATA AAT CTG AAC ATT TTA TCG CCA ATC TAA TAT GTG G

NisB_E823A_F GAT ATT TCT ATT TAT GAT CAA GCG GTA GAA AGA TAT GGT GG

NisB_E823A_R CCACCATATCTTTCTACCGCTTGATCATAAATAGAAATATC

NisB_R826A_F GAT CAA GAA GTA GAA GCG TAT GGT GGA TTT GAT AC

NisB_R826A_R GTA TCA AAT CCA CCA TAC GCT TCT ACT TCT TGA TC

NisB_S958A_F GTT TAT TCA ATT ATT GAC GCG ATC ATT CAT GTC CAT AAT AAC

NisB_S958A_R GTT ATT ATG GAC ATG AAT GAT CGC GTC AAT AAT TGA ATA AAC

NisB_H961A_F CAA TTA TTG ACA GTA TCA TTG CGG TCC ATA ATA ACC GAC TAA TTG

NisB_H961A_R CAA TTA GTC GGT TAT TAT GGA CCG CAA TGA TAC TGT CAA TAA TTG

NisB_R966A_F CAT TCA TGT CCA TAA TAA CGC GCT AAT TGG TAT TGA ACG AG

NisB_R966A_R CTC GTT CAA TAC CAA TTA GCG CGT TAT TAT GGA CAT GAA TG

NisB_E975A_F GGT ATT GAA CGA GAT AAA GCG AAA TTA ATT TAT TAC ACA C

NisB_E975A_R GTG TGT AAT AAA TTA ATT TCG CTT TAT CTC GTT CAA TAC C

Page 81: © 2013 Neha Garg

71  

Table 2.2. Activity of cellular small metabolites as co-substrates for NisB catalyzed dehydration reaction.

Metabolites tested Number of dehydrations

observed

α-ketoglutarate 0

Malate 0

Succinate 0

Glutamine 3-8

Isoglutamine 0 - 8

Aspartate 0 - 3

Glutarate 0

Page 82: © 2013 Neha Garg

72  

2.6 References 

1. Lubelski, J., Rink, R., Khusainov, R., Moll, G. N., and Kuipers, O. P. (2008) Biosynthesis, immunity, regulation, mode of action and engineering of the model lantibiotic nisin, Cell Mol Life Sci 65, 455-476.

2. Lee, M. V., Ihnken, L. A., You, Y. O., McClerren, A. L., van der Donk, W. A., and Kelleher, N. L. (2009) Distributive and directional behavior of lantibiotic synthetases revealed by high-resolution tandem mass spectrometry, J Am Chem Soc 131, 12258-12264.

3. Lubelski, J., Khusainov, R., and Kuipers, O. P. (2009) Directionality and Coordination of Dehydration and Ring Formation during Biosynthesis of the Lantibiotic Nisin, J Biol Chem 284, 25962-25972.

4. Kuipers, A., Meijer-Wierenga, J., Rink, R., Kluskens, L. D., and Moll, G. N. (2008) Mechanistic dissection of the enzyme complexes involved in biosynthesis of lacticin 3147 and nisin, Appl Environ Microbiol 74, 6591-6597.

5. Rink, R., Wierenga, J., Kuipers, A., Kluskens, L. D., Driessen, A. J., Kuipers, O. P., and Moll, G. N. (2007) Dissection and modulation of the four distinct activities of nisin by mutagenesis of rings A and B and by C-terminal truncation, Appl Environ Microbiol 73, 5809-5816.

6. Field, D., Connor, P. M., Cotter, P. D., Hill, C., and Ross, R. P. (2008) The generation of nisin variants with enhanced activity against specific gram-positive pathogens, Mol Microbiol 69, 218-230.

7. Field, D., Hill, C., Cotter, P. D., and Ross, R. P. (2010) The dawning of a 'Golden era' in lantibiotic bioengineering, Mol Microbiol 78, 1077-1087.

8. Wiedemann, I., Breukink, E., van Kraaij, C., Kuipers, O. P., Bierbaum, G., de Kruijff, B., and Sahl, H. G. (2001) 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, 1772-1779.

9. Knerr, P. J., and van der Donk, W. A. (2012) Discovery, biosynthesis, and engineering of lantipeptides, Annu Rev Biochem 81, 479-505.

10. Xie, L., Miller, L. M., Chatterjee, C., Averin, O., Kelleher, N. L., and van der Donk, W. A. (2004) Lacticin 481: in vitro reconstitution of lantibiotic synthetase activity, Science 303, 679-681.

11. Müller, W. M., Schmiederer, T., Ensle, P., and Süssmuth, R. D. (2010) In vitro biosynthesis of the prepeptide of type-III lantibiotic labyrinthopeptin A2 including formation of a C-C bond as a post-translational modification, Angew Chem, Int Ed 49, 2436-2440.

Page 83: © 2013 Neha Garg

73  

12. Goto, Y., Li, B., Claesen, J., Shi, Y., Bibb, M. J., and van der Donk, W. A. (2010) Discovery of unique lanthionine synthetases reveals new mechanistic and evolutionary insights, PLoS Biol 8, e1000339.

13. Nagao, J., Harada, Y., Shioya, K., Aso, Y., Zendo, T., Nakayama, J., and Sonomoto, K. (2005) Lanthionine introduction into nukacin ISK-1 prepeptide by co-expression with modification enzyme NukM in Escherichia coli, Biochem Biophys Res Commun 336, 507-513.

14. Lin, Y., Teng, K., Huan, L., and Zhong, J. (2011) Dissection of the bridging pattern of bovicin HJ50, a lantibiotic containing a characteristic disulfide bridge, Microbiol Res 166, 146-154.

15. Caetano, T., Krawczyk, J. M., Mosker, E., Süssmuth, R. D., and Mendo, S. (2011) Heterologous expression, biosynthesis, and mutagenesis of type II lantibiotics from Bacillus licheniformis in Escherichia coli, Chem Biol 18, 90-100.

16. Kluskens, L. D., Kuipers, A., Rink, R., de Boef, E., Fekken, S., Driessen, A. J., Kuipers, O. P., and Moll, G. N. (2005) Post-translational Modification of Therapeutic Peptides By NisB, the Dehydratase of the Lantibiotic Nisin, Biochemistry 44, 12827-12834.

17. Kuipers, A., Wierenga, J., Rink, R., Kluskens, L. D., Driessen, A. J., Kuipers, O. P., and Moll, G. N. (2006) Sec-Mediated Transport of Post-Translationally Dehydrated Peptides in Lactococcus Lactis, Appl Environ Microbiol 72, 7626-7633.

18. Li, B., Yu, J.-P. J., Brunzelle, J. S., Moll, G. N., van der Donk, W. A., and Nair, S. K. (2006) Structure and Mechanism of the Lantibiotic Cyclase Involved in Nisin Biosynthesis, Science 311, 1464-1467.

19. Engelke, G., Gutowski-Eckel, Z., Hammelmann, M., and Entian, K.-D. (1992) Biosynthesis of the lantibiotic nisin: genomic organization and membrane localization of the NisB protein, Appl Environ Microbiol 58, 3730-3743.

20. Siegers, K., Heinzmann, S., and Entian, K.-D. (1996) Biosynthesis of lantibiotic nisin. Posttranslational modification of its prepeptide occurs at a multimeric membrane-associated lanthionine synthetase complex., J Biol Chem 271, 12294-12301.

21. Xie, L., Chatterjee, C., Balsara, R., Okeley, N. M., and van der Donk, W. A. (2002) Heterologous expression and purification of SpaB involved in subtilin biosynthesis, Biochem Biophys Res Commun 295, 952-957.

22. Koponen, O., Tolonen, M., Qiao, M., Wahlstrom, G., Helin, J., and Saris Per, E. J. (2002) NisB is required for the dehydration and NisC for the lanthionine formation in the post-translational modification of nisin, Microbiology 148, 3561-3568.

23. Mavaro, A., Abts, A., Bakkes, P. J., Moll, G. N., Driessen, A. J., Smits, S. H., and Schmitt, L. (2011) Substrate recognition and specificity of NisB, the lantibiotic dehydratase involved in nisin biosynthesis, J Biol Chem 286, 30552-30560.

Page 84: © 2013 Neha Garg

74  

24. Kelly, W. L., Pan, L., and Li, C. (2009) Thiostrepton biosynthesis: prototype for a new family of bacteriocins, J Am Chem Soc 131, 4327-4334.

25. Wieland Brown, L. C., Acker, M. G., Clardy, J., Walsh, C. T., and Fischbach, M. A. (2009) Thirteen posttranslational modifications convert a 14-residue peptide into the antibiotic thiocillin, Proc Natl Acad Sci USA 106, 2549-2553.

26. Liao, R., Duan, L., Lei, C., Pan, H., Ding, Y., Zhang, Q., Chen, D., Shen, B., Yu, Y., and Liu, W. (2009) Thiopeptide biosynthesis featuring ribosomally synthesized precursor peptides and conserved posttranslational modifications, Chem Biol 16, 141-147.

27. Morris, R. P., Leeds, J. A., Naegeli, H. U., Oberer, L., Memmert, K., Weber, E., Lamarche, M. J., Parker, C. N., Burrer, N., Esterow, S., Hein, A. E., Schmitt, E. K., and Krastel, P. (2009) Ribosomally Synthesized Thiopeptide Antibiotics Targeting Elongation Factor Tu, J Am Chem Soc, 5946-5955.

28. Onaka, H., Nakaho, M., Hayashi, K., Igarashi, Y., and Furumai, T. (2005) Cloning and characterization of the goadsporin biosynthetic gene cluster from Streptomyces sp. TP-A0584, Microbiology 151, 3923-3933.

29. Bierbaum, G., Szekat, C., Josten, M., Heidrich, C., Kempter, C., Jung, G., and Sahl, H. G. (1996) Engineering of a novel thioether bridge and role of modified residues in the lantibiotic Pep5, Appl Environ Microbiol 62, 385-392.

30. Kigawa, T., Yabuki, T., Matsuda, N., Matsuda, T., Nakajima, R., Tanaka, A., and Yokoyama, S. (2004) Preparation of Escherichia coli cell extract for highly productive cell-free protein expression, J Struct Funct Genomics 5, 63-68.

31. Shi, Y., Yang, X., Garg, N., and van der Donk, W. A. (2011) Production of lantipeptides in Escherichia coli, J Am Chem Soc 133, 2338-2341.

32. Jewett, M. C., and Swartz, J. R. (2004) Mimicking the Escherichia coli cytoplasmic environment activates long-lived and efficient cell-free protein synthesis, Biotechnol Bioeng 86, 19-26.

33. Bennett, B. D., Kimball, E. H., Gao, M., Osterhout, R., Van Dien, S. J., and Rabinowitz, J. D. (2009) Absolute metabolite concentrations and implied enzyme active site occupancy in Escherichia coli, Nat Chem Biol 5, 593-599.

34. Garg, N., Salazar-Ocampo, L. M., and van der Donk, W. A. (2013) In vitro activity of the nisin dehydratase NisB, Proc Natl Acad Sci U S A 110, 7258-7263.

35. Ryzhov, V., and Fenselau, C. (2001) Characterization of the protein subset desorbed by MALDI from whole bacterial cells, Anal Chem 73, 746-750.

36. Khusainov, R., Heils, R., Lubelski, J., Moll, G. N., and Kuipers, O. P. (2011) Determining sites of interaction between prenisin and its modification enzymes NisB and NisC, Mol Microbiol 82, 706-718.

Page 85: © 2013 Neha Garg

75  

37. Marsh, A. J., O'Sullivan, O., Ross, R. P., Cotter, P. D., and Hill, C. (2010) In silico analysis highlights the frequency and diversity of type 1 lantibiotic gene clusters in genome sequenced bacteria, BMC Genomics 11, 679.

38. Janke, C., Rogowski, K., and van Dijk, J. (2008) Polyglutamylation: a fine-regulator of protein function?, EMBO Rep 9, 636-641.

39. Goto, Y., Ökesli, A., and van der Donk, W. A. (2011) Mechanistic studies of Ser/Thr dehydration catalyzed by a member of the LanL lanthionine synthetase family, Biochemistry 50, 891-898.

40. Gerlt, J. A., Babbitt, P. C., and Rayment, I. (2005) Divergent evolution in the enolase superfamily: the interplay of mechanism and specificity, Arch Biochem Biophys 433, 59-70.

41. Chatterjee, C., Miller, L. M., Leung, Y. L., Xie, L., Yi, M., Kelleher, N. L., and van der Donk, W. A. (2005) Lacticin 481 Synthetase Phosphorylates its Substrate during Lantibiotic Production, J Am Chem Soc 127, 15332-15333.

42. Mukhtar, T. A., Koteva, K. P., Hughes, D. W., and Wright, G. D. (2001) Vgb from Staphylococcus aureus inactivates streptogramin B antibiotics by an elimination mechanism not hydrolysis, Biochemistry 40, 8877-8886.

43. Korczynska, M., Mukhtar, T. A., Wright, G. D., and Berghuis, A. M. (2007) Structural basis for streptogramin B resistance in Staphylococcus aureus by virginiamycin B lyase, Proc Natl Acad Sci U S A 104, 10388-10393.

44. Bateman, K. P., Yang, K., Thibault, P., White, R. L., and Vining, L. C. (1996) Inactivation of etamycin by a novel elimination mechanism in Streptomyces lividans, J Am Chem Soc 118, 5335–5338.

45. Lipka, M., Filipek, R., and Bochtler, M. (2008) Crystal structure and mechanism of the Staphylococcus cohnii virginiamycin B lyase (Vgb), Biochemistry 47, 4257-4265.

46. Fukase, K., Kitazawa, M., Akihiko, S., Shimbo, K., Fujita, H., Horimoto, S., Wakamiya, T., and Shiba, T. (1988) Total synthesis of Peptide antibiotic Nisin, tl 29, 795-798.

47. Chan, W. C., Lian, L. Y., Bycroft, B. W., and Roberts, G. C. K. (1989) Confirmation of the structure of nisin by complete proton NMR resonance assignment in aqueous and dimethyl sulfoxide solution, J Chem Soc, Perkin Trans 1, 2359-2367.

48. Arnison, P. G., Bibb, M. J., Bierbaum, G., Bowers, A. A., Bugni, T. S., Bulaj, G., Camarero, J. A., Campopiano, D. J., Challis, G. L., Clardy, J., Cotter, P. D., Craik, D. J., Dawson, M., Dittmann, E., Donadio, S., Dorrestein, P. C., Entian, K.-D., Fischbach, M. A., Garavelli, J. S., Göransson, U., Gruber, C. W., Haft, D. H., Hemscheidt, T. K., Hertweck, C., Hill, C., Horswill, A. R., Jaspars, M., Kelly, W. L., Klinman, J. P., Kuipers, O. P., Link, A. J., Liu, W., Marahiel, M. A., Mitchell, D. A., Moll, G. N., Moore, B. S., Müller, R., Nair, S. K., Nes, I. F., Norris, G. E., Olivera, B. M., Onaka, H., Patchett, M. L., Piel, J., Reaney, M. J. T., Rebuffat, S., Ross, R. P., Sahl, H.-G., Schmidt,

Page 86: © 2013 Neha Garg

76  

E. W., Selsted, M. E., Severinov, K., Shen, B., Sivonen, K., Smith, L., Stein, T., Süssmuth, R. E., Tagg, J. R., Tang, G.-L., Truman, A. W., Vederas, J. C., Walsh, C. T., Walton, J. D., Wenzel, S. C., Willey, J. M., and Donk, W. A. v. d. (2013) Ribosomally Synthesized and Post-translationally Modified Peptide Natural Products: Overview and Recommendations for a Universal Nomenclature, Nat Prod Rep 30, 108-160.

Page 87: © 2013 Neha Garg

77  

Chapter 3: Discovery, heterologous production, and characterization of a novel type I

lantibiotic geobacillin I 1

3.1 Introduction

The lantibiotic nisin, the most extensively studied member of the lantibiotic family, was

first approved for use as a food preservative to combat food-borne pathogens in 1969 and is

currently used in over 80 countries (1). Despite its use for over 50 years, reports of significant

resistance against nisin have not appeared (2-3). The lack of resistance may stem from the dual

mode of action of nisin. Nisin exhibits antimicrobial activity by binding to the pyrophosphate

moiety of lipid II (4), a membrane bound advanced intermediate involved in the biosynthesis of

the cell wall. By doing so, nisin inhibits the transglycosylation step in cell wall biogenesis and

sequesters lipid II from its functional location (5-6). Furthermore, the nisin-lipid II complex

leads to formation of pores in the membrane causing cell death. The mechanism-of action of

nisin is described in greater detail in Chapter 1. Resistance mechanisms such as efflux pumps

and enzyme mutations will not affect nisin as it acts on the outside of the bacterial cell and has a

small molecule as target. In addition to its use as a food preservative, the Center for Veterinary

Medicine of the U.S. Food and Drug Administration recently ruled positively on application of

nisin for intramammary treatment of subclinical mastitis in dairy cattle (7-8). After approval of

the pending New Animal Drug Application, a nisin-containing product would allow treatment of

bovine mastitis with a zero milk discard time and zero meat withdrawal period (i.e. milk and/or

1Parts of this chapter have been adapted with permission from:

Garg N, Tang W, Goto Y, Nair SK, van der Donk WA,. Proc Natl Acad Sci U S A. 2012 Apr 3; 109(14):5241-6. Reproduced with permission of the publisher and available from http://www.pnas.org/ (DOI: 10.1073/pnas.1116815109).

Page 88: © 2013 Neha Garg

78  

meat from treated cattle would not have to be discarded). One drawback that has been noted for

nisin is its limited stability at pH 7 (9-13). Hence, more stable analogues may prove more

effective.

Reported herein is the structure of a lantibiotic from Geobacillus thermodenitrificans

NG80-2 termed geobacillin I that is an analogue of nisin with two additional crosslinks. The

compound was produced heterologously in E. coli, its ring topology was determined by NMR

spectroscopy, and its activity against various bacteria was assessed. The compound is three-fold

more active than nisin against Streptococcus dysgalactiae, one of the main contagious causative

agents of clinical bovine mastitis (14). Geobacillin I also demonstrated increased stability

compared to nisin. Similar to nisin, geobacillin I was found to bind lipid II and disrupt the

membrane of Gram-positive bacteria.

3.2 Experimental procedures

3.2.a Construction of pRSFDuet-1/His6-GeoAI+GeoB and pACYCDuet-1/GeoC

Codon-optimized synthetic genes for GeoAI, GeoB, and GeoC were synthesized by

GeneArt (Invitrogen, USA). The nucleotide sequences are given at the end of this chapter. For

geobacillin I production, geoAI was inserted in MCS1 of a pRSFDuet-1 vector using BamHI and

HindIII restriction sites, geoB in MCS2 using NdeI and XhoI sites, and geoC in MCS2 of

pACYCDuet-1 using NdeI and XhoI sites. Site directed mutagenesis was performed to mutate

Asn at the −1 position to Lys (for primers see Table 3.1, also see notebook 7 page 5).

Page 89: © 2013 Neha Garg

79  

3.2.b Screening of Geobacillus strains for lantibiotic production

Seven Geobacillus strains were purchased from the Bacillus Genetic Stock Center. The

strains were streaked on tryptose blood agar base (TBAB) and modified LB (mLB) plates. The

composition of mLB media was tryptone (10 g), yeast extract (5 g), NaCl (5 g), agar (15 g),

NaOH (1.25 mL of 10% w/v), nitrilotriacetic acid (1.05 M), MgSO4.7H2O (0.59 M), CaCl2.2H2O

(0.91 M), FeSO4.7H2O (0.04 M), and water 995 mL. Nitriloacetic acid, MgSO4 .7H2O,

CaCl2.2H2O, and FeSO4.7H2O were filter sterilized and added to the media after autoclaving. A

small amount of cells from a colony was picked from the plate with a pipette tip and was spotted

on a MALDI plate. The sample was overlaid with 2 µL of a 9:1 mixture of 2,5-dihydroxybenzoic

acid and 2-hydroxy-5-methoxybenzoic acid matrix prepared in 60% acetonitrile/0.1% TFA in

water (80 mg/mL). A heat gun was used to dry the sample, and the samples were analyzed by

MALDI-MS (also see notebook 5 page 41).

3.2.c Production and purification of geobacillin I

Single colony transformants of E. coli BL21 (DE3) containing plasmids pRSFDuet-1

(GeoAI/GeoB) and pACYCDuet-1 (GeoC) were picked and grown in LB media overnight. The

culture was used to inoculate 6 L of TB media containing 50 mg/L kanamycin and 34 mg/L

chloramphenicol. The culture was shaken at 200 rpm at 37 °C until the OD600 reached between

0.8 and 1.0. At this point, the culture was induced with 1 mM IPTG. The induced cells were

shaken continually at 37 °C for an additional 2 h, and the cells were harvested by centrifugation

(11,900×g for 10 min, Beckman JLA-10.500 rotor). The cell pellet was resuspended in 50 mL of

start buffer (20 mM Tris, pH 8.0, 500 mM NaCl, and 10% glycerol), and cell lysis was carried

out using a MultiFlex C3 homogenizer (Avestin). The lysed cells were centrifuged at 23,700×g

Page 90: © 2013 Neha Garg

80  

for 30 min at 4 °C. The supernatant was loaded onto a HiTrap HP nickel affinity column (GE

Healthcare) pre-equilibrated with start buffer. The column was washed with wash buffer (start

buffer containing 30 mM imidazole), and the peptide was eluted from the column with elution

buffer (start buffer containing 1 M imidazole) (also see notebook 5 page 42).

To increase the yield, modified GeoAI was also recovered from the insoluble fraction

after cell lysis. The pellet was homogenized using a sonicator (35% amplitude, 4.4 s pulse, 9.9 s

pause for 20 min total) in start buffer to remove any soluble proteins. The suspension was

centrifuged and the pellet was homogenized in 30 mL of denaturing buffer (4 M guanidine

hydrochloride, 20 mM NaH2PO4, 500 mM NaCl, pH 7.5). The supernatant after centrifugation

was loaded onto a HiTrap HP nickel affinity column, the column was washed with denaturing

buffer containing 30 mM imidazole, and the product was eluted with 15 mL of denaturing buffer

containing 1 M imidazole. Small aliquots of eluent were desalted using a ZipTip (µC18) and

analyzed for the presence of geobacillin I using MALDI-MS. For large scale purification, the

eluent was desalted using preparative scale RP-HPLC using a Waters Delta-pak C4 15 μm; 300

Å; 25x100 mm PrepPak Cartridge. Following injection, the column was kept at 2% solvent A

(solvent A = 0.086% TFA in 2% ACN/98% water) for 2 min followed by a gradient of 2-100%

of solvent B (0.086% TFA in 80% ACN/20% water) over 45 min. The fractions containing

modified GeoAI were lyophilized.

3.2.d Cleavage of modified GeoAI and antimicrobial activity assay.

A solution of 500 µM posttranslationally modified GeoAI-N−1K in 50 mM HEPES

buffer (pH 8.0) was treated with 0.5 µM trypsin for 2 h. The product was analyzed using

MALDI-MS. The nine-fold dehydrated and cyclized core peptide was purified using RP-HPLC

Page 91: © 2013 Neha Garg

81  

in order to remove incompletely dehydrated peptide and leader peptide using a Phenomenex

Luna C18 column (250x10 mm, 10 micron) operating at a flow rate of 8 mL/min. The program

used was 3 min of solvent A (2% MeCN/0.1% TFA) followed by a gradient of 2-100% solvent B

(80% MeCN/0.1% TFA) over 50 min (also see see notebook 5 page 47).

3.2.e Specific activity of geobacillin I and nisin

For minimal inhibitory concentration (MIC) calculations, a 48-well plate (300 µL well

size) was used with shaking and a 96-well plate (200 µL well size) was used for non-shaking

conditions, depending on the indicator strain. Serial dilutions of nisin and geobacillin I were

prepared in sterile deionized water (SDW). The 48/96 plate wells contained 50/75 μL of diluted

peptide at defined concentrations and 150/225 μL of a 1:10 dilution (approximately 1 x 108 CFU

mL-1) of a culture of indicator strain diluted in fresh growth medium. Growth medium blank and

negative controls lacking the lantibiotics were also prepared. Plates were incubated under

appropriate growth conditions (Table 3.2), and MIC calculations were performed by monitoring

the OD600. For agar diffusion bioactivity assays, the appropriate agar media was melted in a

microwave, kept at 45 °C for 10 min, and mixed with 150 μL of overnight culture of indicator

organisms (108-109 cfu). After solidification, 10 µL of a solution of 100 M peptide was spotted

on the plate (also see notebook 5 page 54).

3.2.f Chemical stability of geobacillin I and nisin

Stock solutions of nisin and geobacillin I (30 µM) were prepared in 50 mM sodium

phosphate buffer, 100 mM NaCl at pH 7.0 and 8.0. The stock solutions were divided into 200 µL

aliquots in sterile eppendorf tubes and were incubated at 37 °C and 60 °C. At set time points, a

tube was opened and analyzed by agar diffusion growth inhibition assay. The agar diffusion

Page 92: © 2013 Neha Garg

82  

growth inhibition assay was carried out in sterile Nunc OmniTrays. Molten agar (20 mL) was

inoculated with 200 µL of overnight culture of B. subtilis ATCC 6633 diluted to OD600 of 1.0.

An additional 30 mL of molten agar combined with 300 µL cells was poured over the solidified

agar. A sterile 96 well PCR plate was carefully placed on the upper molten agar layer.

Following solidification, the PCR plate was removed and 20 µL of peptide solution was added to

the wells. After 20 min, another 20 µL of peptide solution was added. This procedure was

repeated one more time to add a combined total volume of 60 µL. The earlier time points from

samples at 37 oC were also analyzed by HPLC. The stability tests for geobacillin I S5F/L6I and

nisin I4K/S5F/L6I were performed as described above with a minor variation. Stock solutions of

geobacillin I, nisin, geobacillin I S5F/L6I and nisin I4K/S5F/L6I (50 μM) were prepared in

MOPS buffer, 100 mM NaCl at pH 7.0 instead of phosphate buffer (also see notebook 5 page

61).

3.2.g Membrane disruption assay

The fluorescent dye 3,3′-diethyloxacarbocyanine iodide (DiOC2) was used to investigate

disruption of the membrane potential of B. subtilis ATCC 6633. An overnight culture of B.

subtilis ATCC 6633 grown in LB at 37 °C was diluted to OD600 of 1.0 with fresh LB. The

membrane disruption assay mixture containing DiOC2 (final concentration 2 μM), HEPES (1

mM), and glucose (1 mM), was incubated with the cells (final OD600 of 0.1) for 20 min at room

temperature. The final volume of the mixture was adjusted to 300 µL with LB medium. The

lantibiotics nisin and geobacillin I were added to final concentrations of 0.2, 2.0, and 20 μM and

the samples were incubated for an additional 15 min. A negative control containing water instead

of antibiotic was also prepared in similar manner. The mean fluorescence intensity (MFI) was

recorded with a BD Biosciences LSR II flow cytometer, using excitation at 488 nm with an

Page 93: © 2013 Neha Garg

83  

argon laser and measurement of emission through a band-pass filter at 530/30 nm. A minimum

of 80,000 events were detected for each sample, and experiments were performed in triplicate.

Data analysis to calculate the MFI of gated cell populations was performed using FCS Express

version 4.00. The fluorescent dye propidium iodide (PI) was used to detect pore formation in the

membrane of B. subtilis ATCC6633. The assay mixture contained PI (final concentration 25

μM), phosphate buffered saline, nisin/geobacillin I (13, 10, 5, 2,5 , 1.25, 0.63, 0.31, 0.16, 0.08,

and 0 μM) and B. subtilis ATCC 6633 cells diluted to a final OD600 of 0.1. The final volume of

the mixture was adjusted with LB medium to 300 μL. The assay mixture was incubated for 10

min at room temperature. The data was recorded and analyzed as mentioned above with emission

recorded through a band-pass filter at 695/40 nm (also see notebook 5 page 73).

3.2.h Inhibition of penicillin binding protein (PBP1b) by geobacillin I

A transglycosylation assay to investigate binding of nisin and geobacillin I to lipid II

using the transglycosylase PBP1b from E. coli was performed by a former lab member Dr. Trent

Oman in Prof. Suzanne Walker’s laboratory at Harvard Medical School, Boston, Massachusetts

as previously described (15-17). Briefly, chemo-enzymatically prepared [14C]-GlcNAc-labeled

heptaprenyl lipid II analogue was incubated with varying concentrations of nisin and geobacillin

I. The inhibition of transglycosylase activity of PBP1b was monitored as previously described

(17).

3.2. i Chiral gas chromatography/mass spectrometry (GC/MS) analysis

Wild type GeoAI fully modified by GeoB and GeoC in E. coli and with its leader

attached (1.2 mg) was dissolved in 6 mL of 6 M HCl. The mixture was heated at 110 °C in a

sealed tube for 24 h. The reaction was cooled and dried under nitrogen overnight. Methanol (5

Page 94: © 2013 Neha Garg

84  

mL) was chilled in an ice-water bath, and acetyl chloride (1.5 mL) was added dropwise. This

solution was added to the hydrolyzed geobacillin I and the resulting solution was heated at 110

°C for 1 h. The sample was cooled and dried under a stream of nitrogen. CH2Cl2 (3 mL) and

pentafluoropropionic anhydride (1 mL) was added and the material was heated at 110 °C for 40

min. The reaction was allowed to cool and dried under a stream of nitrogen. The hydrolyzed and

derivatized geobacillin I was taken up in methanol. Nisin (0.3 mg) purified from commercial

nisaplin was also derivatized using the procedure described above. The derivatized amino acids

from geobacillin I, nisin, as well as synthetic (methyl)lanthionine standards prepared by Weixin

Tang (18) were analyzed by GC/MS using an Agilent HP 5973 mass spectrometer and a Varial

CP-Chirasil-L-Val 25 m × 0.25 mm × 0.15 µM silica column. The samples were introduced to

the instrument via a pulsed splitless injection at an inlet temperature of 200 °C and flow rate of 2

mL/min of helium gas. The temperature gradient used was 160 °C for 5 min, then 160 °C to 200

°C at 3 °C/min. The MS was operated in simultaneous scan/single-ion monitoring (SIM) mode,

monitoring at fragment masses of 365 for lanthionine and 379 for methyllanthionine (18) (also

see notebook 5 page 56).

3.3 Results

3.3.a Description of the gene cluster of geobacillin I from Geobacillus thermodenitrificans

NG80-2

G. thermodenitrificans sp. are thermophilic Gram-positive bacteria that have been found

in diverse niches such as oil reservoirs, geothermal areas, and marine environments. They have

gained recent interest as a potential source for industrially useful thermostable enzymes. G.

thermodenitrificans NG80-2 was isolated from the Dagang oil fields in China. Its genome was

Page 95: © 2013 Neha Garg

85  

recently sequenced (19) and two groups of genes were annotated in the deposition to the

National Center for Biotechnology Information database as lantibiotic biosynthetic genes. The

enzymes encoded in one cluster were annotated as homologs of the biosynthetic enzymes of the

class I lantibiotic subtilin produced by B. subtilis 6633 (GenBank accession numbers

YB_001124395-Y_001124403). The second cluster is described in Chapter 4. The gene cluster

for geobacillin I was noted in bioinformatic genome mining studies (20), but the product of the

gene cluster is unknown. The gene cluster for geobacillin I contains a short open reading frame

for the precursor peptide, which we designated geoAI (Figure 3.1A). The core peptide of the

GeoAI precursor peptide has homology with nisin and subtilin, in particular in the N-terminal

region. However, GeoAI is predicted to have two more Cys residues than the precursor peptides

for these known lantibiotics (Figure 3.1B). The 23-residue leader peptide of GeoAI has high

homology to the leader peptides for the lantibiotics ericin S and subtilin (Figure 3.1C). On the

basis of these sequence alignments, the site for proteolytic removal of the leader peptide was

predicted as ProAsn↓Val (Figure 3.1B and C); this prediction was confirmed by detection of the

lantibiotic by colony mass spectrometry of other Geobacillus strains (vide infra).

Genes typically found in lantibiotic biosynthetic gene clusters are present downstream of

geoAI including genes encoding a dehydratase (geoB), ABC transporter (geoT), cyclase (geoC),

a two-component transcriptional regulator (geoR and geoK), and self-resistance proteins (geoI,

geoG, geoE, and geoF) (Figure 3.1A). Similar gene clusters with nearly identical precursor

peptides (the only difference being a Val for Ile change) were also found in other genomes of

Geobacillus (Figure 3.1B and C), but some of these clusters have frameshifted biosynthetic

genes (20). No gene was identified that encodes for a typical class I lantibiotic protease to

remove the leader peptide.

Page 96: © 2013 Neha Garg

86  

Figure 3.1. Gene cluster and primary sequence alignment of the precursor peptide of geobacillin I from G. thermodenitrificans NG80-2. A) Gene cluster for the biosynthesis of geobacillin I. The gene for the precursor peptide is shown in red, modification enzymes in yellow, transporter/protease in purple, regulatory proteins in blue, and immunity proteins in green. B) Sequence alignment of the core peptides of selected known class I lantibiotics with that of GeoAI. Serines and threonines undergoing dehydration are in blue, cysteines involved in ring formation are in red. The additional two cysteines not found in nisin are highlighted in yellow. C) Sequence alignment of the leader peptides of select class I lantibiotics with the leader peptide of geobacillin I.

Page 97: © 2013 Neha Garg

87  

3.3.b Screening of Geobacillus strains for lantibiotic production

Because G. thermodenitrificans NG80-2 was not available to us, seven different

Geobacillus strains (21) were screened for production of the predicted geobacillins (G.

thermodenitrificans DSM465, G. thermodenitrificans OHT-1, Geobacillus sp. M10EXG, G.

thermodenitrificans OH2-1, G. thermodenitrificans OH5-2, G. thermodenitrificans NM16-2, and

G. kaustophilus DSM7263). The bacteria were grown at 55 °C on TBAB and mLB media for 48

h. Whole-cell MALDI-MS (22) was used to investigate the production of geobacillin I after 12,

24, and 48 h. A mass corresponding to GeoAI after nine dehydrations and removal of the leader

peptide at the predicted site was observed in both media for five of the strains after 24 h. Only G.

thermodenitrificans NM16-2 and G. kaustophilus DSM7263 did not produce geobacillin I under

the conditions used. A representative MALDI mass spectrum is shown in Figure 3.2. All five

organisms produced peptides of the same mass, consistent with the near-identical sequences of

GeoAI observed in the currently sequenced genomes (Figure 3.1B). All five observed peptides

had masses consistent with a Val at position 15 rather than an Ile. These observations along with

the genome data suggest very high conservation of the sequence of geobacillin I in the

Geobacillus genus. The mass spectrum in Figure 3.2 also shows peaks corresponding to eight-

fold and seven-fold dehydrated peptide, reminiscent of incomplete dehydration of nisin in

Lactococcus lactis 6F3 (23).

Page 98: © 2013 Neha Garg

88  

Figure 3.2. Geobacillin I production in native and heterologous producers. MALDI-MS spectra showing geobacillin I production by a native producer in black and by the heterologous E. coli host in red (section 3.3.c). Purified nine-fold dehydrated and cyclized core peptide is shown in blue. Nine-fold, eight-fold, and seven-fold dehydrated peptides are denoted by *, #, and •, respectively.

3.3.c Heterologous production of geobacillin I in E. coli

Given our success in producing nisin by coexpression in E. coli (Chapter 2) (24),

synthetic genes encoding the precursor peptide GeoAI and the dehydratase GeoB were inserted

into the MCS-1 and MCS-2 of a pRSFDuet-1 vector, respectively. This construct results in a

hexa-histidine-tagged precursor peptide (His6-GeoAI) and an untagged GeoB protein. A

synthetic gene encoding the cyclase GeoC was inserted into MCS-2 of a pACYCDuet-1 vector,

resulting in an untagged GeoC enzyme. Site-directed mutagenesis was performed to mutate Asn

at the −1 position of the GeoAI peptide (Figure 3.1C) to Lys to incorporate a trypsin cleavage

site in the precursor peptide to be used for leader peptide removal. E. coli BL21 (DE3) cells were

Page 99: © 2013 Neha Garg

89  

transformed with both plasmids and after induction of protein expression, His6-GeoAI was

purified from the cell lysate using nickel affinity chromatography followed by reverse phase high

performance liquid chromatography (RP-HPLC). The modified precursor peptide was treated

with trypsin and MALDI-MS analysis of the resultant peptide demonstrated a nine-fold

dehydrated core peptide (red line, Figure 3.2). As seen with the native producers, peaks

corresponding to eight and seven-fold dehydrated peptides were also observed.

3.3.d Bioactivity of geobacillin I

The antimicrobial spectrum of geobacillin I was investigated with a range of Gram-

positive bacteria. Modified GeoAI was produced in E. coli as described above, treated with

trypsin, purified by RP-HPLC (blue trace, Figure 3.2), and used for agar diffusion growth

inhibition assays (Figure 3.3) and MIC determinations in liquid medium (Table 3.3). The

compound was active against a wide range of Gram positive bacteria but it was not active against

Gram negative E. coli. Interestingly, in parallel experiments, geobacillin I demonstrated three-

fold higher activity against Streptococcus dysgalactiae ATCC 27957 than nisin (Figure 3.3 and

Table 3.3). This bacterium is one of the causative agents of bovine mastitis (14). Geobacillin I

displayed similar activity as nisin against vancomycin-resistant enterococci (VRE) and Bacillus

anthracis Sterne 7702, and had five-fold lower activity against methicillin-resistant

Staphylococcus aureus (MRSA) and Bacillus subtilis ATCC 6633. Its antimicrobial activity

against Lactococcus lactis HP is about seven-fold lower than nisin (Table 3.3).

Page 100: © 2013 Neha Garg

90  

Figure 3.3. Agar diffusion growth inhibition assay. A) Zones 1, 2, and 3 represent inhibition of S. dysgalactiae ATCC 27957 by three different batches of geobacillin I (50 µM). Zones 4, 5, and 6 represent inhibition by nisin (50 µM). B) Zones 1 and 3 represent inhibition of VRE by two different batches of geobacillin I (50 µM). Zone 2 represents inhibition by nisin (50 µM). C) Zone 1 represents inhibition of MRSA by geobacillin I (50 µM). Zones 2 and 3 represent zones of inhibition by two different batches of nisin (50 µM).

3.3.e Comparison of chemical stability of geobacillin I and nisin

Given the larger number of conformation-restraining crosslinks and the thermophilic

producing strain, we anticipated that geobacillin I might be inherently more stable than nisin.

Geobacillin I was indeed found to be more stable than nisin at pH 7 and 8.5 at 37 oC and at 60 oC

(Figure 3.4, 3.5, 3.6, and 3.7). The samples at 37 oC were also analyzed by HPLC but for a

shorter time period (Figure 3.8). At later time points, the amount of remaining intact peptides

was difficult to determine because of closely eluting breakdown products. The main breakdown

product of geobacillin I appeared to be an oxidation product (M+16) but this did not appear to

significantly affect the antimicrobial activity. Nisin has been shown to undergo chemical

degradation at neutral pH at Dha5 (12). One intriguing difference in geobacillin I compared to

nisin is the presence of a Lys at position 4 (in ring A). In light of a previous report that the nisin

analogue I4K/Dha5F/L6I had higher activity against spores of B. subtilis 168 and cells of

Page 101: © 2013 Neha Garg

91  

Micrococcus luteus NCDO 8166, Lactobacillus johnsonii 10533 and Leuconostoc mesenteroides

20343 (25), we speculated that the replacement of Dha5 with Phe in this mutant may contribute

to increased activity by virtue of increased stability. These considerations lead us to investigate

whether the introduction of lysine at position four in nisin might also increase its stability. The

geobacillin I and nisin analogues with amino acid residues KFI at positions 4-6 were generated

by coexpression of the corresponding precursor peptides with the appropriate tailoring enzymes

in E. coli and were not significantly higher in stability than the wild type geobacillin I and nisin,

respectively, (Figure 3.9) while displaying similar MIC values (Table 3.4, and Figure 3.10 and

3.11 for mass spectra). Thus, more complex biological factors and differences in overall structure

may be responsible for the higher stability of geobacillin I.

Figure 3.4. Agar diffusion growth inhibition assay with B. subtilis 6633 to determine the relative stabilities of geobacillin I and nisin A at pH 7.0 and 60 °C. Aliquots of 60 L of 30 µM solutions of compound that had been incubated for the time shown (in hours) were added to each well.

Page 102: © 2013 Neha Garg

92  

Figure 3.5. Agar diffusion growth inhibition assay with B. subtilis 6633 to determine the relative stabilities of geobacillin I and nisin A at pH 8.0 and 60 °C. Aliquots of 60 L of 30 µM solutions of compound that had been incubated for the time shown (in hours) were added to each well.

Figure 3.6. Agar diffusion growth inhibition assay with B. subtilis 6633 to determine the relative stabilities of geobacillin I and nisin A at pH 7.0 and 37 °C. Aliquots of 60 L of 30 µM solutions of compound that had been incubated for the time shown (in days) were added to each well. Nisin standards (60 L per well) are shown on the left.

Page 103: © 2013 Neha Garg

93  

Figure 3.7. Agar diffusion growth inhibition assay with B. subtilis 6633 to determine the relative stabilities of geobacillin I and nisin A at pH 8.0 and 37 °C. Aliquots of 60 L of 30 µM solutions of compound that had been incubated for the time shown (in days) were added to each well. Nisin standards (60 L per well) are shown on the left.

Page 104: © 2013 Neha Garg

94  

Figure 3.8. Assessment of the relative stabilities of nisin A and geobacillin I (Geo) using HPLC analysis.

Page 105: © 2013 Neha Garg

95  

Figure 3.9. Agar diffusion growth inhibition assay with B. subtilis 6633 at pH 7.0 and 37 °C. A) geobacillin I, B) geobacillin I S5F/L6I, C) nisin, and D) nisin I4K/S5F/L6I. Aliquots of 60 µL of 50 μM solutions of compound dissolved in MOPS buffer and 100 mM NaCl that had been incubated for the time shown (in days) were added to each well.

Page 106: © 2013 Neha Garg

96  

Figure 3.10. MALDI-MS spectra of purified A) geobacillin I, B) geobacillin I L19NVA, C) geobacillin I L19P, and D) geobacillin I S5F/L6I.

Page 107: © 2013 Neha Garg

97  

Figure 3.11. MALDI-MS spectra of purified A) nisin B) nisin I4K/S5F/L6I.

3.3.f Investigation of membrane disruption using flow cytometry analysis

Nisin contains a flexible hinge region of three amino acids between the C and D rings

that has been shown to be indispensable for pore formation activity (26-28). For instance, the

ΔN20ΔM21 and N20P/M21P mutants of nisin lost pore formation ability against Gram-positive

bacteria (26-27). Geobacillin I has only a single amino acid between the C and D rings, similar to

the ΔN20ΔM21 mutant of nisin as shown by NMR analysis by Weixin Tang and discussed later

(Figure 3.17). Thus, we sought to explore whether geobacillin I, similar to nisin, binds to lipid II

and forms pores in the membrane of Gram-positive bacteria. Flow cytometry was used to

examine changes in the polarization of the bacterial membrane of B. subtilis ATCC 6633 upon

incubation with geobacillin I using the membrane potential sensitive dye 3,3′-

diethyloxacarbocyanine iodide (DiOC2) (15, 29). Incubation with geobacillin I resulted in a

significant decrease in mean fluorescence intensity (MFI) similar to the observations with nisin

(Figure 3.12A and Figure 3.13), and hence implying efficient pore formation in the membrane of

B. subtilis, despite the lack of a hinge region. The efficiency of pore formation by geobacillin I

Page 108: © 2013 Neha Garg

98  

was also investigated using propidium iodide (PI), a membrane impermeable fluorescent dye.

Upon pore fomation, PI can enter the cell resulting in an increase in fluorescence intensity

because of the interaction of PI with nucleic acids. PI uptake was monitored at nine different

concentrations with each experiment conducted in triplicate (Figure 3.12B and Figure 3.14). In

antimicrobial activity assays in liquid medium, geobacillin I exhibited a four-fold increase in the

minimal inhibitory concentration (MIC) against B. subtilis compared to nisin (Table 3.4), but

showed only two-fold lower efficiency in pore formation (Figure 3.12B).

Previously, site-saturation mutagenesis was performed on the amino acids in the hinge

region of nisin. The antibiotic activity of the nisin mutants N20V and M21A was increased

against pathogenic bacteria (30). Conversely, introduction of Pro into the hinge region abolished

pore formation. For comparison, these same mutations were chosen for introduction into the

region between rings C and D in geobacillin I. Site-directed mutagenesis was used to replace the

naturally occurring Leu19 with Pro and with the tripeptide AsnValAla that is found in the

improved nisin variants. These analogues were generated by coexpression of mutants of the

precursor peptide with the modification enzymes GeoB and GeoC in E. coli (31). The geobacillin

analogue with NVA and P at the hinge region had eight-fold and two-fold increased MIC values

as compared to geobacillin I (Table 3.4), respectively. The ability to induce pore formation by

these analogues was also investigated. Although the efficiency of pore formation was reduced,

replacement of Leu19 with Pro at the hinge region did not abolish this activity, unlike the

observations for nisin. Furthermore, introduction of the amino acid residues NVA at the hinge

region greatly reduced formation of pores in the bacterial cell wall by geobacillin I. Thus, it

seems likely that the detailed mechanism of pore formation by geobacillin I differs from that of

nisin.

Page 109: © 2013 Neha Garg

99  

Figure 3.12. Membrane depolarization assay to investigate the effect on the membrane potential of B. subtilis ATCC 6633 by geobacillin I and nisin measured using DiOC2 and propidium iodide fluorescence by flow cytometry. (A) Average mean fluorescence intensity (MFI) of triplicate measurements with different concentrations of nisin and geobacillin I using DiOC2. The different bars represent concentrations of 0, 0.2, 2, and 20 µM (left to right). (B) Increase in MFI resulting from propidium iodide uptake for nisin and geobacillin I (average of three measurements). The different bars represent concentrations of 0, 0.31, 1.25, 2.5, 5 and 10 µM (left to right).

Page 110: © 2013 Neha Garg

100  

Figure 3.13. Representative histogram of cell count versus DiOC2 fluorescence intensity at various geobacillin I and nisin concentrations (red: no antibiotic; yellow: 0.2 µM nisin; dark red: 2 µM nisin; black: 20 µM nisin; blue: 0.2 µM geobacillin I; turquoise: 2 µM geobacillin I; green: 20 µM geobacillin I).

Figure 3.14. The histograms of cell count versus PI fluorescence intensity at various concentrations are shown for A) geobacillin I, B) nisin, C) geobacillin I L19P and, D) geobacillin I L19NVA (black: 0 µM; red: 13 µM; blue: 10 µM; violet: 5 µM; green: 2.5 µM; yellow: 1.25 µM; dark red: 0.63 µM; teal: 0.31 µM; grey: 0.16 µM; turquoise: 0.08 µM).

Page 111: © 2013 Neha Garg

101  

3.3.g Investigation of lipid II binding by geobacillin I

The ability of geobacillin I to bind lipid II was investigated using an in vitro inhibition

assay of transglycosylation reaction catalyzed by E. coli PBP1b by Dr. Trent Oman. PBP1b uses

lipid II as a substrate for glycan polymerization (16). Geobacillin I produced in my studies

inhibited PBP1b-catalyzed peptidoglycan formation using 4 μM heptaprenyl lipid II (Figure

3.15) with a half-maximal inhibitory concentration (IC50) of 4.55 ± 0.76 M. For comparison,

inhibition by nisin displayed an IC50 of 2.91 ± 0.61 M (Figure 3.12). Thus, the inhibitory

activity of the two peptides was found to be very similar.

Fig. 3.15. Inhibition of PBP1b-catalyzed peptidoglycan (PG) formation by geobacillin I (A) and nisin (B), at a lipid II concentration of 4 μM and a PBP1b concentration of 100 nM. Error bars represent the standard deviation from triplicate experiments.

3.3.h Determination of stereochemistry of (methyl)lanthionine crosslinks in geobacillin I

using GC/MS analysis

The stereochemical purity of (methyl)lanthionine was determined by acid hydrolysis of

geobacillin I and chemical derivatization of the resulting amino acids as previously described

(32-33). The derivatized material was then analyzed by GC/MS. The total ion chromatogram for

Page 112: © 2013 Neha Garg

102  

geobacillin I is shown in Figure 3.16A. The derivatized lanthionine and methyllanthionine

present in hydrolyzed geobacillin I and nisin samples are shown in Figure 3.16B and C,

respectively. The Lan and MeLan derivatives eluted at identical retention times with

(methyl)lanthionine standards confirming that (methyl)lanthionines in nisin and geobacillin I

have the same configuration, (2S, 3S, 6R)-methyllanthionine and (2S, 6R)-lanthionine.

Figure 3.16. (A) Total ion gas chromatogram of derivatized amino acids obtained by acid hydrolysis of geobacillin I. (B) GC/MS trace (SIM 379 Da for MeLan, and 365 Da for Lan) of derivatized (methyl)lanthionine obtained from geobacillin I (black) overlayed with a GC/MS trace of a synthetic (2S, 3S, 6R)-methyllanthionine standard (blue)

Page 113: © 2013 Neha Garg

103  

and a (2S, 6R)-lanthionine standard (red). The intensities were normalized to 1.0 for the largest peak. The small shoulders on the derivatized lanthionine peak have been reported previously and arise from partial epimerization during the acid hydrolysis (32). (C) GC/MS traces (SIM 379 Da, 365 Da) for derivatized (methyl)lanthionine obtained from hydrolyzed geobacillin I (black) compared with derivatized (methyl)lanthionine obtained from nisin A (pink). The stereochemistry of the derivatized (methyl)lanthionines is the same for both compounds and the relative ratio of Lan to MeLan of the two compounds is close to the predictions from the structures in Figure 3.1B (1 Lan, 4 MeLan in nisin A and 4 Lan, 3 MeLan in geobacillin I. The intensities were normalized to 1.0 for the largest peak.

3.4 Discussion

The availability of a rapidly increasing number of bacterial genomes is starting to shift

the search for new gene-encoded peptide natural products from activity-based discovery

platforms to gene-based discovery approaches (34-36). In the lantibiotic group of compounds,

genome mining has resulted in the discovery of several new family members (37-39), including

the first lantibiotic from an alkaliphilic organism (40-41). In addition, a large number of

biosynthetic gene clusters has been identified that remain to be explored experimentally (20, 42).

Our attention was drawn to genomes of thermophiles since only one lantibiotic from a

moderately thermophilic organism has been reported (43). The lantibiotics produced by

thermophiles could find potential applications as they may be more stable than lantibiotics

produced by mesophilic bacteria. Particularly interesting was the discovery of nisin precursor

genes in the genomes of Geobacillus because their sequence suggested that the

posttranslationally modified products would contain additional rings. Indeed, the masses of the

compounds produced by five Geobacillus strains and the corresponding lantibiotic

heterologously produced in E. coli showed nine dehydrations and the NMR structure determined

by Weixin Tang (31) for the latter compound demonstrated seven thioether crosslinks, including

conservation of the A and B rings found in nisin that are important for lipid II binding (5-6, 44)

(Figure 3.17). In all, 16 out of the 33 residues in the core peptide of GeoAI are

posttranslationally modified, and the final product geobacillin I was shown to have high

Page 114: © 2013 Neha Garg

104  

antimicrobial activity against several pathogens. Furthermore, the compound is more stable than

nisin A at pH 7 and 8.5 and at high temperatures.

Compared to nisin, ring C of geobacillin I is reduced in size by one amino acid, the hinge

region is reduced in length by two amino acids, and ring D is a lanthionine ring as opposed to a

methyllanthionine ring (Figure 3.17). In addition, geobacillin I contains two additional thioether

bridges at its C-terminus. However, although there appear to be differences in regards to the

importance and role of the hinge region, the overall mode of action of the two lantibiotics

appears to be quite similar: binding to lipid II and pore formation. These findings suggest that the

previously proposed role of the hinge region for pore formation activity by elongated type A

lantibiotic peptides may not be universal.

This study demonstrates that lantibiotics may be produced in nature at temperatures as

high as 50-80 °C, the temperature of the deep sub-surface Dagang oilfield where G.

thermodenitricans NG80-2 was isolated. The successful production of geobacillin I in E. coli

was therefore somewhat surprising because many enzymes from thermophilic organisms are

non-functional at 37 °C. However, GeoB and GeoC carried out dehydrations and cyclizations

with apparently high efficiency at this temperature. Future work will focus at producing non-

natural analogues of geobacillin I using the coexpression technology in E. coli.

Page 115: © 2013 Neha Garg

105  

Figure 3.17. A) Structure of geobacillin I, and B) Structure of nisin. Hinge region residues are shown in brown color.

Page 116: © 2013 Neha Garg

106  

3.5 Tables Table 3.1. Sequences of oligonucleotide primers used in this study.

Primer name Primer sequence

GeoAI_BamHI_F CTA GAT GGA TCC GAT GGC CAA ATT TGA TGA TTT TGA TC

GaoAI_HindIII_R CTA GAA GCT TTT ATT AGC AAC GAA TAC AGC TAT TAC AGC

GeoB_NdeI_F ACA GCC GCA TAT GAA TGA TCT GGT GTT TAA AAA TAT TG

GeoB_XhoI_R TCT AGC TCG AGTTAA TTT TTC AGC ACC AGG CCA TGG GC

GeoC_NdeI_F AAG CAG CCG CAT ATG TCT ATT AGC ATG AAA GCC CTG G

GeoC_XhoI_R CTA GCT CGA GTAAAA CTT CGC TCA G CA GAA ATG CAC AAT CC

GeoAI (N−1K)_F CAG GAT GAT GTT GTT CAG CCG AAA GTT ACC AGC AAA AGC CTG

GeoAI(N−1K)_R CGG CTG AAC AAC ATC ATC CTG TTT TTT CAC C

GeoAI_L19P_F AT TAC CGG TGT TCT GAT GTG TCCGAC CCA GAA TAG CTG TG

GeoAI_L19P_R ACA CAT CAG AAC ACC GGT AAT ACA ACC CGG

GeoAI_L19NVA_F AT TAC CGG TGT TCT GAT GTG TAA CGT GGC AAC CCA GAA TAG CTG TG

GeoAI_L19NVA_R ACA CAT CAG AAC ACC GGT AAT ACA ACC CGG

GeoAI_ S5F/L6I _F AGC CGA AAG TTA CCA GCA AAT TTA TTT GTA CAC CGG GTT G

GeoAI_S5F/L6I_R TTT GCT GGT AAC TTT CGG CTG AAC AAC ATC

NisA_I4K/S5F/L6I_F

CAT CAC CAC GCA TTA CAA GTA AAT TCA TTT GTA CAC CCG GTT GTA AAA C

NisA_ I4K/S5F/L6I _R

GTT TTA CAA CCG GGT GTA CAA ATG AAT TTA CTT GTA ATG CGT GGT GAT G

Page 117: © 2013 Neha Garg

107  

Table 3.2. Growth conditions for indicator strains used in this study.

Strain Media Conditions

Streptococcus dysgalatiae subsp dysgalactiae BHI 37 °C, aerobic

Vancomycin resistant Enterococcus faecium BHI 37 °C, aerobic

Methicillin resistant Staphylococcus aureus BHI 37 °C, aerobic

Bacillus anthracis Sterne 7702 BHI 37 °C, aerobic

Bacillus subtilis ATCC6633 LB 30 °C, aerobic

Lactococuus lactis HP GM17 30 °C, anaerobic

Table 3.3. Specific activity of geobacillin I and nisin in liquid growth inhibition assay.

Strain Source Geobacillin I

IC50 (µM)

Geobacillin I

IC90 (µM)

Nisin

IC50 (µM)

Nisin

IC90 (µM)

Streptococcus dysgalatiae subsp dysgalactiae

ATCC a

27957 0.69 ± 0.05 0.87 2.12 ± 0.04 3.87

Vancomycin resistant Enterococcus faecium

CNRZb 481 0.84 ± 0.05 1.1 0.39 ± 0.03 0.57

Methicillin resistant Staphylococcus aureus

C5c 2.23 ± 0.03 3.39 0.42 ± 0.01 0.77

Bacillus anthracis Sterne 7702

Gut et al.d 0.49 ± 0.02 0.071 0.21 ± 0.014 0.52

Bacillus subtilis ATCC 6633 0.55 ± 0.01 0.81 0.11 ± 0.01 0.16

Lactococuus lactis HP ATCC 11602 0.12 ± 0.09 0.13 0.017 ± 0.005

0.019

aATCC: American type Culture Collection. bCNRZ: National Centre for Zootechnical Research. cClinical isolate from Carle Foundation Hospital (Urbana, IL). dReference (45).

Page 118: © 2013 Neha Garg

108  

Table 3.4. Specific activity of nisin, geobacillin I and the analogues generated in this study against B. subtilis ATCC6633 in liquid growth inhibition assay.

Table 3.5. Theoretical and observed molecular weights ([M+H]1+) of modified peptides reported in this study.

Peptide IC50 (µM) MIC (µM)

Nisin 0.12 ± 0.01 0.25 Geobacillin I N−1K 0.56 ± 0.01 1.0 Geobacillin I L19P 1.00 ± 0.03 2.0 Geobacillin I L19NVA 5.70 ± 0.08 8.0 Nisin I4K/S5F/L6I 0.25 ± 0.05 0.5 Geobacillin I S5F/L6I 0.45 ± 0.09 1.0

Peptide Protease

used

Theoretical

[M+H]1+

Observed

[M+H]1+

Geobacillin I (from GeoA-N−1K) Trypsin 3265.5 3265.3

Nisin - 3352.6 3352.5

Nisin-I4K/S5F/L6I Trypsin 3445.7 3445.9

Geobacillin-S5F/L6I Trypsin 3340.6 3341.4

Geobacillin-L19P Trypsin 3246.5 3246.1

Geobacillin-L19NVA Trypsin 3433.6 3433.4

Page 119: © 2013 Neha Garg

109  

3.6 Codon optimized nucleotide sequence for GeoAI

1 ATGGCCAAAT TTGATGATTT TGATCTGGAT ATTGTGGTGA AAAAACAGGA TGATGTTGTT 61 CAGCCGAATG TTACCAGCAA AAGCCTGTGT ACACCGGGTT GTATTACCGG TGTTCTGATG 121 TGTCTGACCC AGAATAGCTG TGTTAGCTGT AATAGCTGTA TTCGTTGCTA ATAA

3.7 Codon optimized nucleotide sequence for GeoB

1 ATGAATGATC TGGTGTTTAA AAATATTGAT ACCTTTATGA TTCGTACACC GGTTCTGAGC 61 GTTGATAATT ATCTGCGCTT TTTTGATCAG AAACTGACCG AAGGCGAAAT GAAAGAACGT 121 CTGCTGGAAA TTTGTCATAA TCCGGTGTTT CGTGAAAGCA TTCTGGTTGC AAGCAAAAGC 181 CTGTATAACA AAATGATTGA TTTTTGCAAT GGCAAAGAGA TTAAGAAATA CGATTACTTT 241 ATCAAAGCCA TCTATAAATA TCTGATTCGC ATTAGCACCC GTCCGACCCC GTTTGGTCTG 301 TTTGCAGGTG TTGATTTTGG CGAATATACC GATGAAAATA CCAGCATTCG CTATGGCACC 361 AATAAATATA AAAAATTTGC CCGTCCGGAT CTGGAATGGC TGATGAAAAT TATTAAAAAA 421 CTGGAACAGG AACAGTACGA ACAGCTGTGG TTTACCGTGA ATGATAGCAT TTTTCTGAAA 481 GGCGAACGTG CATATCTGCT GCATAGCACC CGTAAAGATG ATGATAAACG CGTGAATGAA 541 ATTAGCGTTC GTGTTACCCT GCCGTTTAAA ATTACCTGTG AACTGGCTCG CCATCTGATT 601 CATTATCAGA CCCTGAAAAA AGAACTGATT AAACAGTTTC CGAATACCAG CGAAGAAAAA 661 ATTGAACGCT TTCTGAAACA GCTGATTGAA AATGAATTTC TGATTAGCAA TCTGCGTCCT 721 CCGCTGACCG TTATGGATCA GCTGGATTAT CTGATTAAAC GCCTGAAAGA AAGCCATATT 781 GAAGAATGGT CCAATGAACT GATTGATATT CAGCAGAAAA TTCGCACCTA TACCATGACA 841 CCGCTGGGTG AAGGTGAGCA GATTTATAAA GAACTGCATA AAAAAATGAA AAAACTGGCC 901 GATACCAAAA ATGTTCTGCA GGTGGATATG AAACTGAATC TGCAGGAAAA AAAACTGAAT 961 AAACAAGTGA TTAAAGATGT GAATGAACTG ATGCATATTC TGCTGCCGTT TAGCATGACC 1021 TATCAGCAGA CCGATTCTCC GCTGTCTCGT TATAAACAGG AATTTATTGA AAAATACGGC 1081 GTGGATCGTG AAGTTCCGCT GCTGGAAATG CTGGATAATG ATCTGGGTAT TGGTGCACCG 1141 ATGGATTATA CCAATCCGAA AAATGCACGT CTGGCCGTTG GTGTTCCGCA TCAGCTGATT 1201 GATGAACAGC TGCGCGAATA TTTTCTGCAT AAATATATTG AAGCCGTGAA AGATAAAAAA 1261 AGCATTACCA TTACCGAAGA TGAAATTAAA GAACTGGGCC TGGATGATTA TGATTATCAG 1321 AGCATTCCGG ATAGCCTGGA TATTAATGTG ATTGTGAAAT GCGATAGCGA ACAGGATCTG 1381 GAAAATAATC ATTTTAAATA TTATATTGGT CCGAATCTGG GTAGCACACA TGCAGGTAAA 1441 TCTTTCGGTC GCTTTAGCTA TATGATGGAA GGCGAAGAAA AACTGTTTGA AACCCTGAAT 1501 AAAGAAATGC TGAATCTGAA TAATGATAGC GGTGAATATG TTACCTGCGA ACTGGTTTAT 1561 CTGCCGAATG AAACCCGTCA TGCAAATGTG ACCCGCAATA TTCATAATAG CGAATATGAA 1621 CTGGCCCTGT TTACCAATAA TAGCAAAGAT GAATATCATC GTCTGAGCCT GAATGATATT 1681 CTGATCGGCA TTGAAAATGA TATGTTTTAT GCCAAAAGCA AAACACTGAA TAAAAAGCTG 1741 ATTATTACCA TGAATAATAT GCTGAATACC CAGAATGCAC CGAATGCAAT TCGCTTTCTG 1801 TATGATATTA GCCTGGATGG TAAAAAAATT TGGTGTAATT TTCCGTGGGA TACCATTTTT 1861 CAGGATTATG CCTATATTCC GGCAATTGAA TATAAAAATT TTGTGATTGC ACCGCGTAAA 1921 TGGGTGATTA ATAAGATTAT CAAAACAAAC GATAAAATCA GCTTTACCGA ATTTGAAAAA 1981 CGCTTTAAAA AATTTTGCAG CGATTATGAT GTGCCGAAAT ATGTGTATCT GACCTTTGCC 2041 GATAATCGTA TTCTGCTGAA TACGGAAGAT GAAAATTGCC TGCGCATTCT GCATCATGAA 2101 TTTAATAATC GCTATGGTAA TATTATTCTG AGCAGCTATG AAGAAGCAGA TCTGCATCTG 2161 GTGAAAGATG ATGATGAAAA AGAATATGTG TGCGAACTGG TGATTCCGCT GGTTAAAGTG 2221 AAACAAGAAA ATCGTCAGTG GTACGCCAAA GCCAAAAAAA GCAATCATAT TCCGAGCTTT 2281 AGCGATATTC GTGTGAAACT GCCGTTTGAA AATTGGCTGT ATCTGAAACT GTATGGTGTT 2341 CATAGCCGTG CCGATGATTT TATTGCCTTT TATCTGAGCG AATATTGCCG TGAAAAACTG 2401 GCCAAAGGTG AAATTGAAAA ATATTTTTTT ATTCGCTATG CCGATCCGGA ACAGCATATT 2461 CGTCTGCGTC TGAATGCATC TGAAGAAAAA CTGCTGCAGA TTTATCCGGA TATGAAACAG 2521 TGGCTGCGTA CCCTGATTGA AAAACGTGTT GTGAGCCGCT TTAGCATTGA TTGCTATGAA 2581 CGCGAAATTG AACGTTATGG TGGCACCGAA CTGATCGATC TGGCCGAAAA TTTATTTTTT 2641 TTTGATAGCA TTGTGGTGGA AGATATTCTG AAACTGAAAC GTCTGAAGCA TATTCAGTTT 2701 TCCGATGAAG TGATTGGTAT GGTGAGCATT ATTCATTATA TGGAACAGTT TGGCCTGAAT 2761 TATGAAAGCC AGCTGGATTT TCTGCAGCAG CAGGTTAATA AACAGGATTA TCGCGAAGAT

Page 120: © 2013 Neha Garg

110  

2821 TTTCAGAAAA ATCGCAGCGT GTATATGAAA GTGTGCAATA CCGATAATGA TTGGCAGGGT 2881 CTGCGTGATA CCGATGAAGG TATGCTGCTG CTGAATGTTC TGAATCAGCG TAATCGCAGC 2941 ATTAAAGAAT ATGCCAGCAA AATTCAGGAT GGTCTGGAAG CAAGCAGCGA ACTGAGCATT 3001 CTGGATTCTG TTGTTCATCT GCATTGCAAT CGTCTGTTTG GTATTGATCG TGAATTTGAA 3061 AAAAAAATTC GCACCCTGGT TGCACATACC CTGTATGCCC TGAAATATAT TAAAGCCCAT 3121 GGCCTGGTGC TGAAAAATTA A 3.8 Codon optimized nucleotide sequence for GeoC 1 ATGTCTATTA GCATGAAAGC CCTGGTTTGT AATCTGGCAA AAAAACTGAG CGATTATGAA 61 AATATGAAAC GCATTGTGAA TCATCCGAGC AATTATATTA AAATTGGCAA CAAAACCATT 121 AATCCGTTTA ATGAACTGAG CCTGAGCCAT GGTCTGCCTG CACTGTGTGC ACTGTATGGT 181 GAACTGAGCG AACAGTATCC GGATGAAGGT TGGGATCTGC TGGGTCATCA GTATATGAAA 241 AAAATTGGCG AACAAATGAA TCAGCATGGT TTTCCGAGCC TGAGCATGTT TACCGGTCTG 301 GCAGGTATTG GTCTGGCAGC AGTTTGTCTG AGCAAAGGTG GTAAACGCTA TCAGAATTTC 361 ATTAGCACCA TTAATCAGGT GATTGAAGAA CGCATTGGCG AAATGATTGA ATATCTGAAA 421 CAGAAACCGT ATCCGAATAT GGATGATTAT GATGCAATTA GCGGTGTTGC AGGTATTGCA 481 AGCTATTGTC TGCTGTTTCC GCAAGAAATG AAAAAAAGCA TTACCCTGAT TCTGAAATAC 541 ATTATTGATC TGAGCCAGGA TAAACAAATG GAAGATATTA ATGTTCCGGG TTGGTATATT 601 CCTCCGATGA ATGCATTTAG CGATACCGAA CGTCGTAAAT GGCCGAGCGG TTTTTTTAAT 661 ATTGGTCTGA GTCATGGCAT TCCGGCACTG CTGATTGTTC TGTGTAATGC CAAAAAACTG 721 AATATTTATG TGCATGGCCA GGATGAATGT ATTCAGCGTA TTGCAGATTT TCTGATGAAA 781 TTTCAGATTA AAGATGAAAA TGGCAGCTAT TGGGGCACCC ATGTTAGCCT GGAAGAATAT 841 AAAAATGGCA GCGTGCTGAA TAAAGATACC CGTGATGCAT GGTGTTATGG TACACCGGGT 901 GTTGCATATA GCCTGCTGAT TGCAGGTAAA ACCCTGAATA ATCAGAGCTA TATTGATTGT 961 GCCGTGAGCG GTATGAAACT GGCAAGCAAA CGTCTGTATA ATATTTTTAG CCCGAGCTTT 1021 TGCCATGGAC TGAGCGGTGT TGCCTATATT TGCAATCGCT TTTATGAAGA AACCAATATT 1081 AGCGATTTTA AAGAAGCAGC CTGCAAACTG GTGGATGATA TTATCAAATT TTATAATGAA 1141 GAATTCCCGT TTGGCTTTAA AAATATTGAA GAAAGCGAAG GCAGCACCAA ATATTATGAT 1201 TATGTGGGTC TGATTGATGG CACCGCAGGT ATTCTGCTGA CCATTCTGGC AATTCAGAAT 1261 AGCAAAAAAA CCCCGTGGGA TTGTGCATTT CTGCTGAGCG AAGTTTAA

Page 121: © 2013 Neha Garg

111  

3.9 References

1. Delves-Broughton, J., Blackburn, P., Evans, R. J., and Hugenholtz, J. (1996) Applications of the bacteriocin, nisin, Antonie van Leeuwenhoek 69, 193-202.

2. Yeaman, M. R., and Yount, N. Y. (2003) Mechanisms of antimicrobial peptide action and resistance, Pharmacol Rev 55, 27-55.

3. Yount, N. Y., and Yeaman, M. R. (2005) Immunocontinuum: perspectives in antimicrobial peptide mechanisms of action and resistance, Protein Pept Lett 12, 49-67.

4. Hsu, S. T., Breukink, E., Tischenko, E., Lutters, M. A., de Kruijff, B., Kaptein, R., Bonvin, A. M., and van Nuland, N. A. (2004) The nisin-lipid II complex reveals a pyrophosphate cage that provides a blueprint for novel antibiotics, Nat Struct Mol Biol 11, 963-967.

5. Breukink, E., Wiedemann, I., van Kraaij, C., Kuipers, O. P., Sahl, H., and de Kruijff, B. (1999) Use of the cell wall precursor lipid II by a pore-forming peptide antibiotic, Science 286, 2361-2364.

6. Hasper, H. E., Kramer, N. E., Smith, J. L., Hillman, J. D., Zachariah, C., Kuipers, O. P., de Kruijff, B., and Breukink, E. (2006) A new mechanism of antibiotic action, Science 313, 1636-1637.

7. Brigham, M. F. (June 9, 2011) ImmuCell comments on status zero milk discard claim for mast out (R)., Available at http://www.reuters.com/article/2011/06/09/idUS120388+09-Jun-2011+MW20110609.

8. Brigham, M. F. (2011) ImmuCell presents mast out product offering at World Animal Health Congress. World Health Congress 2011, Available at http://finance.yahoo.com/news/ImmuCell-Presents-Mast-Out-R-iw-3929215384.html.

9. Chan, W. C., Bycroft, B. W., Lian, L. Y., and Roberts, G. C. K. (1989) Isolation and characterization of two degradation products derived from the peptide antibiotic nisin, FEBS Lett 252, 29-36.

10. Rollema, H. S., Kuipers, O. P., Both, P., de Vos, W. M., and Siezen, R. J. (1995) Improvement of solubility and stability of the antimicrobial peptide nisin by protein engineering, Appl Environ Microbiol 61, 2873-2878.

11. Rollema, H. S., Metzger, J. W., Both, P., Kuipers, O. P., and Siezen, R. J. (1996) Structure and biological activity of chemically modified nisin A species, Eur J Biochem 241, 716-722.

12. Lian, L. Y., Chan, W. C., Morley, S. D., Roberts, G. C., Bycroft, B. W., and Jackson, D. (1992) Solution structures of nisin A and its two major degradation products determined by n.m.r, Biochem J 283, 413-420.

Page 122: © 2013 Neha Garg

112  

13. Cruz, L., Garden, R. W., Kaiser, H. J., and Sweedler, J. V. (1996) Studies of the degradation products of nisin, a peptide antibiotic, using capillary electrophoresis with off-line mass spectrometry, J Chromatogr A 735, 375-385.

14. Olde Riekerink, R. G., Barkema, H. W., and Stryhn, H. (2007) The effect of season on somatic cell count and the incidence of clinical mastitis, J Dairy Sci 90, 1704-1715.

15. Knerr, P. J., Oman, T. J., Garcia De Gonzalo, C. V., Lupoli, T. J., Walker, S., and van der Donk, W. A. (2012) Non-proteinogenic amino acids in lacticin 481 analogues result in more potent inhibition of peptidoglycan transglycosylation, ACS Chem Biol 7, 1791-1795.

16. Chen, L., Walker, D., Sun, B., Hu, Y., Walker, S., and Kahne, D. (2003) Vancomycin analogues active against vanA-resistant strains inhibit bacterial transglycosylase without binding substrate, Proc Natl Acad Sci U S A 100, 5658-5663.

17. Oman, T. J., Lupoli, T. J., Wang, T. S., Kahne, D., Walker, S., and van der Donk, W. A. (2011) Haloduracin alpha binds the peptidoglycan precursor lipid II with 2:1 stoichiometry, J Am Chem Soc 133, 17544-17547.

18. Liu, W., Chan, A. S., Liu, H., Cochrane, S. A., and Vederas, J. C. (2011) Solid supported chemical syntheses of both components of the lantibiotic lacticin 3147, J Am Chem Soc 133, 14216-14219.

19. Feng, L., Wang, W., Cheng, J., Ren, Y., Zhao, G., Gao, C., Tang, Y., Liu, X., Han, W., Peng, X., Liu, R., and Wang, L. (2007) Genome and proteome of long-chain alkane degrading Geobacillus thermodenitrificans NG80-2 isolated from a deep-subsurface oil reservoir, Proc Natl Acad Sci U S A 104, 5602-5607.

20. Marsh, A. J., O'Sullivan, O., Ross, R. P., Cotter, P. D., and Hill, C. (2010) In silico analysis highlights the frequency and diversity of type 1 lantibiotic gene clusters in genome sequenced bacteria, BMC Genomics 11, 679.

21. Zeigler, D. R. (2005) Application of a recN sequence similarity analysis to the identification of species within the bacterial genus Geobacillus, Int J Syst Evol Microbiol 55, 1171-1179.

22. Stein, T. (2008) Whole-cell matrix-assisted laser desorption/ionization mass spectrometry for rapid identification of bacteriocin/lantibiotic-producing bacteria, Rapid Commun Mass Spectrom 22, 1146-1152.

23. Engelke, G., Gutowski-Eckel, Z., Hammelmann, M., and Entian, K.-D. (1992) Biosynthesis of the lantibiotic nisin: genomic organization and membrane localization of the NisB protein, Appl Environ Microbiol 58, 3730-3743.

24. Shi, Y., Yang, X., Garg, N., and van der Donk, W. A. (2011) Production of lantipeptides in Escherichia coli, J Am Chem Soc 133, 2338-2341.

25. Rink, R., Wierenga, J., Kuipers, A., Kluskens, L. D., Driessen, A. J., Kuipers, O. P., and Moll, G. N. (2007) Dissection and modulation of the four distinct activities of nisin by

Page 123: © 2013 Neha Garg

113  

mutagenesis of rings A and B and by C-terminal truncation, Appl Environ Microbiol 73, 5809-5816.

26. Hasper, H. E., de Kruijff, B., and Breukink, E. (2004) Assembly and stability of nisin-lipid II pores, Biochemistry 43, 11567-11575.

27. Wiedemann, I., Breukink, E., van Kraaij, C., Kuipers, O. P., Bierbaum, G., de Kruijff, B., and Sahl, H. G. (2001) 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, 1772-1779.

28. Brötz, H., and Sahl, H. G. (2000) New insights into the mechanism of action of lantibiotics--diverse biological effects by binding to the same molecular target, J Antimicrob Chemother 46, 1-6.

29. Oman, T. J., and van der Donk, W. A. (2009) Insights into the Mode of Action of the Two-Peptide Lantibiotic Haloduracin, ACS Chem Biol 4, 865-874.

30. Field, D., Connor, P. M., Cotter, P. D., Hill, C., and Ross, R. P. (2008) The generation of nisin variants with enhanced activity against specific gram-positive pathogens, Mol Microbiol 69, 218-230.

31. Garg, N., Tang, W., Goto, Y., Nair, S. K., and van der Donk, W. A. (2012) Lantibiotics from Geobacillus thermodenitrificans, Proc Natl Acad Sci U S A 109, 5241-5246.

32. Ross, A. C., Liu, H., Pattabiraman, V. R., and Vederas, J. C. (2010) Synthesis of the lantibiotic lactocin S using peptide cyclizations on solid phase, J Am Chem Soc 132, 462-463.

33. Li, B., Sher, D., Kelly, L., Shi, Y., Huang, K., Knerr, P. J., Joewono, I., Rusch, D., Chisholm, S. W., and van der Donk, W. A. (2010) Catalytic promiscuity in the biosynthesis of cyclic peptide secondary metabolites in planktonic marine cyanobacteria, Proc Natl Acad Sci U S A 107, 10430-10435.

34. Challis, G. L. (2008) Genome mining for novel natural product discovery, J Med Chem 51, 2618-2628.

35. Velásquez, J. E., and van der Donk, W. A. (2011) Genome mining for ribosomally synthesized natural products, Curr Opin Chem Biol 15, 11-21.

36. Kersten, R. D., Yang, Y.-L., Xu, Y., Cimermancic, P., Nam, S.-J., Fenical, W., Fischbach, M. A., Moore, B. S., and Dorrestein, P. C. (2011) A mass spectrometry–guided genome mining approach for natural product peptidogenomics, Nat Chem Biol 7, 794-802.

37. Dischinger, J., Josten, M., Szekat, C., Sahl, H. G., and Bierbaum, G. (2009) Production of the novel two-peptide lantibiotic lichenicidin by Bacillus licheniformis DSM 13, PLoS One 4, e6788.

Page 124: © 2013 Neha Garg

114  

38. Goto, Y., Li, B., Claesen, J., Shi, Y., Bibb, M. J., and van der Donk, W. A. (2010) Discovery of unique lanthionine synthetases reveals new mechanistic and evolutionary insights, PLoS Biol 8, e1000339.

39. Begley, M., Cotter, P. D., Hill, C., and Ross, R. P. (2009) Identification of a novel two-peptide lantibiotic, lichenicidin, following rational genome mining for LanM proteins, Appl Environ Microbiol 75, 5451-5460.

40. McClerren, A. L., Cooper, L. E., Quan, C., Thomas, P. M., Kelleher, N. L., and van der Donk, W. A. (2006) Discovery and in vitro biosynthesis of haloduracin, a two-component lantibiotic, Proc Natl Acad Sci U S A 103, 17243-17248.

41. Lawton, E. M., Cotter, P. D., Hill, C., and Ross, R. P. (2007) Identification of a novel two-peptide lantibiotic, Haloduracin, produced by the alkaliphile Bacillus halodurans C-125, FEMS Microbiol Lett 267, 64-71.

42. Wang, H., Fewer, D. P., and Sivonen, K. (2011) Genome mining demonstrates the widespread occurrence of gene clusters encoding bacteriocins in cyanobacteria, PLoS One 6, e22384.

43. Kabuki, T., Uenishi, H., Seto, Y., Yoshioka, T., and Nakajima, H. (2009) A unique lantibiotic, thermophilin 1277, containing a disulfide bridge and two thioether bridges, J Appl Microbiol 106, 853-862.

44. Brötz, H., Josten, M., Wiedemann, I., Schneider, U., Götz, F., Bierbaum, G., and Sahl, H.-G. (1998) Role of lipid-bound peptidoglycan precursors in the formation of pores by nisin, epidermin and other lantibiotics, Mol Microbiol 30, 317-327.

45. Gut, I. M., Prouty, A. M., Ballard, J. D., van der Donk, W. A., and Blanke, S. R. (2008) Inhibition of Bacillus anthracis spore outgrowth by nisin, Antimicrob Agents Chemother 52, 4281-4288.

Page 125: © 2013 Neha Garg

115

Chapter 4: Discovery, heterologous production, and characterization of geobacillin II, and

in vitro reconstitution of its synthetase enzyme GeoM 1

4.1 Introduction

The class II lanthipeptide synthetases are bifunctional enzymes that carry out the

dehydration of serines and threonines in the core region of the precursor peptide followed by

intramolecular addition of cysteines to the dehydro-amino acids (1). The N-terminal domain of

class II lanthipeptide synthetases catalyzes the phosphorylation of Ser/Thr followed by

elimination of phosphate yielding the Dha/Dhb residues (2-3). The C-terminal domain displays

homology with class I cyclases and catalyzes the formation of the thioether crosslinks between

Dha/Dhb residues and the cysteine side chain (4-5). Removal of the leader peptide by the

bifunctional protease/transporter LanT and, in some cases a second proteolysis step, results in the

production of the final lanthipeptide (3, 6-8). The class II enzymes also occur in the gene clusters

of two-component lantibiotics such as haloduracin (3, 9), cytolysin (6, 10-11) and lacticin 3147

(12-16). Such gene clusters encode for two LanAs and sometimes two LanMs. The modified

molecules act in synergy to produce their biological activities. The stereochemistry at the

thioether crosslinked amino acids of select class I and class II lanthipeptides including nisin (17),

epidermin (18), subtilin (19), Pep5 (20), mersacidin (21), lactocin S (22), and several

prochlorosins (23) was found to be 2S,6R for lanthionine and 2S,3S,6R for methyllanthionine

1Parts of this chapter have been adapted with permission from:

Garg N, Tang W, Goto Y, Nair SK, van der Donk WA,. Proc Natl Acad Sci U S A. 2012 Apr 3; 109(14):5241-6. Reproduced with permission of the publisher and available from http://www.pnas.org/ (DOI: 10.1073/pnas.1116815109).

Page 126: © 2013 Neha Garg

116

(abbreviated hereafter as DL stereoisomers, Figure 4.1A). Recently, a different stereochemical

configuration of the thioether rings for the two-component lantibiotics cytolysin and haloduracin

that are biosynthesized by the class II enzymatic machinery was reported (Figure 4.1) (24).

Cytolysin and haloduracin were shown to contain unusual stereochemistry (2R,6R for lanthionine

and 2R,3R,6R for methyllanthionine) at the N-terminal thioether bridges as opposed to the more

commonly reported DL stereochemistry for lanthipeptides characterized in the past (25). The

methyllanthionine A-ring and lanthionine B-ring of CylLL (Figure 4.1B), the methyllanthionine

A-ring of CylLS (Figure 4.1C), and the methyllanthionine A ring of Halβ all had the LL

stereochemical configuration (Figure 4.1D). Based on the studies performed on cytolysin, it was

suggested that the presence of a Dhx-Dhx-Xxx-Xxx-Cys sequence motif in the precursor

peptide, where Dhx is dehydroalanine (Dha) or dehydrobutyrine (Dhb), and Xxx is any amino

acid except Ser/Thr/Cys, may be dictating the formation of the LL stereochemical configuration

of the thioether crosslinks. Further, a recent report highlighted the biological significance of

unique stereochemistry of the thioether bridges of lacticin 481 (26). Using solid phase support,

analogues of lacticin 481 with alternative stereochemistry were synthesized. The analogues with

non-native stereochemical configuration were shown to lack biological activity.

Reported herein is the discovery and characterization of the lantibiotic geobacillin II and

its class II biosynthetic enzyme GeoM. In vivo of production of geobacillin II in E. coli was

achieved, and the precursor peptide expressed in E. coli was also modified in vitro using the

thermostable lanthipeptide synthetase GeoM at temperatures ranging between 37 °C and 80 °C.

Similar to cytolysin and haloduracin, the LL stereochemical configuration was observed in the A-

ring of geobacillin II. Various analogues of geobacillin II were produced using co-expression of

the precursor peptide GeoAII and GeoM in E. coli. The findings in this study suggest that the

Page 127: © 2013 Neha Garg

117

stereochemical outcome of the A-ring in geobacillin II is not solely dependent on the peptide

sequence as previously suggested for the two-component lantibiotic haloduracin and cytolysin.

Figure 4.1. Structure of lantibiotics with unusual stereochemistry. (A) Chemical structure of DL and LL stereoisomers observed for thioether bridges in lanthipeptides. (B) CylLL". (C) CylLS". (D) Haloduracin β

Page 128: © 2013 Neha Garg

118

4.2 Experimental procedures

4.2.a Construction of pRSFDuet-1/His6-GeoAII+GeoM

Codon-optimized synthetic genes for GeoAII and GeoM (for sequences see the end of

this chapter) were synthesized by GeneArt (Invitrogen, USA). For geobacillin II production,

geoAII was inserted in MCS1 of a pRSFDuet-1 vector using BamHI and HindIII restriction sites

and geoM in MCS2 using NdeI and XhoI sites. A second copy of geoM was provided in a pET15

vector inserted using NcoI and BamHI restriction sites. QuikChange site directed mutagenesis

was performed to change the glycine at position −8 to lysine (pRSFDuet-1/His6GeoAII-

G−8K+GeoM). Two other constructs were made in which glutamine at position −1 was either

mutated to lysine (pRSFDuet-1/His6GeoAIIQ−1K+GeoM) or glutamate (pRSFDuet-

1/His6GeoAII-G−8K/Q−1E+GeoM) (for primers see Table 4.1). Expression and purification of

these constructs was performed as described for geobacillin I in Chapter 3. Briefly,

electrocompetent BL21(DE3) cells were transformed with the appropriate plasmids and a single

colony was picked to inoculate a culture used to overexpress modified GeoAII. Ring A

analogues (GeoAII-S1T, GeoAII-S1T/T2A, GeoAII-S1T/T2A/I3R, GeoAII-G−8K/T2A, and

GeoAII-G−8K/T2G), ring B analogues (GeoAII-G−8K/S7T and GeoAII-G−8K/S7T/L8T), ring

C analogues (GeoAII-G−8K/S14T and GeoAII-G−8K/S14T/L15T), and ring D analogues

(GeoAII-G−8K/S20T and GeoAII-G−8K/S20T/F21T) were generated using QuikChange site

directed mutagenesis (for primers see Table 4.1) and were produced as described above (also see

notebook 5 page 62 and 64).

Page 129: © 2013 Neha Garg

119

4.2.b Cleavage of modified GeoAII and antimicrobial activity assay.

For preparation of geobacillin II, GeoM-modified and HPLC-purified GeoAII-

G−8K/Q−1E (100 µM) was treated with either 0.1 µM trypsin (20 min) or 0.5 µM GluC (2 h).

GeoAII-Q−1K (100 µM) was also treated with 0.1 µM trypsin. The GluC-cleaved GeoAII-

G−8K/Q−1E peptide (25 µL, 100 µM) was further treated with 8 µL of Leu aminopeptidase (97

units/mL, porcine kidney, Sigma L5006) for 24 h at 37 °C. For preparation of ring A analogues

(GeoAII-S1T, GeoAII-S1T/T2A, GeoAII-S1T/T2A/I3R, GeoAII-G−8K/T2A, and GeoAII-

G−8K/T2G), ring B analogues (GeoAII-G−8K/S7T and GeoAII-G−8K/S7T/L8T), ring C

analogues (GeoAII-G−8K/S14T and GeoAII-G−8K/S14T/L15T), and ring D analogues

(GeoAII-G−8K/S20T and GeoAII-G−8K/S20T/F21T), 150 µM peptide solution was incubated

with 0.5 µM GluC for 4 h. Antimicrobial assays were performed as described for nisin in

Chapter 2 (also see notebook 5 page 51, 55, and 75).

4.2.c Tandem mass spectrometry analysis to determine the ring topology of geobacillin II

GeoAII modified by GeoM and treated with GluC was subjected to ESI-MSMS analysis

using collision induced dissociation. For tandem MS, 10 μL of the peptide was injected into a

BEH C8 column (1.7 μm, 1.0 × 100 mm), and separated by UPLC using a gradient of 3% mobile

phase A (0.1% formic acid in water) to 97% mobile phase B (0.1% formic acid in methanol)

over 12 min. The instrument settings used included capillary voltage and cane voltage of 3500 V

and 40 V, respectively, 120 °C as source temperature; 300 °C as desolvation temperature, cone

gas flow of 150 L/h, and desolvation gas flow of 600 L/h. A transfer collision energy of 4 V was

used for both MS and tandem MS, while the trap collision energy was set to 6 V for MS and a

fixed energy of 25 V was used for tandem MS (also see notebook 5 page 55).

Page 130: © 2013 Neha Garg

120

4.2.d GC/MS analysis of modified GeoAII

The stereochemistry of (methyl)lanthionine residues was determined by acid hydrolysis

of geobacillin II and chemical derivatization of the resulting amino acids as previously described

(22-23). The derivatized material was then analyzed by GC/MS. Wild type GeoAII fully

modified by GeoM in E. coli and with its leader attached (1.0 mg) was dissolved in 3 mL of 6 M

HCl. The protocol used for derivatization is described in Chapter 3 (27) (also see notebook 5

page 56).

4.2.e Overexpression and purification of His6-GeoM

Single colony transformants of E. coli Rossetta2 cells containing pET28/His6GeoM

(provided by Dr. Yuki Goto) were grown in a 37 oC shaker for 12-15 h in 30 mL of LB medium

supplemented with 50 μg/mL kanamycin and 25 μg/mL chloramphenicol. These cultures were

used to inoculate 3 L of LB media and grown at 37 oC until the OD600 was ~ 0.6. IPTG was

added to a final concentration of 0.1 mM and the cultures were grown at 18 oC for an additional

12 h. Cells were harvested by centrifugation at 11,900 × g for 25 min at 4 oC. The cell paste was

resuspended in 20 mM Tris-HCl, 1 mM NaCl, and 10% glycerol, pH 8.0. The cells were lysed

using a C3 Homogenizer (Avestin Inc.). The lysed cells were pelleted by centrifugation at 23,700

× g for 25 min. The purification by nickel column chromatography was accomplished using a 5

mL HiTrap chelating HP nickel affinity column (GE Healthcare) on an AKTA FPLC system

(Amersham Pharmacia Biosystems). The supernatant from the centrifugation step was applied to

the column using a 50 mL superloop. The protein was detected by absorbance at 280 nm. The

column was washed with buffer A containing 20 mM Tris-HCl, pH 8.0, 1 M NaCl, and 30 mM

imidazole to remove any non-specifically bound proteins until a flat baseline was observed. A

Page 131: © 2013 Neha Garg

121

gradient of 0-100%B over 40 min was used to elute the bound proteins where buffer B contained

1 M NaCl, 20 mM Tris-HCl, pH 8.0, and 200 mM imidazole. The most concentrated fraction (4

mL) was loaded onto an S-200 gel filtration column (Amersham Biosciences) equilibrated in 20

mM HEPES pH 7.5, 300 mM NaCl, and 10% glycerol. The eluted protein was stored at −80 ˚C

until use (also see notebook 5 page 75).

4.2.f Production and purification of GeoAII

The plasmid used to express and purify GeoAII was provided by a former lab member

Dr. Yuki Goto. Single colony transformants of E. coli BL21 (DE3) containing plasmid pET15/

His6-GeoAII were picked and grown in LB media overnight. The culture after centrifugation and

resuspension in fresh LB media was used to inoculate 12 L of LB media containing 100 mg/L

ampicillin. The culture was shaken at 200 rpm at 37 °C until the OD600 reached between 0.6 and

0.8. At this point, the culture was induced with 1 mM IPTG. The induced cells were shaken

continually at 37 °C for an additional 2 h, and the cells were harvested by centrifugation

(11,900×g for 10 min, Beckman JLA-10.500 rotor). The cell pellet was resuspended in 50 mL of

start buffer (20 mM Tris, pH 8.0, 500 mM NaCl, and 10% glycerol), and cell lysis was carried

out using a MultiFlex C3 homogenizer (Avestin). The lysed cells were centrifuged at 23,700×g

for 30 min at 4 °C. The supernatant was loaded onto a HiTrap HP nickel affinity column (GE

Healthcare) pre-equilibrated with start buffer. The column was washed with wash buffer (start

buffer containing 30 mM imidazole), and the peptide was eluted from the column with elution

buffer (start buffer containing 1 M imidazole) (also see notebook 5 page 42). To increase the

yield, GeoAII was also recovered from the insoluble fraction after cell lysis. The pellet was

homogenized using a sonicator (35% amplitude, 4.4 s pulse, 9.9 s pause for 20 min total) in start

buffer to remove any soluble proteins. The suspension was centrifuged and the pellet was

Page 132: © 2013 Neha Garg

122

homogenized in 30 mL of denaturing buffer (4 M guanidine hydrochloride, 20 mM NaH2PO4,

500 mM NaCl, pH 7.5). The supernatant after centrifugation was loaded onto a HiTrap HP nickel

affinity column, the column was washed with denaturing buffer containing 30 mM imidazole,

and the product was eluted with 15 mL of denaturing buffer containing 1 M imidazole. Small

aliquots of eluent were desalted using a ZipTip (µC18) and analyzed for the presence of GeoAII

using MALDI-MS. For large scale purification, the eluent was desalted using preparative scale

RP-HPLC using a Waters Delta-pak C4 15 μm; 300 Å; 25 ×100 mm PrepPak Cartridge.

Following injection, the column was kept at 2% solvent A (solvent A = 0.086% TFA in 2%

ACN/98% water) for 2 min followed by a gradient of 2-100% of solvent B (0.086% TFA in 80%

ACN/20% water) over 45 min. The fractions containing GeoAII were lyophilized.

4.2.g In vitro enzymatic assay of GeoM

The activity assay contained 20 µM His6-GeoAII, 5 µM His6-GeoM, 20 mM HEPES pH

7.5, 10 mM MgCl2, 2 mM tris(2-carboxyethyl)phosphine (TCEP), and 2.5 mM ATP. The activity

assays were incubated at various temperatures (37 °C, 45 °C, 55 °C, 65 °C, 75 °C, and 80 °C) for

2 h, desalted using µC18 Ziptip devices, and analyzed by MALDI-MS. In order to determine the

effect of temperature on enzyme activity, time dependent in vitro assays were carried out at 37

°C, 55 °C, and 70 °C with 1 µM GeoM. The reactions (20 µL) were quenched with 5 µL of 3%

TFA at 30 min, 60 min, and 2 h. The reaction mixture obtained at different time points was

desalted using µC18 Ziptip and analyzed by MALDI-MS (also see notebook 5 page 64 and 75).

Page 133: © 2013 Neha Garg

123

4.3 Results

4.3.a Description of the gene cluster of geobacillin II from Geobacillus thermodenitrificans

NG80-2

The gene cluster of geobacillin II was annotated in the database as having homology with

the enzymes involved in the biosynthesis of the class II lantibiotic mersacidin produced by

Bacillus sp. HIL Y85,54728 (accession numbers YP_001126158 & YP_001126159). The

cluster of geobacillin II had also been noted in bioinformatic genome mining studies (28), but the

product of the gene cluster was not known. The gene cluster contained a short open reading

frame for the precursor peptide, which we designated geoAII (Figure 4.2A). A search of the non-

redundant protein database for homologs of the precursor peptide GeoAII did not find

homologous precursor peptides of known lantibiotics suggesting it might be a novel class II

lantibiotic (Figure 4.2B and C). A gene encoding a precursor peptide differing in only one amino

acid was found in the genome of Geobacillus sp. G11MC16 (Figure 4.2C). Two start codons are

present in frame in the gene encoding the leader peptide of GeoAII preventing prediction of the

initiation codon and resulting in two possible precursor peptides (Figure 4.2B). GeoAII contains

a typical double glycine motif for proteolytic removal of the leader peptide in class II lantibiotics

(29-30). Leader sequence removal after the double Gly site is supported by the presence of a

gene product (GeoTII) that shows homology with the family of AMS (ABC transporter

maturation and secretion) proteins whose peptide substrates share the double-glycine type

cleavage site (30). Like other family members, GeoTII contains an N-terminal Cys protease

domain for leader peptide removal. The predicted biosynthetic scheme for the production of

geobacillin II based on bioinformatic analysis is given in Figure 4.3. The class II lanthionine

Page 134: © 2013 Neha Garg

124

synthetase GeoM is encoded by a gene located next to geoAII (Figure 4.2A) and has 35%

sequence identity to known LanM enzymes such as MrsM, HalM1 and HalM2.

Figure 4.2. Gene cluster and primary sequence alignment of geobacillin II from G. thermodenitrificans NG80-2. A) Gene cluster for the biosynthesis of geobacillin II. The gene for the precursor peptide (GeoAII) is shown in red, modification enzyme (GeoM) in yellow, and transporter/protease (GeoTII) in purple. B) The two possible Met residues that could be the start of the leader peptide are underlined in the sequence of the leader peptide (top). Sequence alignment of the leader peptides of selected known class II lantibiotics (bottom). C) Sequence alignment of the core region of the precursor peptide GeoAII with the core peptides of class II lantibiotics that undergo a second proteolytic processing step after removal of the leader peptide at the double Gly cleavage site (indicated by a blue arrow, see section 4.3.c). The second cleavage site for this group of lantibiotics is indicated with a red arrow. The ring topology of these compounds, when known, is also shown at the top and the ring topology deduced in this study for geobacillin is shown at the bottom.

Page 135: © 2013 Neha Garg

125

Figure 4.3. Predicted biosynthesis of geobacillin II. The steps leading to incorporation of post-translational modifications in the biosynthesis of lantibiotic geobacillin II. The dehydration and cyclization reactions are catalyzed by the bifunctional enzyme, GeoM. GeoATII is a bifunctional transporter/protease catalyzing transport of modified GeoAII followed by cleavage of the leader peptide. A second proteolysis step catalyzed by an unknown enzyme leads to the synthesis of fully modified geobacillin II (see section 4.3.c).

Page 136: © 2013 Neha Garg

126

4.3.b Heterologous production of geobacillin II in E. coli

Synthetic genes encoding GeoAII and GeoM were inserted into MCS-1 and MCS-2 of a

pRSFDuet-1 vector, respectively. Site-directed mutagenesis was performed to mutate the second

Gly in the double Gly motif of the GeoAII peptide to Lys to incorporate a trypsin cleavage site

for leader peptide removal (GeoAII-G−8K). Co-expression of GeoM with its substrate precursor

peptide in E. coli at 37 °C and subsequent purification as described for geobacillin I in Chapter 3

resulted in five-fold dehydrated peptide (Figure 4.4).

Figure 4.4. Biosynthesis of geobacillin II. MALDI-MS spectrum of GeoAII-G−8K modified by GeoM in E. coli and treated with trypsin.

Page 137: © 2013 Neha Garg

127

4.3.c Bioactivity of geobacillin II

To determine the bioactivity of geobacillin II, purified five-fold dehydrated GeoAII-

G−8K was treated with trypsin and checked for bioactivity. No antimicrobial activity was

observed against any of the strains listed in Table 4.2. Although it is not unprecedented that

lanthionine-containing peptides do not demonstrate any antimicrobial activity (23, 31-33), we

wondered whether perhaps a second proteolytic cleavage event might be required. Removal of

six additional N-terminal residues after leader peptide cleavage at a double glycine site has been

reported previously for several class II lantibiotics (3, 6-7, 34). In some examples, such as

cytolysin from Enterococcus faecalis, removal of these additional residues from the N-terminus

of the modified core peptide is necessary for bioactivity (6), but in other examples (eg

haloduracin (8)) such removal is not required. A sequence alignment of the predicted core

peptide of GeoAII with the core peptides of lantibiotics that are currently known to undergo a

second proteolytic step is shown in Figure 4.2C. The N-terminal amino acids of GeoAII display

only low level sequence homology with the sequences that are removed in the other peptides, but

the presence of a Pro-Gln sequence and the observation that the GeoAI leader peptide is removed

at a Pro-Asn site (Chapter 3) prompted us to investigate whether removal of additional amino

acids might result in bioactivity. First, a trypsin cleavage site was introduced into wild type

GeoAII by mutation of Gln−1 to Lys (this numbering assumes cleavage after ProGln, Fig ure

4.2C). Unfortunately, this mutant was not processed well in E. coli (predominantly 4

dehydrations, Figure 4.5A), presumably because introduction of a positively charged residue is

detrimental for GeoM activity. A different mutant was constructed next (GeoAII-Gln−1Glu) and

co-expressed with GeoM in E. coli resulting in the desired five-fold dehydrated peptide.

Proteolysis with endoproteinase GluC did not result in efficient cleavage after the introduced

Page 138: © 2013 Neha Garg

128

Glu−1 (Figure 4.5B), presumably because the posttranslationally modified Ser at position 1

deactivates the engineered cleavage site. Instead GluC cleaved predominantly after Glu−5

(Figure 4.5B) resulting in the removal of only three residues (Tyr, Thr, and Glu) at the N-

terminus compared to cleavage at the double Gly site (see Figure 4.2C). Interestingly, this GluC-

treated core peptide induced zones of growth inhibition of Bacillus strains in agar well diffusion

assays (Figure 4.5D). Thus, removal of Tyr or Thr (or both) may be required for bioactivity.

Geobacillin II only showed bioactivity against Bacillus species and no activity against any other

tested bacteria (Table 4.2). Attempts to increase the observed bioactivity by treatment with Leu

aminopeptidase (35) to remove additional amino acids (Figure 4.5C) did not result in

significantly increased zones of growth inhibition (compare spots 1 and 2, Figure 4.5D).

Authentic geobacillin II was not produced by the seven Geobacillus strains we evaluated

for the production of geobacillin I (Chapter 3). Thus, it is not clear where the core peptide of

geobacillin II starts. We favor removal of the entire heptapeptide sequence YTEVSPQ after

initial removal of the leader peptide at the double Gly site for the following reasons. Firstly,

removal of this peptide would result from cleavage after Gln similar to the cleavage site for

haloduracin from B. halodurans and similar to lantibiotic proteases such as NisP involved in

nisin biosynthesis (36). A LanP-type protease is not present in the gene cluster or elsewhere in

the genome and therefore, like for haloduracin, the identity of the protease is not known, but the

same protease may remove the leader of GeoAI at its Pro-Asn cleavage site. Secondly, cleavage

after the ProGln sequence would result in an N-terminal structure very similar to that of

haloduracin β, plantaricin β, and cytolysin CylLS (Figure 4.2C).

Page 139: © 2013 Neha Garg

129

Figure 4.5. MALDI-MS spectra of analogues of geobacillin II produced to investigate the antimicrobial activity of geobacillin II. A) MALDI-MS spectrum of GeoAII-Q−1K modified by GeoM in E. coli and treated with trypsin. B) MALDI-MS spectrum of GeoAII-Q−1E modified by GeoM in E. coli and treated with GluC. Peptide 1 arises from cleavage at Glu−5, whereas peptide 5 arises from cleavage at Glu−1 . C) MALDI-MS spectrum of GeoAII-Q−1E modified by GeoM in E. coli and treated first with GluC and then Leu aminopeptidase. Peptide 1 is initially formed upon GluC cleavage. The aminopeptidase then partially removed additional amino acids from the N terminus as shown in the inset. The aminopeptidase treated sample gave the same size of the zone of growth inhibition as pure peptide 1 from panel B suggesting that further removal of the amino acids from the N terminus does not result in more active peptides. D) Agar well diffusion assay with B. subtilis ATCC 6633. Zone 1: 10 µL of 100 µM GeoAII-Q−1E modified by GeoM and treated with GluC. Zones 2 and 3: the material used for zone 1 was treated with aminopeptidase for 24 and 48 h, respectively. Zone 4: 10 µL of 300 µM GeoAII-Q−1K modified by GeoM and treated with trypsin. This sample contains a mixture of four- and five-fold dehydrated peptide (see text and panel A). Zone 5: trypsin-cleaved GeoAII-G−8K. Spot 6: negative control containing GluC and aminopeptidase.

Page 140: © 2013 Neha Garg

130

4.3.d Structure determination of geobacillin II

Tandem MS was used to determine the ring pattern of geobacillin II because the amount

of material obtained from co-expression in E. coli was insufficient for NMR studies. The

precursor peptide GeoAII was co-expressed with GeoM in E. coli, purified by IMAC

chromatography, and treated with GluC to remove most of the leader peptide. The resulting five-

fold dehydrated peptide was fragmented by collision induced dissociation resulting in the

fragment ions indicated in Figure 4.6A and C. The observed ions are inconsistent with

overlapping rings and agree very well with four non-overlapping rings. Based on these data, the

structure shown in Figure 4.6B is proposed for geobacillin II. The structure of in vitro modified

GeoAII was determined by Dr. Yuki Goto and identical ring topology was deduced from those

results. In vitro GeoM-catalyzed modification of GeoAII is described in section 4.3.f.

Page 141: © 2013 Neha Garg

131

Figure 4.6. A) ESI fragmentation pattern of GeoAII modified by GeoM in E. coli and treated with GluC. The b5 and y”26 ions appear to indicate that the A-ring is formed from Cys5 and Dhb2, but three observations argue against this interpretation. First, if the ring were formed between Cys5 and Dhb2, the N-terminal Dha1 would hydrolyze to a pyruvate group (37), decreasing the mass by 1 Da; the masses in Figure 4.5B and C (Table 4.3) do not support this. Second, as shown in Figure 4.2C, A-rings formed from Cys5 and a dehydro amino acid at position 1 are well conserved amongst various lantibiotics, and third, acid hydrolysis of the modified GeoAII peptide followed by derivatization and amino acid analysis by gas chromatography/mass spectrometry only showed the presence of lanthionines for geobacillin AII (section 4.3.d). Thus, we attribute the b5 and y”26 ions to fragmentation in the A-ring. Similar fragmentation is also observed in the A-ring of nisin (see Figure S7 of reference (38). B) Ring topology of geobacillin II derived from the tandem MS results. Arrows indicate the position and direction of lanthionine formation. C) Tandem MS spectrum of GeoAII modified by GeoM in E. coli and treated with GluC. The b and y" ions are marked. Ions marked with asterisks and crosses are not fragment ions but arise from multiply charged ions.

Page 142: © 2013 Neha Garg

132

4.3.e Determination of stereochemistry of lanthionine rings of geobacillin II

Like cytolysin and haloduracin β, ring A of geobacillin II also contains a Dha-Dhb-Xxx-

Xxx-Cys motif and hence, prompted us to determine the stereochemical configuration of the

thioether crosslinks in this molecule. In order to produce sufficient quantities of geobacillin II,

the hexa-histidine tagged precursor peptide GeoAII-G−8K was coexpressed with untagged

GeoM in E. coli (38-39). The modified precursor peptide thus obtained was hydrolyzed and

derivatized for gas chromatography/mass spectrometry (GC/MS) analysis. As predicted,

geobacillin II contained a mixture of DL and LL stereoisomers of lanthionine suggesting that ring

A might have LL stereochemistry (Figure 4.7A). In order to determine unambiguously the

position of the lanthionine with LL stereochemical configuration, the Ser residue at position 1

was mutated to Thr using site directed mutagenesis such that ring A would contain a unique

methyllanthionine. Coexpression of the precursor peptide GeoAII-S1T with GeoM resulted in

five-fold dehydrated and cyclized peptide. The modified GeoAII-S1T was hydrolyzed,

derivatized, and the derivatized amino acids were analyzed by GC/MS. The A-ring, now

methyllanthionine, was found to have the LL configuration (Figure 4.7B) and the lanthionine B,

C, and D rings had the DL configuration (Figure 4.7A). Thus, unlike other characterized

lanthipeptides but similar to cytolysin and haloduracin β, ring A of geobacillin II with the amino

acid sequence Dha-Dhb-Ile-Val-Cys had the LL configuration.

Page 143: © 2013 Neha Garg

133

Figure 4.7. Stereochemical configuration of the thioether bridges of wild type geobacillin II and geobacillin II S1T. (A) GC-MS traces of hydrolyzed and derivatized Lan residues from GeoM-modified His6GeoAII (black), His6GeoAII-S1T (red), DD-Lan standard (blue), DL-Lan standard (pink), and LL- Lan standard (green). (B) GC-MS traces of hydrolyzed and derivatized MeLan residues from GeoM-modified His6GeoAII-S1T (blue), DL-MeLan standard (red), and LL-MeLan standard (black). The small shoulders on the Lan peaks result from partial epimerization during hydrolysis and/or derivatization as previously reported (22).

Page 144: © 2013 Neha Garg

134

To further investigate the importance of the suggested motif in determining the

stereochemical outcome, analogues of geobacillin II were produced with the amino acid

composition of ring A changed to Dha-Gly-Ile-Val-Cys (generated from GeoAII-G−8K/T2G)

and Dha-Ala-Ile-Val-Cys (from GeoAII-G−8K/T2A). Coexpression of the precursor peptides

GeoAII-G−8K/T2G and GeoAII-G−8K/T2A with GeoM resulted in fourfold dehydrated

products. The modified GeoAII-G−8K/T2G and GeoAII-G−8K/T2A peptides were hydrolyzed,

derivatized, and the derivatized amino acids were analyzed by GC/MS. The stereochemistry of

the lanthionine rings in these analogues was found to be a mixture of DL and LL stereoisomers

that was similar to that of the parent sequence (Figures 4.8 and 4.9). Thus, it appeared that the

second Dhx in the sequence motif Dhx-Dhx-Xxx-Xxx is not essential for a stereochemical

outcome of LL configuration of the A-ring of geobacillin II.

Page 145: © 2013 Neha Garg

135

Figure 4.8. GC-MS traces of hydrolyzed derivatized Lan residues of modified His6GeoAII-G−8K/T2A (black), wild type modified His6GeoAII (red, for comparison), DD-Lan standard (blue), DL-Lan standard (pink), and LL- Lan standard (green).

Figure 4.9. GC-MS traces of hydrolyzed derivatized Lan residues of modified His6GeoAII-G−8K/T2G (black) compared with the wild type modified His6GeoAII (red).

Page 146: © 2013 Neha Garg

136

The presence of four additional Lan in geobacillin II and the partial epimerization during

hydrolysis and derivatization confounds the analysis of the stereochemistry of the A-ring. In

order to clearly establish the stereochemical configuration of the ring A lacking the Dhx-Dhx-

Xxx-Xxx-Cys motif, a ring A analogue of geobacillin II with amino acid sequence Dhb-Ala-Ile-

Val-Cys (GeoAII-S1T/T2A) was generated in a similar fashion using coexpression with GeoM

in E. coli to distinguish the derivatized MeLan residue of ring A from derivatized Lan residues

arising from the B, C and D rings. The derivatized methyllanthionine residue originating from

ring A showed a mixture of DL and LL configuration with the LL isomer as the major product

(Figure 4.10). Note that GeoAII-S1T mutant resulted in a product in which the LL-isomer was

favoured more but the DL isomer enriched compared to the wild type sequence (Figure 4.10, blue

trace). This observation suggests that the second Dhx in the sequence Dhx-Dhx-Xxx-Xxx-Cys

motif increases the amount of LL product formed but may not be the only governing factor for a

stereochemical outcome of LL configuration by GeoM.

Page 147: © 2013 Neha Garg

137

Figure 4.10. GC-MS traces of hydrolyzed derivatized MeLan residues of modified His6GeoAII ring A analogues with amino acid sequence Dhb-Ala-Arg-Val-Cys (black), Dhb-Ala-Ile-Val-Cys (red), Dhb-Dhb-Ile-Val-Cys (blue), DL-MeLan standard (pink), and LL-MeLan standard (green).

Page 148: © 2013 Neha Garg

138

Figure 4.11. GC-MS traces of hydrolyzed derivatized Lan residues of modified His6GeoAII ring A analogues with amino acid sequence Dhb-Ala-Arg-Val-Cys (black), Dhb-Ala-Ile-Val-Cys (red), DD-Lan standard (blue), DL-Lan (pink), and LL-Lan standard (green).

The B, C and D rings of geobacillin II all have a positively charged amino acid residue

while ring A does not. To investigate whether the presence of a charged amino acid is (in part)

responsible for the observed stereochemistry, we constructed a ring A analogue with amino acid

sequence Dhb-Ala-Arg-Val-Cys (GeoAII-S1T/T2A/I3R) in order to investigate a possible role of

a charged residue in determining the fate of stereochemistry in the case of GeoM-catalyzed

formation of thioether bridges. Although the abundance of the DL isomer in the GC-MS trace

was lower, the derivatized methyllanthionine residue of this analogue showed a mixture of DL

and LL configuration (Figure 4.10, black trace). As expected, the stereochemistry of all

lanthionine rings in these analogues was DL (Figure 4.11). Thus, in light of results obtained in

Page 149: © 2013 Neha Garg

139

this study, we speculate that in geobacillin II the presence of the sequence Dhx-Dhx-Xxx-Xxx-

Cys motif alone is not enough to produce the LL stereoisomer of the thioether bridge.

In order to investigate whether GeoM can catalyze the formation of LL stereoisomers at

positions other then the A-ring, each of the B, C, and D rings were individually converted to

methyllanthionine rings with amino acid sequences Dhb-Leu-Arg-Ile-Cys (GeoAII-G−8K/S7T)

and Dhb-Dhb-Arg-Ile-Cys (GeoAII-G−8K/S7T/L8T) for ring B, Dhb-Leu-Arg-Phe-Cys

(GeoAII- G−8K/S14T) and Dhb-Dhb-Arg-Phe-Cys (GeoAII-G−8K/S14T/L15T) for ring C, and

Dhb-Phe-Lys-Val-Arg-Cys (GeoAII-G−8K/S20T) and Dhb-Dhb-Lys-Val-Arg-Cys (GeoAII-

G−8K/S20T/F21T) for ring D. The peptides were produced using co-expression with GeoM in

E. coli. Unfortunately, following co-expression, five-fold dehydrated peptides were obtained for

the mutants GeoAII-G−8K/S7T/L8T, GeoAII-G−8K/S14T/L15T, and GeoAII-

G−8K/S20T/F21T as opposed to the desired six-fold dehydrated peptides. A likely explanation

may be the inability of GeoM to catalyze the conversion of the second Thr in the motif Dha-

Dhb-Xxx-Xxx-Cys to Dhb in these rings thus resulting in modified amino acid sequences of the

B, C and D ring analogues as Dhb-Thr-Xxx-Xxx-Cys instead of Dhb-Dhb-Xxx-Xxx-Cys. The

lack of dehydration may possibly occur because of rapid closure of the thioether bridge

preventing access of the threonine residue to the enzyme active site involved in catalyzing the

dehydration reaction. The stereochemistry of the B, C and D ring analogues was determined as

described above. All newly introduced methyllanthionines in these analogues had the DL

configuration (Figures 4.12, 4.13, and 4.14). Similar to wild type geobacillin II, all lanthionines

in the unperturbed A, B, C, and D rings in these analogues contained a mixture of LL and DL

stereochemical configuration (Figures 4.12, 4.13 and 4.14).

Page 150: © 2013 Neha Garg

140

Figure 4.12. Stereochemical configuration of the thioether bridges of geobacillin II B-ring analogues. (A) Amino acid composition of B-ring analogues generated in this study. (B) GC-MS traces of hydrolyzed and derivatized MeLan residues from GeoM-modified His6GeoAII B-ring analogues with amino acid sequence Dhb-Thr-Arg-Ile-Cys (blue), Dhb-Leu-Arg-Ile-Cys (pink), DL-MeLan standard (black), and LL-MeLan standard (red). (C) GC-MS traces of hydrolyzed and derivatized Lan residues from GeoM-modified His6GeoAII B-ring analogues with amino acid sequence Dhb-Thr-Arg-Ile-Cys (blue), and Dhb-Leu-Arg-Ile-Cys (pink).

Page 151: © 2013 Neha Garg

141

Figure 4.13. Stereochemical configuration of the thioether bridges of geobacillin II C-ring analogues. (A) Amino acid composition of C-ring analogues generated in this study. (B) GC-MS traces of hydrolyzed and derivatized MeLan residues from GeoM-modified His6GeoAII C-ring analogues with amino acid sequence Dhb-Thr-Arg-Phe-Cys (blue), Dhb-Leu-Arg-Phe-Cys (pink), DL-MeLan standard (black), and LL-MeLan standard (red). (C) GC-MS traces of hydrolyzed and derivatized Lan residues from GeoM-modified His6GeoAII C-ring analogues with amino acid sequence Dhb-Thr-Arg-Phe-Cys (blue), and Dhb-Leu-Arg-Phe-Cys (pink).

Page 152: © 2013 Neha Garg

142

Figure 4.14. Stereochemical configuration of the thioether bridges of geobacillin II D-ring analogues. (A) Amino acid composition of D-ring analogues generated in this study. (B) GC-MS traces of hydrolyzed and derivatized MeLan residues from GeoM-modified His6GeoAII D-ring analogues with amino acid sequence Dhb-Thr-Lys-Val-Arg-Cys (blue), Dhb-Phe-Lys-Val-Arg-Cys (pink), DL-MeLan standard (black), and LL-MeLan standard (red). (C) GC-MS traces of hydrolyzed and derivatized Lan residues from GeoM-modified His6GeoAII D-ring analogues with amino acid sequence Dhb-Thr-Lys-Val-Arg-Cys (blue), and Dhb-Phe-Lys-Val-Arg-Cys (pink).

Page 153: © 2013 Neha Garg

143

4.3.f Antimicrobial assay of geobacillin II analogues generated in this study

The bioactivity of the geobacillin II analogues produced in this study was tested using an

agar diffusion growth inhibition assay (Figure 4.15). The purified modified peptides were

incubated with GluC in order to remove the leader peptide (see Figures 4.16 and 4.17 and Table

4.3 for mass spectra). GluC cleaved at the −5 position leaving a four amino acid overhang at the

N-terminus of core peptide. This overhang does not inhibit the antibacterial activity of

geobacillin II as described above (39). The cleavage products were tested for biological activity.

All analogues resulted in comparable activity as wild type geobacillin II treated in the same

manner, except the A-ring analogues with amino acid compositions Dhb-Ala-Ile-Val-Cys (zone

13) and Dhb-Ala-Arg-Val-Cys (zone 14) resulting in methyllanthionine bridges, and the B-ring

analogues (zones 4 and 5) constructed in this study (Figure 4.15). The modified core peptide of

the methyllanthionine A-ring analogues (Dhb-Ala-Arg-Val-Cys and Dhb-Ala-Ile-Val-Cys)

eluted as two separate peaks while the core peptide for lanthionine A-ring analogues (Dhb-Ala-

Ile-Val-Cys and Dha-Gly-Ile-Val-Cys) eluted as a single peak on RP-HPLC (Figure 4.18). The

other modified geobacillin II analogues produced in this study also eluted as single peak (Figure

4.19). In order to further investigate the identity of the two peaks, the geobacillin II analogue

GeoAII-S1T/T2A/I3R was produced in larger quantities. The two peaks obtained in RP-HPLC

after GluC cleavage of the corresponding modified peptide were collected separately and

lyophilized. The lyophilized peptides were ran on a RP-HPLC column to determine their purity

(Figure 4.20). The isolated peptides were hydrolyzed and derivatized for GC/MS analysis. Peak

1 contained a mixture of DL and LL stereoisomers of MeLan whereas only a very small signal for

MeLan was observed in the analysis of peak 2 (Figure 4.21). The lack of cyclization of ring A

Page 154: © 2013 Neha Garg

144

was also confirmed by alkylation of free cysteines in these peptides with iodoacetamide (IAA)

using a previously published protocol (3). MALDI-MS spectra showed no IAA adducts were

present in the reaction containing the product of the peptide corresponding to peak 1 whereas one

IAA adduct was present in the reaction containing the product of the peptide corresponding to

peak 2 (Figure 4.22). Similar results upon incubation with IAA were obtained for the geobacillin

II analogue GeoAII-S1T/T2A that also consisted of two peaks by HPLC. The observed

incomplete cyclization upon mutation of these residues in the A-ring suggests that the wild type

sequence of this ring is important for efficient modification by GeoM.

For both analogues, the cyclized material (peak 1) inhibited the growth of B. subtilis

whereas the uncyclized material (peak 2) resulted in no inhibition (Figure 4.23). When the

cyclized and uncyclized material were applied together, no zone of inhibition was observed

(zone 4 and 5, Figure 4.23). This observation suggests that the uncyclized material serves as

antagonist of the fully cyclized peptides. In turn, this finding suggests formation of an oligomeric

active complex involving multiple geobacillin II molecules may be required for antibacterial

activity. For instance, the lantibiotic nisin has been shown to form a multimeric complex in the

cell membrane upon lipid II binding (40). Formation of an inactive complex between fully

cyclized and partially cyclized compounds, and potentially a target molecule may explain the

lack of activity of geobacillin II S1T/T2A/I3R and geobacillin II S1T/T2A. Since ring B

analogues had identical stereochemistry as wild-type geobacillin II and eluted as single peaks,

the reduction in biological activity observed in zones 4 and 5 in Figure 4.15 may arise from

mutation of the Dha residue to a Dhb residue. This mutation may inhibit binding to the target,

which may specifically require a Lan at this position. Alternatively, a Lan B-ring may be

Page 155: © 2013 Neha Garg

145

important for interactions between geobacillin II molecules. Deeper investigation of these results

requires additional mechanism-of-action studies.

Figure 4.15. Agar diffusion growth inhibition assay of geobacillin II analogues produced in this study. Zone 1: nisin (10 µL of 75 µM), Zone 2: nisin (10 µL of 50 µM). Zones 3-14 represent inhibition by geobacillin II and the analogues produced in this study (20 µL of 150 µM solution from GluC cleavage of each modified precursor peptide). Zone 3: geobacillin II, Zone 4: geobacillin II with ring B amino acid sequence Dhb-Leu-Arg-Ile-Cys, Zone 5: geobacillin II with ring B amino acid sequence Dhb-Thr-Arg-Ile-Cys, Zone 6: geobacillin II with ring C amino acid sequence Dhb-Leu-Arg-Phe-Cys, Zone 7: geobacillin II with ring C amino acid sequence Dhb-Thr-Arg-Phe-Cys, Zone 8: geobacillin II with ring D amino acid sequence Dhb-Phe-Lys-Val-Arg-Cys, Zone 9: geobacillin II with ring D amino acid sequence Dhb-Thr-Lys-Val-Arg-Cys, Zone 10: geobacillin II with ring A amino acid sequence Dha-Ala-Ile-Val-Cys, Zone 11: geobacillin II with ring A amino acid sequence Dha-Gly-Ile-Val-Cys, Zone 12: geobacillin II with ring A amino acid sequence Dhb-Dhb-Ile-Val-Cys, Zone 13: geobacillin II with ring A amino acid sequence Dhb-Ala-Ile-Val-Cys, and Zone 14: geobacillin II with ring A amino acid sequence Dhb-Ala-Arg-Val-Cys.

Page 156: © 2013 Neha Garg

146

Figure 4.16. MALDI-MS spectra of modified His6-GeoAII and A-ring analogues generated by coexpression of the corresponding precursor peptide with untagged GeoM in E. coli. The amino acid sequences of the A-rings were: (A) Dha-Dhb-Ile-Val-Cys (wild type), (B) Dhb-Dhb-Ile-Val-Cys, (C) Dhb-Ala-Ile-Val-Cys, (D) Dhb-Ala-Arg-Val-Cys, (E) Dha-Ala-Ile-Val-Cys, and (F) Dha-Gly-Ile-Val-Cys.

Page 157: © 2013 Neha Garg

147

Figure 4.17. MALDI-MS spectra of modified His6-GeoAII analogues of rings B, C, and D generated by coexpression of the corresponding precursor peptide with untagged GeoM in E. coli. The amino acid sequence of the B-ring analogues were: (A) DhbLeuArgIleCys, and (B) Dhb-Thr-Arg-Ile-Cys; analogues of the C-ring were: (C) Dhb-Leu-Arg-Phe-Cys, and (D) Dhb-Thr-Arg-Phe-Cys; and analogues of the D-ring were: (E) Dhb-Phe-Lys-Val-Arg-Cys, and (F) Dhb-Thr-Lys-Val-Arg-Cys.

Page 158: © 2013 Neha Garg

148

Figure 4.18. Analytical HPLC traces of modified His6-GeoAII and A-ring analogues after GluC cleavage. The amino acid sequences of the A-ring were: wild type Dha-Dhb-Ile-Val-Cys (black), Dhb-Dhb-Ile-Val-Cys (red), Dha-Ala-Ile-Val-Cys (blue), Dha-Gly-Ile-Val-Cys (pink), Dhb-Ala-Ile-Val-Cys (green), and Dhb-Ala-Arg-Val-Cys (navy).

Page 159: © 2013 Neha Garg

149

Figure 4.19. Analytical HPLC traces of modified His6-GeoAII and analogues of the B, C, and D rings after GluC cleavage. The amino acid sequences of the B-ring were: Dhb-Leu-Arg-Ile-Cys (green), and Dhb-Thr-Arg-Ile-Cys (navy). The amino acid sequences of the C-ring sequences were: Dhb-Leu-Arg-Phe-Cys (blue), and Dhb-Thr-Arg-Phe-Cys (pink). The amino acid sequences of the D-ring were: Dhb-Phe-Lys-Val-Arg-Cys (black), and Dhb-Thr-Lys-Val-Arg-Cys (red).

Page 160: © 2013 Neha Garg

150

Figure 4.20. Analytical HPLC trace of modified wild type His6-GeoAII-S1T/T2A/I3R (blue) after GluC cleavage. Peak 1 and Peak 2 were collected separately and lyophilized. The lyophilized powder was dissolved in water and analyzed by analytical HPLC. The black and red traces correspond to purified peak 1 and peak 2.

Page 161: © 2013 Neha Garg

151

Figure 4.21. Stereochemical configuration of hydrolyzed and derivatized Lan/MeLan residues of peptides present in peak 1 and peak 2 (Figure 4.20). (A) Hydrolyzed and derivatized MeLan residue in sample generated from peak 1 (black trace) and peak 2 (red trace). (B) Hydrolyzed and derivatized Lan residue in samples generated from peak 1 (black trace) and peak 2 (red trace).

Page 162: © 2013 Neha Garg

152

Figure 4.22. MALDI-TOF mass spectra of peptides present in peak 1 and peak 2 after treatment with 5 mM TCEP and 50 mM iodoacetamide. The peak corresponding to M-4 H2O + 9 is generated by one IAA adduct of dethiomethyl geobacillin II. Dethiomethyl geobacillin II is generated by loss of part of the side chain of an oxidized methionine (UNIMOD accession number 526) or by the reaction of the methionine side chain with IAA, followed by elimination resulting in a mass shift of −48 Da.

Page 163: © 2013 Neha Garg

153

Figure 4.23. Bioassay with the indicator strain B. subtilis ATCC6633. Zone 1: 20 µL of GluC treated geobacillin II-G−8K/T2A (150 µM). Zone 2: 10 µL of geobacillin II-S1T/T2A/I3R corresponding to peak 2 (150 µM) + 10 µL sterile water. Zone 3: 10 µL of geobacillin II-S1T/T2A/I3R corresponding to peak 1 (150 µM) + 10 µL sterile water. Zone 4: 10 µL each of geobacillin II-S1T/T2A/I3R corresponding to peak 1 and 2 (150 µM each). Zone 5: 10 µL each of geobacillin II-S1T/T2A corresponding to peak 1 and 2 (150 µM each). Zone 6: 10 µL of geobacillin II-S1T/T2A corresponding to peak 1 (150 µM) + 10 µL sterile water. Zone 7: 10 µL of geobacillin II-S1T/T2A corresponding to peak 2 (150 µM) + 10 µL sterile water.

4.3.g In vitro reconstitution of enzymatic activity of GeoM

Promiscuous thermostable lanthionine synthetases may provide promising candidates for

generation of cyclic peptide libraries. Towards this aim, we sought to characterize the enzymatic

activity of GeoM. The genes encoding the precursor peptide GeoAII and the modification

enzyme GeoM were heterologously expressed as recombinant His6-tagged fusion proteins in E.

coli by former lab member Dr. Yuki Goto. The precursor peptide was purified using nickel

affinity chromatography followed by reverse phase high performance liquid chromatography

(RP-HPLC). The enzyme GeoM was purified using nickel affinity chromatography followed by

gel filtration chromatography. In vitro activity assays were carried out at 37 °C, 45 °C, 55 °C, 65

°C, 75 °C, and 80 °C for 2 h in the presence of ATP, Mg2+, TCEP, and 5 µM enzyme.

Commercial GluC protease was used for the removal of most of the leader peptide. GluC cleaved

Page 164: © 2013 Neha Garg

154

at the −5 position leaving a four amino acid overhang at the N-terminus of the core peptide. The

GluC cleavage product was analyzed by MALDI-MS analysis. The temperature-activity analysis

showed that GeoAII was processed by GeoM in 2 h at all temperatures demonstrating the

thermostability of the enzyme (Figure 4.24). The efficiency of modification of GeoAII by GeoM

was investigated at 37 °C, 55 °C, and 70 °C with 1 µM enzyme (a five-fold decrease compared

to Figure 4.24). The reaction was analyzed at 30 min, 60 min, and 2 h for each temperature using

MALDI-MS (Figure 4.25). The modification efficiency increased with increase in temperature.

Full conversion was observed by 2 h at 55 °C and 60 min at 70 °C whereas a significant amount

of starting material was left by 2 h at 37 °C.

Figure 4.24. In vitro reconstituition of GeoM activity. MALDI-MS spectra of His6GeoAII incubated with His6GeoM (5 µM), ATP (2.5 mM), Mg2+ (10 mM), and TCEP (2 mM) in HEPES pH 7.5 and cleaved with GluC.

Page 165: © 2013 Neha Garg

155

Figure 4.25. Temperature-time dependence of GeoM in vitro activity. MALDI-MS spectra of His6GeoAII incubated with His6GeoM (1 µM), ATP (2.5 mM), Mg2+ (10 mM), and TCEP (2 mM) at 70 °C, 55 °C, and 37 °C at time points (A) 30 min, (B) 60 min, and (C) 2 h.

Page 166: © 2013 Neha Garg

156

4.4 Discussion

A novel class II lantibiotic, geobacillin II, was discovered and characterized in

this study. The structure of geobacillin II revealed two unique features. First, the ring topology of

geobacillin II is different from known class II lantibiotics. Second, similar to cytolysin and

haloduracin, the first lanthionine ring (ring A) of geobacillin II has unusual LL stereochemistry.

The B, C, and D rings have the more common DL stereochemistry. It seems likely that the

stereochemical outcome of the ring A analogues generated in this study does not depend only on

the sequence motif Dhx-Dhx-Xxx-Xxx-Cys. The findings in this chapter suggests that more

complex factors such as amino acid sequence and/or differential binding in the enzyme active

site, and perhaps even some features of the leader peptide might play a concerted role in

determining the stereochemistry of the thioether bridges synthesized by GeoM. A role of the

leader peptide is suggested by the position of nearly all (methyl)lanthionines with LL-

stereochemistry at the N-termini of the mature lantibiotics. Although, at present we cannot

determine the precise location of the start of geobacillin II, the demonstration that removal of

additional amino acids past the double Gly protease cleavage site resulted in bioactivity along

with the sequence homology with other lantibiotics that undergo a second step of proteolytic

processing support the lantibiotic structure shown in Figure 4.7. Geobacillin II had a narrow

spectrum of antibacterial activity suggesting a different mechanism of action when compared to

known lantibiotics. Future work should focus on characterizing the mode-of-action of geobacillin

II.

The lantibiotic synthetase GeoM is a thermostable enzyme and is active up to at least

temperatures of 80 °C. Phosphorylated intermediates were observed both in vivo and in vitro

(Figure 4.5C and 4.24) suggesting that dehydro-amino acids are installed via phosphorylation of

Page 167: © 2013 Neha Garg

157

Ser/Thr followed by subsequent elimination of the phosphate. Comparison of enzyme activity at

37 °C, 55 °C, and 70 °C demonstrated higher dehydration efficiency with increase in

temperature.

Page 168: © 2013 Neha Garg

158

4.5 Tables Table 4.1. Sequences of oligonucleotide primers used in this study.

Primer name Primer sequence GeoAII_BamHI_F CTA GAT GGA TCC GAT GAA AGG TGG CAT TCA GAT GG GeoAII_HindIII_R CTA GAA GCT TGeoM_NdeI_F

TT ACA TCG GAC AAC GAA CTT TAA AGC AAG CAG CCG CAT ATG

GeoM_XhoI_R AAC GAA ATC GTG GAA AAT AAC C

CTA GCT CGA GGeoAII(G-8K)_F

TT AAT GGT TCA GCT GCA GAG TCA GCA CG GAA AAA ACT GGC TGG CAA ATA TAC CGA AGT TTC TCC G

GeoAII(G-8K)_R GCC AGC CAG TTT TTT CAG TTC TTC TTC GC GeoAII(Q-1E)_F GTT ATA CCG AAG TTT CTC CG GAAAGCA CCA TTG TTT GTG GeoAII(Q-1E)_R CGG AGA AAC TTC GGT ATA TTT GCC AGC CAG GeoAII(Q-1K)_F GTT ATA CCG AAG TTT CTC CGA AAA GCA CCA TTG TTT GTG GeoAII(Q-1K)_R CGG AGA AAC TTC GGT ATA ACC GCC AGC CAG

GeoAII(S1T)_F AAT ATA CCG AAG TTT CTC CGC AGA CCA CCA TTG TTT GTG TTA GCC

GeoAII(S1T)_R CTGCGGAGAAACTTCGGTATATTTGCCAGC GeoAII(G−8K/T2A)_F ACC GAA GTT TCT CCG CAG AGC GCG ATT GTT TGT GTT AGC GeoAII(G−8K/T2A) _R GCTCTGCGGAGAAACTTCGGTATAACCGCC GeoAII(G−8K/T2G)_F CGA AGT TTC TCC GCA GAG CGG CAT TGT TTG TGT TAG CCT G

GeoAII(G−8K/T2G)_R CAG GCT AAC ACA AAC AAT GCC GCT CTG CGG AGA AAC TTC G

GeoAII(S1T/T2A)_F ACC GAA GTT TCT CCG CAG ACC GCG ATT GTT TGT GTT AGC GeoAII(S1T/T2A)_R GGTCTGCGGAGAAACTTCGGTATAACCGCC GeoAII(S1T/T2A/I3R)_F TTT CTC CGC AGA CCG CGC GTG TTT GTG TTA GCC TG GeoAII(S1T/T2A/I3R)_R ACG CGC GGT CTG CGG AGA AAC TTC GGT ATA AC GeoAII(G−8K/S7T)_F CAG AGC ACC ATT GTT TGT GTT ACC CTG CGT ATT TGT AAT

TGG GeoAII(G−8K/S7T)_R AAC ACA AAC AAT GGT GCT CTG CGG AGA AAC GeoAII(G−8K/S7T/L8T)_F CAG AGC ACC ATT GTT TGT GTT ACC ACA CGT ATT TGT AAT

TGG GeoAII(G−8K/S7T/L8T)_F GGT AAC ACA AAC AAT GGT GCT CTG CGG AGA AAC GeoAII(G−8K/S14T)_F CCT GCG TAT TTG TAA TTG GAC CCT GCG TTT TTG TCC GAG GeoAII(G−8K/S14T)_R CTC GGA CAA AAA CGC AGG GTC CAA TTA CAA ATA CGC AGG GeoAII(G−8K/S14T/L15T)F CCT GCG TAT TTG TAA TTG GAC CAC ACG TTT TTG TCC GAG GeoAII(G−8K/S14T/L15T)R GGT CCA ATT ACA AAT ACG CAG GCT AAC ACA AAC GeoAII(G−8K/S20T)_F GGA GCC TGC GTT TTT GTC CGA CCT TTA AAG TTC GTT GTC GeoAII(G−8K/S20T)_R CGG ACA AAA ACG CAG GCT CCA ATT ACA AAT AC GeoAII(G−8K/S20T/F21T)_F GGA GCC TGC GTT TTT GTC CGA CCA CAA AAG TTC GTT GTC GeoAII(G−8K/S20T/F21T)_R GAC AAC GAA CTT TTG TGG TCG GAC AAA AAC GCA GGC TCC

Page 169: © 2013 Neha Garg

159

Table 4.2. Bioactivity spectrum of geobacillin II assessed by zones of growth inhibition in agar well diffusion assays.

Strain Source Geobacillin II

Micrococcus luteus ATCC 4698 -

Streptococcus dysgalatiae subsp dysgalactiae ATCC 27957 -

Staphylococcus epidermidis 15X

Ekkelenkamp et. al.a -

Staphylococcus epidermidis ATCC12228 -

Bacillus cereus Z4222 INRAb Z4222 +

Bacillus subtilis ATCC 6633 +

Lactococuus lactis HP ATCC 11602 -

E. coli DH5α UIUC-CMFc - aReference (41). bINRA: Institut National de la Recherche Agronomique. cUIUC-CMF: University of Illinois Urbana-Champaign Cell and Media Facility.

Page 170: © 2013 Neha Garg

160

Table 4.3. Theoretical and observed molecular weights ([M+H]1+) of modified peptides reported in this study.

Peptide Protease

used Theoretical

[M+H]1+ Observed [M+H]1+

Geobacillin II (from GeoA-G−8K) GluC 3470, 3057 3470, 3057

Geobacillin II (from GeoA-G−8K) GluC and aminopeptidase

3470,3371, 3187,3057

3470,3371, 3187,3058

Geobacillin II (from GeoA-G−8K/Q−1E) Trypsin 3863 3863

Geobacillin II (from GeoA-Q−1K) Trypsin 3057, 3075 3057, 3075

Geobacillin II (S1T) GluC 3479 3479

Geobacillin II (S1T/T2A) GluC 3468 3468

Geobacillin II (S1T/T2A/I3R) GluC 3511 3511

Geobacillin II (G−8K/T2G) GluC 3440 3440

Geobacillin II (G−8K/T2A) GluC 3454 3454

Geobacillin II (G−8K/S7T) GluC 3480 3480

Geobacillin II (G−8K/S7T/L8T) GluC 3550 3467

Geobacillin II (S14T) GluC 3480 3480

Geobacillin II (G−8K/S14T/L15T) GluC 3450 3467

Geobacillin II (G−8K/S20T) GluC 3480 3480

Geobacillin II (G−8K/S20T/F21T) GluC 3416 3439

Page 171: © 2013 Neha Garg

161

4.6 Codon optimized nucleotide sequence for GeoAII

1 ATGAAAGGTG GCATTCAGAT GGAAAAACAG GAACAGACCT TTGTTAGCAA AATTAGCGAA 61 GAAGAACTGA AAAAACTGGC TGGCGGTTAT ACCGAAGTTT CTCCGCAGAG CACCATTGTT 121 TGTGTTAGCC TGCGTATTTG TAATTGGAGC CTGCGTTTTT GTCCGAGCTT TAAAGTTCGT 181 TGTCCGATGT AA

4.7 Codon optimized nucleotide sequence for GeoM

1 ATGAACGAAA TCGTGGAAAA TAACCTGATG GAAATCATCA AAGATAAACT GCTGTACCCG 61 AGCGAAAAAA AAAAAGAAAT CGAAAACAAC GTGCCGATTA ACGAGCTGCT GTTCCAAGAA 121 AAAACCAACA AATGGATCGA CCTGTTCTAT CAGTATCTGA ACGTGTGCAA AAACAACATT 181 GAAGTGGATC AGCTGTATAG CAACGAAGAA AACCTGAGCA ATCCGTTTAG CAGCTTTTTC 241 AAAGAATTTC TGGAAATTGC GAGCCGTTAT CTGGATAGCA GCCTGAAAAG CCAGGTGAAC 301 AACTATGATG ATCTGAGCCG TATTTGCAAC CTGACCAGCA TCAAAAATGA TATCATGAAA 361 TTCATCCACG AAAACCTGGT GTATGTGAGC ATTCGTACCC TGATTCAGGA TCTGAATGAA 421 GAACGTGAAA AAGGCAACCT GATTGGCGAT ACCCCGAAAG AACGCTATAA CTTCTACGTG 481 GATCAGATTC TGAGCAGCCC GGATAAAAAA TTCGAGCTGA TCAAAAAATA TCCGGTGCTG 541 GTGCGTATTC TGATTGAATT TATCCTGAAC AAAATCGATA GCATCGTGGA AAGCATTTAC 601 CGCTTTCTGA AAGATCGTAG CGAGCTGGTG CATATTTTCC ACCTGAGCGA TGATGATATT 661 CTGACCAGCC TGACTCTGCA GAGCGGCGAT AGCCATAACA ACGGCCGTAG CGTGATCATT 721 TTTCAGTTCA GCAGCAAAAA AAAAATCGTG TACAAACCGC GTAGCCTGAG CATTGATCTG 781 CATTTTCAGC AGTTTCTGGA ATGGATCAAC GGCAAAAAAC CGAGCCTGCA GCTGAAAACG 841 ATTACCATTC TGAACAAAGA TCAGTATGGC TGGCAGGAAT TTGTTGAATA TAAACCGTGC 901 AGCTCTAACA ACGAGCTGTC TCGTTTTTAT GAACGTCAGG GCAACTATAT TGCGATTCTG 961 TATATTCTGA ACGCGACCGA TTTTCATTTT GAAAACCTGA TTGCGAACGG CGAACACCCG 1021 GTGCTGATTG ATCTGGAAGG CCTGGTGCAG AACACCGTGA AACTGCCGCG TAAAGCGAGC 1081 AGCGCGTATG ATATTGCGTT TAGCAAACTG ACCGATAGCG TTCTGAGCAC CGGCATGCTG 1141 CCGGCGACCT TTATGCAGGC GAATATCTAT GATTTTGATC TGAGCGGCCT GGGCGGTGAT 1201 GAAGGTCAGC CGACCGGCCT GGAAACCTTT ACCATTGAAA ATCCGCTGAC CGATGAAATG 1261 CGTGTGATTA AAGTGCCGGC GTTTAGCCAG AGCAGCAAAA ACAAACCGTA TCTGAAAGAA 1321 GAAAAAGAAA TTGCGGTGAC CGATTATAGC AGCGAAATCA TTAAAGGCTT CCGCGAAATG 1381 TATACCCTGC TGCTGCATAA CAAAAAGGAG CTGCTGTCTC AGAACGGCCC GATTTACCTG 1441 TTTAAAGGCG ATAAAGTGCG TATTATTCTG CGTAGCACCC AGGTGTATAG CACCTTCCTG 1501 GATAGCAGCT TTCATCCGGA TTATCTGAAA GATGGCTATG AACGTGAACG CCTGATTAAC 1561 TTTCTGTGGA TCGGCAAAGA AAACCATCCG GAATATCAGG ATGCGATCAT GTATGAATGC 1621 CGCGATATTC TGAATGGCGA TATCCCGTAT TTCTATTGCT ATACCGATAG CAGCGATCTG 1681 TATCATCCGA TTGGCATCGT GAAACACAAC TTTTTCTATG AAAGCAGCTT TAACAGCCTG 1741 CTGGAAAAAG TGAAAATGAT CAACGAAGAA GATCTGAATT TCCAGATCGA AATTCTGACC 1801 AATAGCCTGC TGGCCCAGTA TAGCAACAAA ACCCATAGCC ATGCGAACGT GAGCAACCGT 1861 GTGTATAACC TGGATAAAAT CAGCGGCAAC TTTCGTCGTG AATCTTTTCT GGAAGTGAGC 1921 GAAAAAATTG CGGATAACAT CAAAGAAAAC GCGATCTTCG GTAAAGAAAA CGATGTGACC 1981 TGGCTGGGCC TGAACCTGAC CATTGAAGAT AAATGGACCT TCAAACCGCT GGATTTCGAT 2041 CTGTATGATG GCGTGCTGGG CATTGGCCTG TTTTATGCGA ACCTGTATAA CCTGAACAAA 2101 CGCAAAGAAT ACAAAATCCT GGCCGAAAAA ACCGTGCAGA CCGCGCTGAA CTATCTGGAA 2161 TATTATCCGG CGAAACTGCC GCTGTCTGCG TTTTATGGCT ATGGCGCGTA TGCGTATGTG 2221 CTGGGCAACT TTAGCATCAT CTTTAATAAC AGCGAATACC TGACCTATGT GAAAAAAGTG 2281 CTGAACAAAG CGGCGAACAC CATCGAAGAT GATCAGCTGC TGGATTTTCT GGGCGGTGCG 2341 GCGGGTCTGA TTATTGTGTG CATCCACCTG TATAAAAAAA CCGGCGAACA GTATCTGCTG 2401 GATATTGCGA ACAAATGCGG TGAGCTGCTG CTGCGTAAAA AAGAAAAACA GGCCATGGGC 2461 ATTGGCTGGC GTCCGAACCA GGTGAATAAA CCGCTGGCCG GTCTGGCCCA TGGTAGCAGC 2521 GGCTTTACCT GGGCGCTGAT GGCGCTGTAT AAATTCACCA AAAACTACAA CTATAAAGAA 2581 ACCTTCCTGC AGAGCTTCGA ATATGAAAAA AGCCTGTTTA ATAAAAAAGA AGAAAACTGG

Page 172: © 2013 Neha Garg

162

2641 CTGGTTCTGA GCGACGATGG CGATTATTAT GGCAACAGCA TGTGGTGCCA TGGTGCGGCC 2701 GGTATTGGCA TGAGCCGTAT CATGATGAGC GAATACTATA ACAGCCACGA TCTGAAAAAC 2761 GATATTGAAA TCGCGATTCG TCAGACCCTG AAAAGCGGCT TTGGCGGCAC CCATTGCCTG 2821 TGCCATGGCG ATCTGGGCAA CCTGGACCTG TTTCTGCTGG CCTCTAGCAA ATTTAAAAAC 2881 GAAGAAATGA AAGAAACCGC GCTGAAAATT GGCGAATATA TCGCGAGCGA TATTACCTAT 2941 GGCAACCTGA AATATGGCAC CCCGAGCGGC ACCAAAAATC TGGGTCTGAT GCTGGGCCAG 3001 GCGGGTATTG GCTATGGCTT TCTGCGTCTG GCCTATCCGG AAATTGTGCC GAGCGTGCTG 3061 ACTCTGCAGC TGAACCATTA A

Page 173: © 2013 Neha Garg

163

4.8 References

1. Xie, L., Miller, L. M., Chatterjee, C., Averin, O., Kelleher, N. L., and van der Donk, W. A. (2004) Lacticin 481: in vitro reconstitution of lantibiotic synthetase activity, Science 303, 679-681.

2. You, Y. O., and van der Donk, W. A. (2007) Mechanistic investigations of the dehydration reaction of lacticin 481 synthetase using site-directed mutagenesis, Biochemistry 46, 5991-6000.

3. McClerren, A. L., Cooper, L. E., Quan, C., Thomas, P. M., Kelleher, N. L., and van der Donk, W. A. (2006) Discovery and in vitro biosynthesis of haloduracin, a new two-component lantibiotic, Proc Natl Acad Sci USA 103, 17243-17248.

4. Li, B., Yu, J.-P. J., Brunzelle, J. S., Moll, G. N., van der Donk, W. A., and Nair, S. K. (2006) Structure and Mechanism of the Lantibiotic Cyclase Involved in Nisin Biosynthesis, Science 311, 1464-1467.

5. Paul, M., Patton, G. C., and van der Donk, W. A. (2007) Mutants of the Zinc Ligands of Lacticin 481 Synthetase Retain Dehydration Activity but Have Impaired Cyclization Activity, Biochemistry 46, 6268-6276.

6. Cox, C. R., Coburn, P. S., and Gilmore, M. S. (2005) Enterococcal cytolysin: a novel two component peptide system that serves as a bacterial defense against eukaryotic and prokaryotic cells, Curr Protein Pept Sci 6, 77-84.

7. Caetano, T., Krawczyk, J. M., Mosker, E., Süssmuth, R. D., and Mendo, S. (2011) Heterologous expression, biosynthesis, and mutagenesis of type II lantibiotics from Bacillus licheniformis in Escherichia coli, Chem Biol 18, 90-100.

8. Cooper, L. E., McClerren, A. L., Chary, A., and van der Donk, W. A. (2008) Structure-Activity Relationship Studies of the Two-Component Lantibiotic Haloduracin, Chem Biol 15, 1035-1045.

9. Lawton, E. M., Cotter, P. D., Hill, C., and Ross, R. P. (2007) Identification of a novel two-peptide lantibiotic, Haloduracin, produced by the alkaliphile Bacillus halodurans C-125, FEMS Microbiol Lett 267, 64-71.

10. Gilmore, M. S., Segarra, R. A., Booth, M. C., Bogie, C. P., Hall, L. R., and Clewell, D. B. (1994) Genetic structure of the Enterococcus faecalis plasmid pAD1-encoded cytolytic toxin system and its relationship to lantibiotic determinants, J Bacteriol 176, 7335-7344.

11. Lawton, E. M., Ross, R. P., Hill, C., and Cotter, P. D. (2007) Two-peptide lantibiotics: a medical perspective, Mini Rev Med Chem 7, 1236-1247.

12. Ryan, M. P., Rea, M. C., Hill, C., and Ross, R. P. (1996) An application in cheddar cheese manufacture for a strain of Lactococcus lactis producing a novel broad-spectrum bacteriocin, lacticin 3147, Appl Environ Microbiol 62, 612-619.

Page 174: © 2013 Neha Garg

164

13. Martin, N. I., Sprules, T., Carpenter, M. R., Cotter, P. D., Hill, C., Ross, R. P., and Vederas, J. C. (2004) Structural Characterization of Lacticin 3147, a Two-Peptide Lantibiotic with Synergistic Activity, Biochemistry 43, 3049-3056.

14. Kuipers, A., Meijer-Wierenga, J., Rink, R., Kluskens, L. D., and Moll, G. N. (2008) Mechanistic dissection of the enzyme complexes involved in biosynthesis of lacticin 3147 and nisin, Appl Environ Microbiol 74, 6591-6597.

15. Silkin, L., Hamza, S., Kaufman, S., Cobb, S. L., and Vederas, J. C. (2008) Spermicidal bacteriocins: lacticin 3147 and subtilosin A, Bioorg Med Chem Lett 18, 3103-3106.

16. Suda, S., Cotter, P. D., Hill, C., and Ross, R. P. (2012) Lacticin 3147--biosynthesis, molecular analysis, immunity, bioengineering and applications, Curr Protein Pept Sci 13, 193-204.

17. Gross, E., and Morell, J. L. (1971) The Structure of Nisin, J Am Chem Soc 93, 4634-4635.

18. Allgaier, H., Jung, G., Werner, R. G., and Schneider, U. (1985) Elucidation of the structure of epidermin, a ribosomally synthesized, tetracyclic hetrodetic polypeptide antibiotic, Angew Chem, Int Ed Engl 24, 1051-1053.

19. Gross, E., Kiltz, H. H., and Nebelin, E. (1973) Subtilin. VI. Structure of subtilin, Hoppe-Seyler's Z Physiol Chem 354, 810-812.

20. Kellner, R., Jung, G., Josten, M., Kaletta, C., Entian, K. D., and Sahl, H. G. (1989) Pep5: structure elucidation of a large lantibiotic, Angew Chem 101, 618-621.

21. Chatterjee, S., Chatterjee, S., Lad, S. J., Phansalkar, M. S., Rupp, R. H., Ganguli, B. N., Fehlhaber, H. W., and Kogler, H. (1992) Mersacidin, a new antibiotic from Bacillus. Fermentation, isolation, purification and chemical characterization, J Antibiot 45, 832-838.

22. Ross, A. C., Liu, H., Pattabiraman, V. R., and Vederas, J. C. (2010) Synthesis of the lantibiotic lactocin S using peptide cyclizations on solid phase, J Am Chem Soc 132, 462-463.

23. Li, B., Sher, D., Kelly, L., Shi, Y., Huang, K., Knerr, P. J., Joewono, I., Rusch, D., Chisholm, S. W., and van der Donk, W. A. (2010) Catalytic promiscuity in the biosynthesis of cyclic peptide secondary metabolites in planktonic marine cyanobacteria, Proc Natl Acad Sci U S A 107, 10430-10435.

24. Tang, W., and van der Donk, W. A. (2013) The sequence of the enterococcal cytolysin imparts unusual lanthionine stereochemistry, Nat Chem Biol 9, 157-159.

25. Chatterjee, C., Paul, M., Xie, L., and van der Donk, W. A. (2005) Biosynthesis and Mode of Action of Lantibiotics, Chem Rev 105, 633-684.

Page 175: © 2013 Neha Garg

165

26. Knerr, P. J., and van der Donk, W. A. (2013) Chemical synthesis of the lantibiotic lacticin 481 reveals the importance of lanthionine stereochemistry, J Am Chem Soc 135, 7094-7097.

27. Liu, W., Chan, A. S., Liu, H., Cochrane, S. A., and Vederas, J. C. (2011) Solid supported chemical syntheses of both components of the lantibiotic lacticin 3147, J Am Chem Soc 133, 14216-14219.

28. Marsh, A. J., O'Sullivan, O., Ross, R. P., Cotter, P. D., and Hill, C. (2010) In silico analysis highlights the frequency and diversity of type 1 lantibiotic gene clusters in genome sequenced bacteria, BMC Genomics 11, 679.

29. Nes, I. F., and Tagg, J. R. (1996) Novel lantibiotics and their pre-peptides, Antonie van Leeuwenhoek 69, 89-97.

30. Håvarstein, L. S., Diep, D. B., and Nes, I. F. (1995) A family of bacteriocin ABC transporters carry out proteolytic processing of their substrates concomitant with export, Mol Microbiol 16, 229-240.

31. Kodani, S., Hudson, M. E., Durrant, M. C., Buttner, M. J., Nodwell, J. R., and Willey, J. M. (2004) The SapB morphogen is a lantibiotic-like peptide derived from the product of the developmental gene ramS in Streptomyces coelicolor, Proc Natl Acad Sci USA 101, 11448-11453.

32. Kodani, S., Lodato, M. A., Durrant, M. C., Picart, F., and Willey, J. M. (2005) SapT, a lanthionine-containing peptide involved in aerial hyphae formation in the streptomycetes, Mol Microbiol 58, 1368-1380.

33. Goto, Y., Li, B., Claesen, J., Shi, Y., Bibb, M. J., and van der Donk, W. A. (2010) Discovery of unique lanthionine synthetases reveals new mechanistic and evolutionary insights, PLoS Biol 8, e1000339.

34. Holo, H., Jeknic, Z., Daeschel, M., Stevanovic, S., and Nes, I. F. (2001) Plantaricin W from Lactobacillus plantarum belongs to a new family of two-peptide lantibiotics, Microbiology 147, 643-651.

35. Majchrzykiewicz, J. A., Lubelski, J., Moll, G. N., Kuipers, A., Bijlsma, J. J., Kuipers, O. P., and Rink, R. (2010) Production of a class II two-component lantibiotic of Streptococcus pneumoniae using the class I nisin synthetic machinery and leader sequence, Antimicrob Agents Chemother 54, 1498-1505.

36. van der Meer, J. R., Polman, J., Beerthuyzen, M. M., Siezen, R. J., Kuipers, O. P., and de Vos, W. M. (1993) Characterization of the Lactococcus lactis nisin A operon genes nisP, encoding a subtilisin-like serine protease involved in precursor processing, and nisR, encoding a regulatory protein involved in nisin biosynthesis, J Bacteriol 175, 2578-2588.

37. Velásquez, J. E., Zhang, X., and van der Donk, W. A. (2011) Biosynthesis of the Antimicrobial Peptide Epilancin 15X and its Unusual N-terminal Lactate Moiety, Chem Biol 18, 857-867.

Page 176: © 2013 Neha Garg

166

38. Shi, Y., Yang, X., Garg, N., and van der Donk, W. A. (2011) Production of lantipeptides in Escherichia coli, J Am Chem Soc 133, 2338-2341.

39. Garg, N., Tang, W., Goto, Y., Nair, S. K., and van der Donk, W. A. (2012) Lantibiotics from Geobacillus thermodenitrificans, Proc Natl Acad Sci U S A 109, 5241-5246.

40. Hasper, H. E., de Kruijff, B., and Breukink, E. (2004) Assembly and stability of nisin-lipid II pores, Biochemistry 43, 11567-11575.

41. Ekkelenkamp, M. B., Hanssen, M., Danny Hsu, S. T., de Jong, A., Milatovic, D., Verhoef, J., and van Nuland, N. A. (2005) Isolation and structural characterization of epilancin 15X, a novel lantibiotic from a clinical strain of Staphylococcus epidermidis, FEBS Lett 579, 1917-1922.

Page 177: © 2013 Neha Garg

167  

Appendix: Structural characterization of lantibiotic dehydrogenase ElxO1

A.1 Introduction

The lantibiotic epilancin 15X is a member of the epilancin-group of peptides and is

produced by Staphylococcus epidermidis 15X154 (1-2). S. epidermidis 15X154, a clinical strain

isolated from a wound infection, is a colonizer of the human skin (3). Other staphylococcal

lantibiotics include Pep5 (4), epidermin (5), epicidin 280 (6), and epilancin K7 (7-8). Epilancin

15X was shown to have potent antimicrobial activity against pathogenic bacteria, including

MRSA and VRE (9). Epilancin 15X contains ten post-translationally modified amino acids, one

lanthionine, two methyllanthionine bridges, and an unusual N-terminal 2-hydroxypropionyl

group (D-lactate, Figure 1.1 and 1.3, Chapter 1). The gene cluster and biosynthetic pathway of

epilancin 15X was recently characterized (2). The N-terminal lactate was shown to be produced

by dehydration of the first serine residue in the core peptide by the lantibiotic dehydratase ElxB,

followed by proteolytic removal of the leader peptide by the lantibiotic protease ElxP. The

spontaneous hydrolysis of the N-terminal dehydroalanine to a pyruvate group and reduction of

the newly generated pyruvate to lactate by the nicotinamide adenine dinucleotide phosphate

(NADPH)-dependent oxidoreductase ElxO complete the biosynthesis of epilancin 15X. The N-

terminal lactate group was shown to impart stability to epilancin 15X against bacterial

aminopeptidases. Three lantibiotics namely, epilancin K7 (7-8), SWLP1 (10), and epicidin 280

(6) have also been reported to contain an N-terminal lactate group. The biosynthetic gene cluster

of epicidin 280 contains a homologous oxidoreductase EciO and was suggested to be involved in

the biosynthesis of its N-terminal lactate (6). The enzymatic activity of ElxO has been

reconstituted in vitro and ElxO was shown to have relaxed substrate specificity by Dr. Juan

Esteban Velásquez, as various N-terminal ketone-containing peptides were tolerated as

Page 178: © 2013 Neha Garg

168  

substrates. The N-terminal lactate group generated by ElxO can be engineered into other

lantibiotics in order to enhance their biological stability. In this study, the crystal structure of

ElxO was determined providing insights into the catalytic mechanism.

A.2 Experimental procedures

A.2.a Expression and purification of ElxO

The plasmid pET28/His6-ElxO encoding hexahistidine tagged ElxO was provided by

former lab member Dr. Juan Esteban Velásquez. The overexpression and purification was carried

out in Prof. Satish K. Nair’s laboratory. Chemically competent E. coli Rossetta2 cells were

transformed with pET28/His6-ElxO. An overnight culture grown from a single colony

transformant was used as inoculum for 2 L of LB medium containing 50 µg/L kanamycin and 25

µg/L chloramphenicol, and the culture was grown at 37 °C until the OD600 reached about 0.6.

The incubation temperature was then changed to 18 °C and the culture was induced with 0.1 mM

IPTG. The induced cells were shaken continually at 18 °C for an additional 12 h. The cells were

harvested by centrifugation (11,900 × g for 10 min, Beckman JLA-10.500 rotor). The cell paste

was resuspended in 20 mM Tris-HCl, 1 mM NaCl, and 10% glycerol, pH 8.0. The cells were

lysed using a C5 Homogenizer (Avestin Inc.). The lysed cells were pelleted by centrifugation at

23,700 x g for 25 min. The purification by nickel column chromatography was accomplished

using a 5 mL HiTrap chelating HP nickel affinity column (GE Healthcare) on an AKTA FPLC

system (Amersham Pharmacia Biosystems). The supernatant from the centrifugation step was

applied to the column using a 50 mL superloop. The protein was detected by absorbance at 280

nm. The column was washed with buffer A containing 20 mM Tris-HCl, pH 8.0, 1 M NaCl, and

30 mM imidazole to remove any non-specifically bound proteins until a flat baseline was

Page 179: © 2013 Neha Garg

169  

observed. A gradient of 0-100% B over 45 min was used to elute the bound proteins where

buffer B contained 1 M NaCl, 20 mM Tris-HCl, pH 8.0, and 200 mM imidazole. The most

concentrated fraction (4 mL) was loaded onto an S-200 gel filtration column (Amersham

Biosciences) equilibrated in 20 mM HEPES pH 7.5 containing 100 mM KCl.

A.2.b Protein crystallization, X-ray diffraction data collection and structure determination

Apo-ElxO was concentrated to a final concentration of 0.4 mM by centrifugal filters, and

incubated with five fold excess of freshly prepared NADPH. The solution was incubated on ice

for 2 h prior to crystallization. Crystals of the ElxO wild type enzyme in complex with NADPH

were grown using the hanging drop vapor diffusion technique. Commercially available sparse

matrix crystallization screens were used to identify initial crystallization conditions using the

Hampton Natrix crystallization screen. The initial condition (Natrix 39) was optimized to 0.1 M

ammonium acetate, 0.02 M magnesium chloride, 0.05 M HEPES-Na (pH 7.5), 8% PEG 8000,

and 5% glycerol. Crystals were obtained by adding 1 L of the protein to an equal volume of the

crystallization mother liquor equilibrated against the same solution at 9 oC. Crystals grew within

three days and were harvested after one week. Crystals were vitrified by soaking in

crystallization condition supplemented with 30% glycerol prior to direct immersion in liquid

nitrogen. All X-ray diffraction data was collected at the Life Sciences Collaborative Access

Team (Sector-21), Argonne National Laboratory, Lemont- IL.

The diffraction data was indexed and scaled using HKL2000 (11) to a limiting resolution

of 1.85 Å. The coordinates from three models (PDB ids: 2HQI, 2UVD and 2CFC) were

superimposed to generate a search model for molecular replacement. All water and ligand atoms

Page 180: © 2013 Neha Garg

170  

from the search model were deleted, and non-conserved amino acid side chain reduced to

alanines. A five fold redundant native dataset for the ElxO-NADPH complex crystals was used.

Following rigid body refinement of the initial MR solution, phases from the resultant

model were further improved by two-fold density averaging allowing for a significant portion of

the model to be manually rebuilt using XtalView (12). Further manual fitting was interspersed

with automated rebuilding using ARP/wARP (13) and rounds of refinement using REFMAC5

(14-15). Cross-validation, using 5% of the data for the calculation of the free R factor (16), was

utilized throughout the model building process in order to monitor building bias. Although clear

density for NADPH could be observed prior to crystallographic refinement, the ligand was

manually built into the model only after the free R factor dropped below 30%. The

stereochemistry of the models was routinely monitored throughout the course of refinement

using PROCHECK (17).

A.3 Results and Discussion

In order to investigate and understand substrate specificity and the catalytic mechanism

of ElxO, crystallographic structure determination was performed. The crystallographic

asymmetric unit consists of two protein molecules with extensive contacts between the

monomers, consistent with the existence of ElxO in solution as a dimer. The crystal structure of

ElxO reveals it to contain a single domain Rossmann fold (18), with high structural homology to

members of the short-chain dehydrogenases (19). A DALI server based structure homology

search identified the closest homologs to be hydroxypropyl-coenzyme M dehydrogenase (PDB

id: 2CFC, root mean square deviation (RMSD) = 1.4, 35% sequence identity) (20), 3-oxoacyl-

[acyl carrier protein] dehydrogenase FabG (PDB id: 4JRO, RMSD = 1.6, 36% sequence

Page 181: © 2013 Neha Garg

171  

identity), and a putative acetoacetyl coenzyme A reductase (PDB id: 4K6F, RMSD = 1.4, 31%

sequence identity). Hence members of the short chain dehydrogenase superfamily share a

conserved structural core despite a moderate degree of sequence conservation.

Central to the structure of ElxO is a seven membered parallel β-sheet. This β-sheet is

decorated by three α-helices on either side (Figure A.1). The ElxO-NADPH co-crystal structure

identifies the cofactor binding interactions. The cofactor is bound in a shallow surface cavity of

the protein in a fully extended conformation (Figure A.2). The nicotinamide dinucleotide is

bound at the top of the β-sheet, with the diphosphate moiety bound between the loops emanating

from the third and the fourth β-strands. The binding site for the diphosphate is provided by the

canonical Rossmann fold, comprising the conserved TGX3GXG (Thr9 – Gly16 for ElxO) motif.

This motif provides the primary stabilizing interactions for the diphosphate of the cofactor by

interactions with the main chain amide atoms. The adenine moiety of the cofactor is bound by

the hydrophobic pocket defined by Leu58, Val60, Ile70, Ala87, and Ile89 at one face of the ring,

with Arg34 on the other face of the ring. The nicotinamide moiety is similarly bound by the

hydrophobic pocket defined by the side chains of Ile14, Pro182, Ile223 and Val219. The primary

amine of the side chain of Lys156 forms hydrogen bonds to both the 2’ and 3’ ribose sugar

hydroxyls of the nicotinamide ribose. The adenine ribose 2’-phosphate of NADPH is coordinated

by salt bridge interactions to the basic side chains of Arg35 and Lys39. These interactions likely

provide the basis for cofactor specificity for ElxO.

Though the majority of the residues can be traced in the electron density map, residues

188 to 204 are not visible. This region has been postulated to be responsible for binding to the

substrate. Hence disorder in this region can be attributed to the lack of substrate in the structure.

This region is expected to fold into a helix structure forming a protective lid over the active site

Page 182: © 2013 Neha Garg

172  

that may undergo an open to closed conformational change upon substrate binding (21) (Figure

A.2). By homology to other members of the short chain dehydrogenase family, several key

catalytic residues can be readily identified from the structure of ElxO in complex with NADPH,

despite the absence of the substrate peptide. Primary among these are Ser139, Tyr152 and

Lys156. Based on the mechanism of known short-chain dehydrogenases (19) and the active site

architecture of ElxO, a reaction mechanism is proposed (Figure A.3). The protonated hydroxyl

group of Tyr152 and Ser139 hydrogen bond with the N-terminal carbonyl oxygen of

dehydroepilancin 15X promoting transfer of hydride from the reduced nicotinamide ring to the

carbonyl carbon of dehydroepilancin 15X. The resulting positive charge on the oxidized

nicotinamide ring and the adjacent ε-amino group of Lys156 may lower the pKa of tyr152

allowing the transfer of a proton to the newly formed alkoxide in the substrate, thereby

generating the lactate moiety at the N-terminus. Extensive co-crystallization trials, including

screening for crystal growth in the presence of substrate peptides, and soaking of ElxO-NADPH

crystals were tried with four short peptides pyruvate (Pyr)-DAIVK, Pyr-VAIVK, 2-oxobutyryl

(OBut)-AAIVK, and OBut-RAIVK without success. This may be due to high Km values of these

substrate peptides (unpublished data from previous lab member Dr. Juan Esteban Velásquez).

Page 183: © 2013 Neha Garg

173  

Figure A.1. Cartoon representation of the three dimensional crystal structure of the ElxO monomer in complex with NADPH. Helices are colored blue, and the beta-sheets are colored red. NADPH is shown in stick representation with the carbon atoms colored green.

Page 184: © 2013 Neha Garg

174  

Figure A.2. Surface representation of ElxO with NADPH shown in stick representation. The ElxO surface is colored brown, with the carbon atoms of NADPH shown in yellow. Note the postulated substrate binding empty cavity positioned above the nicotinamide ring of the cofactor.

Page 185: © 2013 Neha Garg

175  

Figure A.3. The proposed reaction mechanism for ElxO and the postulated contributions of individual side chains based on homology to other members of the short chain dehydrogenase family of enzymes.

Page 186: © 2013 Neha Garg

176  

A.4 References

1. Ekkelenkamp, M. B., Hanssen, M., Danny Hsu, S. T., de Jong, A., Milatovic, D., Verhoef, J., and van Nuland, N. A. (2005) Isolation and structural characterization of epilancin 15X, a novel lantibiotic from a clinical strain of Staphylococcus epidermidis, FEBS Lett 579, 1917-1922.

2. Velásquez, J. E., Zhang, X., and van der Donk, W. A. (2011) Biosynthesis of the antimicrobial peptide epilancin 15X and its N-terminal lactate, Chem Biol 18, 857-867.

3. Archer, G. L. (2000) Staphylococcus epidermidis and other coagulase-negative staphylococci in: Principles and Practice of Infectious Diseases (Mandell, G.L., Bennett, J.E. and Dolin, R., Eds.), pp. 2092–2100, Churchill Livingstone, Philadelphia, PA.

4. Sahl, H. G., Grossgarten, M., Widger, W. R., Cramer, W. A., and Brandis, H. (1985) Structural similarities of the staphylococcin-like peptide Pep-5 to the peptide antibiotic nisin, Antimicrob Agents Chemother 27, 836-840.

5. Allgaier, H., Jung, G., Werner, R. G., Schneider, U., and Zahner, H. (1986) Epidermin: sequencing of a heterodetic tetracyclic 21-peptide amide antibiotic, Eur J Biochem 160, 9-22.

6. Heidrich, C., Pag, U., Josten, M., Metzger, J., Jack, R. W., Bierbaum, G., Jung, G., and Sahl, H. G. (1998) Isolation, characterization, and heterologous expression of the novel lantibiotic epicidin 280 and analysis of its biosynthetic gene cluster, Appl Environ Microbiol 64, 3140-3146.

7. van de Kamp, M., Horstink, L. M., van den Hooven, H. W., Konings, R. N., Hilbers, C. W., Frey, A., Sahl, H. G., Metzger, J. W., and van de Ven, F. J. (1995) Sequence analysis by NMR spectroscopy of the peptide lantibiotic epilancin K7 from Staphylococcus epidermidis K7, Eur J Biochem 227, 757-771.

8. van de Kamp, M., van den Hooven, H. W., Konings, R. N. H., Bierbaum, G., Sahl, H.-G., Kuipers, O. P., Siezen, R. J., de Vos, W., Hilbers, C. W., and van de Ven, F. J. (1995) Elucidation of the primary structure of the lantibiotic epilancin K7 from Staphylococcus epidermis K7. Cloning and characterization of the epilancin-K7-encoding gene and NMR analysis of mature epilancin K7, Eur. J. Biochem. 230, 587-600.

9. Verhoef, J., Milatovic, D., and Ekkelenkamp, M.B. (March 17, 2005.) Antimicrobial compounds. Patent WO/2005/023852 (http://www.wipo.int/patentscope/search/en/WO2005023852).

10. Petersen, J., Boysen, A., Fogh, L., Tabermann, K., Kofoed, T., King, A., Schrotz-King, P., and Hansen, M. C. (2009) Identification and characterization of a bioactive lantibiotic produced by Staphylococcus warneri, Biol Chem 390, 437-444.

11. Otwinowski, Z., Borek, D., Majewski, W., and Minor, W. (2003) Multiparametric scaling of diffraction intensities, Acta Crystallogr A 59, 228-234.

Page 187: © 2013 Neha Garg

177  

12. McRee, D. E. (1999) XtalView/Xfit--A versatile program for manipulating atomic coordinates and electron density, J Struct Biol 125, 156-165.

13. Perrakis, A., Sixma, T. K., Wilson, K. S., and Lamzin, V. S. (1997) wARP: improvement and extension of crystallographic phases by weighted averaging of multiple-refined dummy atomic models, Acta Crystallogr D Biol Crystallogr 53, 448-455.

14. Murshudov, G. N., Vagin, A. A., and Dodson, E. J. (1997) Refinement of macromolecular structures by the maximum-likelihood method, Acta Crystallogr D Biol Crystallogr 53, 240-255.

15. Murshudov, G. N., Vagin, A. A., Lebedev, A., Wilson, K. S., and Dodson, E. J. (1999) Efficient anisotropic refinement of macromolecular structures using FFT, Acta Crystallogr D Biol Crystallogr 55, 247-255.

16. Kleywegt, G. J., and Brunger, A. T. (1996) Checking your imagination: applications of the free R value, Structure 4, 897-904.

17. Laskowski, R. A., Rullmannn, J. A., MacArthur, M. W., Kaptein, R., and Thornton, J. M. (1996) AQUA and PROCHECK-NMR: programs for checking the quality of protein structures solved by NMR, J Biomol NMR 8, 477-486.

18. Rao, S. T., and Rossmann, M. G. (1973) Comparison of super-secondary structures in proteins, J Mol Biol 76, 241-256.

19. Ladenstein, R., Winberg, J. O., and Benach, J. (2008) Medium- and short-chain dehydrogenase/reductase gene and protein families : Structure-function relationships in short-chain alcohol dehydrogenases, Cell Mol Life Sci 65, 3918-3935.

20. Krishnakumar, A. M., Nocek, B. P., Clark, D. D., Ensign, S. A., and Peters, J. W. (2006) Structural basis for stereoselectivity in the (R)- and (S)-hydroxypropylthioethanesulfonate dehydrogenases, Biochemistry 45, 8831-8840.

21. Tanaka, N., Nonaka, T., Nakamura, K.T., and Hara, A. (2001) SDR: Structure, mechanism of action, and substrate recognition, Curr Org Chem 5, 89-111.