Download Novel Antimicrobial Agents and Strategies

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts
no text concepts found
Transcript
Edited by
David A. Phoenix,
Frederick Harris, and
Sarah R. Dennison
Novel Antimicrobial Agents
and Strategies
Related Titles
Gualerzi, C.O., Brandi, L., Fabbretti, A.,
Pon, C.L. (eds.)
Antibiotics
Skold, O.
Antibiotics and Antibiotic
Resistance
Targets, Mechanisms and Resistance
2011
2013
Print ISBN: 978-3-527-33305-9, also
available in digital formats
Print ISBN: 978-0-470-43850-3, also
available in digital formats
Chen, L., Petersen, J., Schlagenhauf, P. (eds.)
Phoenix, D. A., Dennison, S. R., Harris, F.
Antimicrobial Peptides
Infectious Diseases - A
Geographic Guide
2013
2011
Print ISBN: 978-3-527-33263-2, also
available in digital formats
Anderson, R.R., Groundwater, P.P.,
Todd, A.A., Worsley, A.A.
Antibacterial Agents Chemistry, Mode of Action,
Mechanisms of Resistance and
Clinical Applications
2012
Print ISBN: 978-0-470-97244-1, also
available in digital formats
Print ISBN: 978-0-470-65529-0, also
available in digital formats
De Clercq, E. (ed.)
Antiviral Drug Strategies
2011
Print ISBN: 978-3-527-32696-9, also
available in digital formats
Edited by
David A. Phoenix, Frederick Harris,
and Sarah R. Dennison
Novel Antimicrobial Agents and Strategies
The Editors
Prof. David A. Phoenix
London South Bank University
Borough Road 103
London
SE1 0AA
United Kingdom
All books published by Wiley-VCH are
carefully produced. Nevertheless, authors,
editors, and publisher do not warrant the
information contained in these books,
including this book, to be free of errors.
Readers are advised to keep in mind that
statements, data, illustrations, procedural
details or other items may inadvertently
be inaccurate.
Dr. Frederick Harris
University of Central Lancashire
Forensic & Investigative Science
Preston, Lancashire
PR1 2HE
United Kingdom
Library of Congress Card No.: applied for
British Library Cataloguing-in-Publication
Data
A catalogue record for this book is
available from the British Library.
Dr. Sarah R. Dennison
University of Central Lancashire
Pharmacy and Biomedical Science
Preston, Lancashire
PR1 2HE
United Kingdom
Bibliographic information published by the
Deutsche Nationalbibliothek
The Deutsche Nationalbibliothek
lists this publication in the Deutsche
Nationalbibliografie; detailed
bibliographic data are available on the
Internet at <http://dnb.d-nb.de>.
Cover design
The cover shows beta-lactamase, an
enzyme produced by some bacteria,
which provide bacterial resistance to
beta-lactam antibiotics in the presence of
a lipid bilayer. The image was created by
Dr. Manuela Mura, University of Central
Lancashire, UK.
© 2015 Wiley-VCH Verlag Gmbh & Co.
KGaA, Boschstr. 12, 69469 Weinheim,
Germany
All rights reserved (including those of
translation into other languages). No part
of this book may be reproduced in any
form – by photoprinting, microfilm,
or any other means – nor transmitted
or translated into a machine language
without written permission from the
publishers. Registered names, trademarks,
etc. used in this book, even when not
specifically marked as such, are not to be
considered unprotected by law.
Print ISBN: 978-3-527-33638-8
ePDF ISBN: 978-3-527-67614-9
ePub ISBN: 978-3-527-67615-6
Mobi ISBN: 978-3-527-67616-3
oBook ISBN: 978-3-527-67613-2
Cover-Design Adam-Design, Weinheim,
Germany
Typesetting Laserwords Private Limited,
Chennai, India
Printing and Binding Markono Print
Media Pte Ltd, Singapore
Printed on acid-free paper
V
Contents
List of Contributors XI
Preface XVII
1
1
The Problem of Microbial Drug Resistance
Iza Radecka, Claire Martin, and David Hill
1.1
1.2
Introduction 1
History of the Origins, Development, and Use of Conventional
Antibiotics 1
Problems of Antibiotic Resistance 4
Multiple Drug-Resistant (MDR), Extensively Drug-Resistant (XDR),
and Pan-Drug-Resistant (PDR) Organisms 5
MDR Mechanisms of Major Pathogens 5
Antimicrobial Stewardship Programs 11
Discussion 12
Acknowledgment 13
References 13
1.3
1.4
1.5
1.6
1.7
2
Conventional Antibiotics – Revitalized by New Agents 17
Anthony Coates and Yanmin Hu
2.1
2.2
2.3
2.4
Introduction 17
Conventional Antibiotics 18
The Principles of Combination Antibiotic Therapy 20
Antibiotic Resistance Breakers: Revitalize Conventional
Antibiotics 21
β-Lactamase Inhibitors 21
Aminoglycoside-Modifying Enzyme Inhibitors 23
Antibiotic Efflux Pumps Inhibitors 23
Synergy Associated with Bacterial Membrane Permeators 23
Discussion 25
Acknowledgments 26
References 26
2.4.1
2.4.2
2.4.3
2.4.4
2.5
VI
Contents
3
Developing Novel Bacterial Targets: Carbonic Anhydrases as
Antibacterial Drug Targets 31
Clemente Capasso and Claudiu T. Supuran
3.1
3.2
3.3
3.4
3.5
3.6
3.7
3.8
3.9
Introduction 31
Carbonic Anhydrases 31
CA Inhibitors 32
Classes of CAs Present in Bacteria 33
Pathogenic Bacterial CAs 35
α-CAs in Pathogenic Bacteria 35
β-CAs in Pathogenic Bacteria 37
γ-CAs from Pathogenic Bacteria 39
Conclusions 40
References 41
4
Magainins – A Model for Development of Eukaryotic Antimicrobial
Peptides (AMPs) 47
Sarah R. Dennison, Frederick Harris, and David A. Phoenix
4.1
4.2
4.3
4.4
4.5
Introduction 47
Magainins and Their Antimicrobial Action 49
Magainins as Antibiotics 51
Other Antimicrobial Uses of Magainins 55
Future Prospects for Magainins 57
References 58
5
Antimicrobial Peptides from Prokaryotes 71
Maryam Hassan, Morten Kjos, Ingolf F. Nes, Dzung B. Diep,
and Farzaneh Lotfipour
5.1
5.2
5.2.1
5.2.2
Introduction 71
Bacteriocins 73
Microcins – Peptide Bacteriocins from Gram-Negative Bacteria
Lanthibiotics – Post-translationally Modified Peptides from
Gram-Positive Bacteria 76
Non-modified Peptides from Gram-Positive Bacteria 77
Applications of Prokaryotic AMPs 79
Food Biopreservation 79
Bacteriocinogenic Probiotics 80
Clinical Application 81
Applications in Dental Care 82
Development and Discovery of Novel AMP 82
References 84
5.2.3
5.3
5.3.1
5.3.2
5.3.3
5.3.4
5.4
6
Peptidomimetics as Antimicrobial Agents
Peng Teng, Haifan Wu, and Jianfeng Cai
6.1
6.2
Introduction 91
Antimicrobial Peptidomimetics
93
91
73
Contents
6.2.1
6.2.2
6.2.3
6.2.4
6.2.5
6.2.6
6.2.6.1
6.2.6.2
6.3
Peptoids 93
β-Peptides 94
Arylamides 96
β-Peptoid–Peptide Hybrid Oligomers 97
Oligourea and γ4 -Peptide-Based Oligomers 98
AApeptides 98
α-AApeptides 99
γ-AApeptides 101
Discussion 102
Acknowledgments 103
References 103
7
Synthetic Biology and Therapies for Infectious Diseases 109
Hiroki Ando, Robert Citorik, Sara Cleto, Sebastien Lemire, Mark Mimee,
and Timothy Lu
7.1
7.2
7.3
7.3.1
7.3.2
7.4
7.4.1
7.4.2
7.4.3
7.4.4
7.5
7.5.1
7.5.2
7.5.3
7.6
7.7
7.7.1
7.7.2
7.7.2.1
7.7.2.2
7.7.2.3
7.7.2.4
7.7.2.5
Current Challenges in the Treatment of Infectious Diseases 109
Introduction to Synthetic Biology 112
Vaccinology 113
Genetic Engineering and Vaccine Development 114
Rational Antigen Design Through Reverse Vaccinology 119
Bacteriophages: A Re-emerging Solution? 122
A Brief History of Bacteriophages 122
Addressing the Problem of the Restricted Host Range of Phages 124
Phage Genome Engineering for Enhanced Therapeutics 129
Phages as Delivery Agents for Antibacterial Cargos 132
Isolated Phage Parts as Antimicrobials 133
Engineered Phage Lysins 133
Pyocins: Deadly Phage Tails 135
Untapped Reservoirs of Antibacterial Activity 136
Predatory Bacteria and Probiotic Bacterial Therapy 136
Natural Products Discovery and Engineering 139
In Silico and In Vitro Genome Mining for Natural Products 140
Strain Engineering for Natural Products 144
Production of the Antimalarial Artemisinin 145
Daptomycin (Cubicin) 147
Echinomycin 147
Clavulanic Acid 148
Production of the Antiparasitic Avermectin and Its Analogs
Doramectin and Ivermectin 149
Production of Doxorubicin/Daunorubicin 149
Development of Hosts for the Expression of Nonribosomal Peptides
and Polyketides 150
Generation of Novel Molecules by Rational Reprogramming 152
Engineering NRPS and PKS Domains 154
Activation of Cryptic Genes/Clusters 155
7.7.2.6
7.7.2.7
7.7.3
7.7.4
7.7.5
VII
VIII
Contents
7.7.6
7.8
Mutasynthesis as a Source of Novel Analogs 157
Summary 157
Acknowledgments 157
References 158
8
Nano-Antimicrobials Based on Metals 181
Maria Chiara Sportelli, Rosaria Anna Picca, and Nicola Cioffi
8.1
8.2
8.2.1
8.2.1.1
8.2.1.2
8.2.1.3
8.2.1.4
8.2.1.5
8.2.2
8.2.3
8.2.3.1
8.3
8.3.1
Introduction 181
Silver Nano-antimicrobials 182
Synthesis of Silver Nanostructures 182
Physical Approaches 183
Laser Ablation in Liquids 183
Chemical Approaches 183
Biological and Biotechnological Approaches 184
Electrochemical Approaches 184
Characterization of Silver Nanostructures 185
Applications of Silver Nanostructures 187
Silver-Based Nano-antimicrobials 187
Copper Nano-antimicrobials 190
Preparation and Applications of Antimicrobial Cu
Nanostructures 190
Physical Methods 190
Wet-Chemical Methods 192
Electrochemical Syntheses 195
Laser Ablation in Liquids 196
Biological Syntheses 197
Zinc Oxide Nano-antimicrobials 197
Synthesis of Zinc Oxide Nanostructures 197
Physical Approaches 198
Chemical Approaches 198
Electrochemical Approaches 200
Conclusions 201
References 201
8.3.1.1
8.3.1.2
8.3.1.3
8.3.1.4
8.3.1.5
8.4
8.4.1
8.4.1.1
8.4.1.2
8.4.1.3
8.5
9
Natural Products as Antimicrobial Agents – an Update 219
Muhammad Saleem
9.1
9.2
9.2.1
9.2.2
9.2.3
9.3
9.4
9.5
9.6
Introduction 219
Antimicrobial Natural Products from Plants 220
Antimicrobial Alkaloids from Plants 220
Antimicrobial Alkaloids from Microbial Sources 223
Antimicrobial Alkaloids from Marine Sources 225
Antimicrobial Natural Products Bearing an Acetylene Function
Antimicrobial Carbohydrates 228
Antimicrobial Natural Chromenes 228
Antimicrobial Natural Coumarins 229
226
Contents
9.6.1
9.6.1.1
9.7
9.7.1
9.8
9.8.1
9.9
9.9.1
9.10
9.10.1
9.10.2
9.10.3
9.11
9.12
9.12.1
9.12.2
9.12.2.1
9.12.2.2
9.12.2.3
9.12.3
9.13
9.13.1
9.13.2
9.14
9.14.1
9.14.2
9.14.3
9.15
9.15.1
9.15.2
9.15.3
9.16
Antimicrobial Coumarins from Plants 229
Antimicrobial Coumarins from Bacteria 232
Antimicrobial Flavonoids 232
Antimicrobial Flavonoids from Plants 233
Antimicrobial Iridoids 237
Antimicrobial Iridoids from Plants 237
Antimicrobial Lignans 238
Antimicrobial Lignans from Plants 238
Antimicrobial Phenolics Other Than Flavonoids and Lignans 240
Antimicrobial Phenolics from Plants 240
Antimicrobial Phenolics from Microbial Sources 244
Antimicrobial Phenolics from Marine Source 246
Antimicrobial Polypeptides 247
Antimicrobial Polyketides 249
Antimicrobial Polyketides as Macrolides 250
Antimicrobial Polyketides as Quinones and Xanthones 252
Antimicrobial Quinones and Xanthones from Plants 252
Antimicrobial Quinones from Bacteria 256
Antimicrobial Quinones and Xanthones from Fungi 257
Antimicrobial Fatty Acids and Other polyketides 261
Antimicrobial Steroids 263
Antimicrobial Steroids from Plants 264
Steroids from Fungi 266
Antimicrobial Terpenoids 267
Antimicrobial Terpenoids from Plants 267
Antimicrobial Terpenoids from Microbial Sources 273
Antimicrobial Terpenoids from Marine Sources 274
Miscellaneous Antimicrobial Compounds 275
Miscellaneous Antimicrobial Natural Products from Plants 275
Miscellaneous Antimicrobials from Bacteria 278
Miscellaneous Antimicrobials from Fungi 280
Platensimycin Family as Antibacterial Natural Products 282
References 284
10
Photodynamic Antimicrobial Chemotherapy 295
David A. Phoenix, Sarah R. Dennison, and Frederick Harris
10.1
10.2
10.3
10.4
10.4.1
10.4.2
10.5
Introduction 295
The Administration and Photoactivation of PS 296
Applications of PACT Based on MB 301
The Applications of PACT Based on ALA 303
Food Decontamination Using PACT Based on ALA 303
Dermatology Using PACT Based on ALA 305
Future Prospects 308
References 310
IX
X
Contents
11
The Antimicrobial Effects of Ultrasound 331
Frederick Harris, Sarah R. Dennison, and David A. Phoenix
11.1
11.2
11.3
11.3.1
11.3.2
11.4
Introduction 331
The Antimicrobial Activity of Ultrasound Alone 332
The Antimicrobial Activity of Assisted Ultrasound 335
Synergistic Effects 336
Sonosensitizers 338
Future Prospects 341
References 343
12
Antimicrobial Therapy Based on Antisense Agents 357
Glenda M. Beaman, Sarah R. Dennison, and David A. Phoenix
12.1
12.2
12.3
12.4
12.5
12.6
12.7
12.8
12.9
12.10
12.10.1
12.10.2
12.11
Introduction 357
Antisense Oligonucleotides 358
First-Generation ASOs 360
Second-Generation ASOs 361
Third-Generation ASOs 362
Antisense Antibacterials 364
Broad-Spectrum Antisense Antibacterials 365
Methicillin-Resistant Staphylococcus aureus (MRSA) 371
RNA Interference (RNAi) 371
Progress Using siRNA 374
Mycobacterium Tuberculosis 374
MRSA 375
Discussion 376
References 377
13
New Delivery Systems – Liposomes for Pulmonary Delivery of
Antibacterial Drugs 387
Abdelbary M.A. Elhissi, Sarah R. Dennison, Waqar Ahmed, Kevin M.G. Taylor
and David A. Phoenix
13.1
13.2
13.3
13.3.1
13.3.1.1
13.3.1.2
13.3.1.3
13.4
Introduction 387
Pulmonary Drug Delivery 389
Liposomes as Drug Carriers in Pulmonary Delivery 389
Liposomes for Pulmonary Delivery of Antibacterial Drugs 390
Delivery of Antibacterial Liposomes Using pMDIs 391
Delivery of Antibacterial Liposomes Using DPIs 392
Delivery of Antibacterial Liposomes Using Nebulizers 394
Present and Future Trends of Liposome Research in Pulmonary Drug
Delivery 398
Conclusions 401
References 401
13.5
Index 407
XI
List of Contributors
Waqar Ahmed
Glenda M. Beaman
University of Central Lancashire
Institute of Nanotechnology and
Bioengineering
School of Medicine and
Dentistry
Corporation street
Preston
PR1 2HE
UK
University of Central Lancashire
School of Forensic and
Investigative Sciences
Corporation Street
Preston
PR1 2HE
UK
Hiroki Ando
Department of Electrical
Engineering and Computer
Science and Department of
Biological Engineering
Massachusetts Institute of
Technology
77 Massachusetts Avenue
Cambridge, MA 02139
USA
Jianfeng Cai
University of South Florida
Department of Chemistry
4202 E. Fowler Avenue
Tampa, FL 33620
USA
Clemente Capasso
Istituto di Biochimica delle
Proteine-CNR
via Pietro Castellino
111 - 80131 Napoli
Italy
and
and
Massachusetts Institute of
Technology
MIT Synthetic Biology Center
500 Technology Square
Cambridge, MA 02139
USA
Istituto di Bioscienze e
Biorisorse-CNR
via Pietro Castellino
111 - 80131 Napoli
Italy
XII
List of Contributors
Nicola Cioffi
Sara Cleto
Università degli Studi di Bari
Aldo Moro
Dipartimento di Chimica
via Orabona 4
70126 Bari
Italy
Department of Electrical
Engineering and Computer
Science and Department of
Biological Engineering
Massachusetts Institute of
Technology
77 Massachusetts Avenue
Cambridge, MA 02139
USA
Robert Citorik
Department of Electrical
Engineering and Computer
Science and Department of
Biological Engineering
Massachusetts Institute of
Technology
77 Massachusetts Avenue
Cambridge, MA 02139
USA
and
Massachusetts Institute of
Technology
MIT Synthetic Biology Center
500 Technology Square
Cambridge, MA 02139
USA
and
Anthony Coates
Massachusetts Institute of
Technology
MIT Synthetic Biology Center
500 Technology Square
Cambridge, MA 02139
USA
and
Massachusetts Institute of
Technology
MIT Microbiology Program
77 Massachusetts Avenue
Cambridge, MA 02139
USA
St George’s University of London
Medical Microbiology
Institute of Infection and
Immunity
Cranmer Terrace
London
SW17 0RE
UK
Sarah R. Dennison
University of Central Lancashire
Institute of Nanotechnology and
Bioengineering
School of Pharmacy and
Biomedical Sciences
Corporation Street
Preston
PR1 2HE
UK
List of Contributors
Dzung B. Diep
David Hill
Norwegian University of Life
Sciences
Laboratory of Microbial Gene
Technology
Department of Chemistry
Biotechnology and Food Science
P.O. Box 5003
1432 Ås
Norway
University of Wolverhampton
School of Biology, Chemistry,
and Forensic Science
Faculty of Science and
Engineering
Wulfruna Street
Wolverhampton
WV1 1LY
UK
Abdelbary M.A. Elhissi
Yanmin Hu
University of Central Lancashire
Institute of Nanotechnology and
Bioengineering
School of Pharmacy and
Biomedical Sciences
Corporation street
Preston
PR1 2HE
UK
St George’s University of London
Medical Microbiology
Institute of Infection and
Immunity
Cranmer Terrace
London
SW17 0RE
UK
Morten Kjos
Frederick Harris
University of Central Lancashire
School of Forensic and
Investigative Science
Corporation street
Preston
PR1 2HE
UK
Norwegian University of Life
Sciences
Laboratory of Microbial Gene
Technology
Department of Chemistry
Biotechnology and Food Science
P.O. Box 5003
1432 Ås
Norway
Maryam Hassan
Zanjan University of Medical
Sciences
Pharmaceutical Biotechnology
Research Center
Zanjan
Iran
and
University of Groningen
Molecular Genetics Group
Groningen Biomolecular
Sciences and Biotechnology
Institute
Centre for Synthetic Biology
Nijenborgh 7
9747 AG Groningen
The Netherlands
XIII
XIV
List of Contributors
Sebastien Lemire
Timothy Lu
Department of Electrical
Engineering and Computer
Science and Department of
Biological Engineering
Massachusetts Institute of
Technology
77 Massachusetts Avenue
Cambridge, MA 02139
USA
Department of Electrical
Engineering and Computer
Science and Department of
Biological Engineering
Massachusetts Institute of
Technology
77 Massachusetts Avenue
Cambridge, MA 02139
USA
and
and
Massachusetts Institute of
Technology
MIT Synthetic Biology Center
500 Technology Square
Cambridge, MA 02139
USA
Massachusetts Institute of
Technology
MIT Synthetic Biology Center
500 Technology Square
Cambridge, MA 02139
USA
Farzaneh Lotfipour
and
Tabriz University of Medical
Sciences
Hematology & Oncology
Research Center and Faculty of
Pharmacy
Tabriz
51664
Iran
Massachusetts Institute of
Technology
MIT Microbiology Program
77 Massachusetts Avenue
Cambridge, MA 02139
USA
Claire Martin
University of Wolverhampton
School of Pharmacy
Faculty of Science and
Engineering
Wulfruna Street
Wolverhampton
WV1 1LY
UK
List of Contributors
Mark Mimee
David A. Phoenix
Department of Electrical
Engineering and Computer
Science and Department of
Biological Engineering
Massachusetts Institute of
Technology
77 Massachusetts Avenue
Cambridge, MA 02139
USA
London South Bank University
Office of the Vice Chancellor
103 Borough Road
London
SE1 0AA
UK
and
Massachusetts Institute of
Technology
MIT Synthetic Biology Center
500 Technology Square
Cambridge, MA 02139
USA
and
Massachusetts Institute of
Technology
MIT Microbiology Program
77 Massachusetts Avenue
Cambridge, MA 02139
USA
Ingolf F. Nes
Norwegian University of Life
Sciences
Laboratory of Microbial Gene
Technology
Department of Chemistry
Biotechnology and Food Science
P.O. Box 5003
1432 Ås
Norway
Rosaria Anna Picca
Università degli Studi di Bari
Aldo Moro
Dipartimento di Chimica
via Orabona 4
70126 Bari
Italy
Iza Radecka
University of Wolverhampton
School of Biology
Chemistry and Forensic Science
Faculty of Science and
Engineering
Wulfruna Street
Wolverhampton
WV1 1LY
UK
Muhammad Saleem
The Islamia University of
Bahawalpur
Department of Chemistry
Baghdad-ul-Jadeed Campus
Bahawalpur, 63100
Pakistan
XV
XVI
List of Contributors
Maria Chiara Sportelli
Kevin M.G. Taylor
Università degli Studi di Bari
Aldo Moro
Dipartimento di Chimica
via Orabona 4
70126 Bari
Italy
University College London
Department of Pharmaceutics
School of Pharmacy
29-39 Brunswick Square
London
WC1N 1AX
UK
Claudiu T. Supuran
Università degli Studi di Firenze
Dipartimento di Scienze
Farmaceutiche
Via della Lastruccia
3, Polo Scientifico
50019 Sesto Fiorentino
(Florence)
Italy
and
and
Peng Teng
Sezione di Scienze
Farmaceutiche e Nutraceutiche,
Neurofarba Department
Università degli Studi di Firenze
Via Ugo Schiff 6
50019 Sesto Fiorentino
(Florence)
Italy
Department of Pharmaceutics
UCL School of Pharmacy
29-39 Brunswick Square
London
WC1N 1AX
UK
University of South Florida
Department of Chemistry
4202 E. Fowler Avenue
Tampa, FL 33620
USA
Haifan Wu
University of South Florida
Department of Chemistry
4202 E. Fowler Avenue
Tampa, FL 33620
USA
XVII
Preface
The “Golden age of antibiotics” was between 1929 and the 1970s when over 20
antibiotic classes were marketed [1, 2]. Since the 1960s, the rise in the emergence
of microbial pathogens with multiple drug resistance (MDR) has led to the
realization that the “Golden age” had ended. The pharmaceutical industry has
been constantly battling with MDR because of the overprescription and misuse
of antibiotics [3–5]. In Chapter 1, Radecka and coworkers give an insight into
bacterial resistance being a major threat to public health. They also discuss
the implications arising from the threat posed by MDR pathogens in relation
to factors such as medical practice and economics, along with an overview
of recent practices and measures proposed to contain this threat, such as the
introduction of stewardship programs. Concern regarding our future ability to
combat infection has been further intensified by the decreasing supply of new
agents [3, 6–8], and in the remainder of the book we review approaches being
taken to identity and develop the antimicrobials of the future.
In response to the challenges outlined, in this book there has been increasing research into maximizing opportunities to develop and revitalize established
classes of antibiotics. Coates and Hu consider this area in Chapter 2 where they
look at opportunities to extend the life of old antibiotics such as β-lactams by the
addition of agents that can overcome drug resistance factors, such as β-lactamase
inhibitors. Identification of new, effective derivatives remains a challenge, and
another approach in the search for new antibiotics has been to seek out new targets that would enable new classes of antibacterials to be developed. In Chapter 3,
Capasso and Supuran review the cloning and characterization of carbonic anhydrases (CAs). In this chapter, they make reference to the impact of inhibitors that
target the α-, β-, and γ-CAs from many pathogenic bacteria and suggest that this
provides evidence that these proteins could provide novel antibacterial targets for
the development of new antimicrobial compounds.
There remain concerns, though, that only a small number of drugs are currently
under research and development as antibacterial agents [9]. It has been suggested
that a further approach could be to revisit naturally occurring compounds with
antibacterial potential. Due to the arrival of antibiotics, there has been a rapid
loss of interest in the therapeutic potential of natural host antibiotics such as
XVIII
Preface
lysozyme [3, 4]. However, more recently, there has been an awakened interest in
host defense molecules, such as antimicrobial peptides (AMPs) [10, 11]. Since
the early 1990s, the potential of AMPs has been investigated using, for example,
magainins isolated from the African clawed frog Xenopus laevis, to investigate
the effect of the structural and physiochemical properties of these peptides on
their antimicrobial action. These AMPs have the potency to target and kill a
wide range of Gram-negative and Gram-positive bacteria, fungi, viruses, and
some tumor cells [12]. Based on this ability, AMPs are attractive propositions for
development as therapeutically useful antimicrobial and anticancer agents [13].
The first clinical trials of these AMPs as potential novel antibiotics have been for
topical treatments [14], and Dennison et al. review this area in Chapter 4. AMPs
are not only produced by eukaryotes but are also generated by prokaryotes,
and Lotfipour and coworkers review this class of peptides, generally known as
bacteriocins, in Chapter 5. These prokaryotic peptides are produced by geneencoded or ribosome-independent pathways [15]. Non-ribosomal prokaryotic
AMPs generally include examples such as vancomycin and daptomycin, which
are assembled by large multifunctional enzyme complexes. Gene-encoded AMPs
from prokaryotes include microcins from Gram-negative bacteria, lantibiotics,
and nonmodified bacteriocins from Gram-positive bacteria. The potential uses
of these molecules are reviewed for their potential in food biopreservation and
healthcare. However, both eukaryotic and prokaryotic AMPs have a range of
challenges to overcome, such as the cost of production and design complexity of
these molecules. For this reason, work has been under way to design mimics and
peptidomimetics of these peptides, which is reviewed in Chapter 6 by Cai and
coworkers. Major examples of these molecules include : peptoids [16], β-peptides
[17], arylamide oligomers [18], AApeptides [19, 20], and other compounds
[21–25], which may be considered second-generation AMPs. These molecules
are designed to possess properties conducive to therapeutic application and
retain key structural characteristics of naturally occurring AMPs, such as positive
charge, hydrophobicity, and amphiphilicity, which facilitate their membranolytic
and antimicrobial activity. Tuning these properties has led to superior levels of
microbial selectivity and antimicrobial activity as compared to both natural AMPs
and conventional antibiotics. This Chapter considers the recent development of
these synthetic mimics of AMPs based on a variety of peptide backbones other
than canonical peptides, including β-peptides, peptoids, and AApeptides.
It is interesting to note that, in addition to direct action, AMPs are part of more
complex innate immune systems and a further approach to developing treatments
for the future has involved review of how aspects of such immune systems could be
adapted to support treatment of infections. Prior to the discovery and widespread
use of antibiotics, it was believed that bacterial infections could be treated by the
administration of bacteriophages, which are viruses that infect and kill bacteria via
lytic mechanisms but have no effect on humans. With the advent of penicillins and
other antibiotics, clinical studies with bacteriophages were not vigorously pursued
in the United States and Western Europe, but phage therapy was extensively used
in Eastern European countries mainly in the former Soviet Union and Georgia.
Preface
However, with the current rise of antibiotic-resistant bacteria, there has been a
revitalization of interest in phage therapy in Western countries. In Chapter 7, Lu
and coworkers discuss the use of synthetic biology and whether bacteriophages are
a re-emerging solution to the current problem of pathogenic microbes. Bacteriophage therapy has a number of potential advantages over the use of conventional
antibiotics, such as high bacterial specificity and efficacy against bacteria with
MDR, although there are concerns over its use, such as the possibility of inducing immunological responses. Nonetheless, phage therapy is generally regarded
as one of the most promising strategies to provide antimicrobial alternatives for
fighting antibiotic-resistant bacteria and could lead to the development of new and
improved therapies and diagnostics to combat infectious threats of the present
and the future.
In addition to the above approaches, there is a wide range of additional natural
compounds that have the potential in the treatment of infection. The antimicrobial properties of metals such as copper and silver have been known for centuries
especially in use for the treatment of burns and chronic wounds [26]. Recently, the
confluence of nanotechnology and the search for new agents in the fight against
microbes with MDR has brought metals in the form of nanoparticles to the fore
as potential antimicrobial agents. In Chapter 8, Sportelli and coworkers present
several examples of nanomaterials based on three of the main inorganic materials
with known antimicrobial action (i.e., silver, copper, and zinc oxide) along with
the mechanisms underlying their antimicrobial action. The potential applications of these nanoparticles as antimicrobials in areas such as prophylaxis and
therapeutics, medical devices, the food industry, and textile fabrics are discussed
in more detail. In addition, there are numerous examples of naturally produced
organic compounds with antibacterial properties. In the period 2000–2008,
over 300 natural metabolites with antimicrobial activity were reported, and in
Chapter 9, Saleem reviews these compounds and describes candidates with
potentially useful antimicrobial activity with reference to a variety of molecules,
including : alkaloids, acetylenes, coumarins, iridoids, terpenoids, and xanthones.
A range of organic compounds with the potential to serve as anti-infectives are
those that are known to sequester within bacterial cells and can be light-activated
to induce antimicrobial activity. For example, phenothiazinium-based molecules
[27, 28], whose antimicrobial properties were first noted in dyes that were used
for the histological staining of cellular components, have been shown to be more
efficacious than conventional antibiotics [28, 29]. These dyes photoinactivate bacteria, viruses, yeasts, fungi, and protozoa via the production of reactive oxygen
species (ROS) such as such as hydroxyl radicals and hydrogen peroxide. Over
the last few decades, photosensitizers (PS) have attracted increasing attention as
antimicrobial agents with therapeutic potential, and, when applied in this context, the use of PS is known as photodynamic antimicrobial chemotherapy (PACT).
Phoenix co-workers provide an overview of the photophysics and photochemistry
involved in PACT, and illustrate the therapeutic uses of this action with reference to a variety of PACT agents such as methylene blue and 5-aminolevulinic
acid. Whilst this area has clear potential, there are also challenges that need to
XIX
XX
Preface
be overcome if the use of such compounds is to become more widespread. One
such limitation is the challenge of ensuring effective light penetration of tissue
and in this respect, it has been suggested that ultrasound could be used as part of
a new antimicrobial strategy that addresses this limitation based on its superior
capacity for tissue penetration. Ultrasound has been shown to have an antibacterial effect comparable to some conventional antibiotics as recently reported in
the case of rhinosinusitis. It has also been shown that the application of ultrasound in conjunction with conventional antibiotics such as gentamycin is able to
synergize the effects of these drugs when applied to both planktonic and sessile
bacteria. More recently, it has been shown that irradiation with ultrasound can
activate some PS, which are generally termed sonosensitizers (SS) in this capacity, and based on these observations it was hypothesized that ultrasound and SS
may be exploited for the treatment of infectious diseases. This system has been
designated sonodynamic antimicrobial chemotherapy (SACT) and most recently
has been shown to be able to eradicate both Gram-positive and Gram-negative
bacteria. In Chapter 11, Harris coworkers provides an overview of the impact
of SACT.
In considering approaches to combat growing drug resistance and to identify
new means of treatment, the potential of oligonucleotides as antibacterial agents
has been investigated. Such molecules are able to act as antisense agents to prevent
translation, or, alternatively, can be designed to bind DNA to prevent gene transcription: these approaches are reviewed in Chapter 12 by Beaman coworkers.
In this area, a range of new and exciting approaches are being developed. For
example, it may be that such agents can inhibit microbial resistance mechanisms
by interrupting the expression of resistance genes and hence restore susceptibility
to key antibiotics, which would be co-administered with the antisense compound.
Such an approach will clearly have significant applications.
Finally, it is worth considering whether antibiotic efficacy can be increased
by enhancing the targeting of such molecules to their site of action. In the final
chapter, Ehlissi coworkers review an example of such an approach by looking
at targeting via the development of antimicrobial agent carrier systems such as
the use of nanoparticle constructs. Here, the authors discuss the development of
nanostructures for the entrapment and delivery of antimicrobials as an alternative to the direct application of these substances. Specific reference is made to
structures formed from liposomes and the effects of the carrier on the activity of
the compound are discussed.
In conclusion, it is clear that new approaches are needed if we are to maintain
our ability to deal with infection. These approaches have to be holistic and
integrated and must involve consideration of stewardship programs as well
as the development of new antibiotics and novel approaches to enhancing
activity through improved targeting or combination therapies. The need for the
development of new antibiotics and antibacterial design strategies has never
been greater.
March 2014
David A. Phoenix, Frederick Harris, and Sarah R. Dennison
Preface
Reference
1. Coates, A.R., Halls, G., and Hu, Y. (2011) 12. Zasloff, M. (1987) Magainins, a class of
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
Novel classes of antibiotics or more
of the same? Br. J. Pharmacol., 163,
184–194.
Powers, J. H. (2004). Antimicrobial drug
development – the past, the present, and
the future. Clin. Microbiol. Infect., 10
(Suppl 4), 23–31.
Boucher, H.W., Talbot, G.H., Bradley,
J.S., Edwards, J.E. Jr., Gilbert, D., Rice,
L.B., Scheld, M., Spellberg, B., and
Bartlett, J. (2009) Bad bugs, no drugs: no
ESKAPE! An update from the Infectious
Diseases Society of America. Clin. Infect.
Dis., 48, 1–12.
Berger, R.E. (2011) Emergence of a
new antibiotic resistance mechanism in
India, Pakistan, and the UK: a molecular,
biological, and epidemiological study
editorial comment. J. Urol., 185, 154.
Costelloe, C., Metcalfe, C., Lovering, A.,
Mant, D., and Hay, A.D. (2010) Effect
of antibiotic prescribing in primary
care on antimicrobial resistance in individual patients: systematic review and
meta-analysis. Br. Med. J., 340.
Overbye, K.M. and Barrett, J.F. (2005)
Antibiotics: where did we go wrong?
Drug Discov. Today, 10, 45–52.
Projan, S.J. and Shlaes, D.M. (2004)
Antibacterial drug discovery: is it all
downhill from here? Clin. Microbiol.
Infect., 10, 18–22.
Morel, C.M. and Mossialos, E. (2010)
Stoking the antibiotic pipeline. BMJ
(Clinical research ed.), 340, c2115.
Alvan, G., Edlund, C., and Heddini,
A. (2011) The global need for effective antibiotics – a summary of plenary
presentations. Drug Resist. Updat., 14,
70–76.
Davies, J. (2006) Where have all the
antibiotics gone? Can. J Infect. Dis. Med.
Microbiol. (Journal canadien des maladies infectieuses et de la microbiologie
medicale/AMMI Canada), 17, 287–290.
Katz, M.L., Mueller, L.V., Polyakov, M.,
and Weinstock, S.F. (2006) Where have
all the antibiotic patents gone? Nat.
Biotechnol., 24, 1529–1531.
13.
14.
15.
16.
17.
18.
19.
20.
21.
antimicrobial peptides from Xenopus
skin: isolation, characterization of two
active forms, and partial cDNA sequence
of a precursor. Proc. Natl. Acad. Sci.
U.S.A., 84, 5449–5453.
Izadpanah, A. and Gallo, R.L. (2005)
Antimicrobial peptides. J. Am. Acad.
Dermatol., 52, 381–390; quiz 391-2.
Zasloff, M. (2000) Reconstructing one of
nature’s designs. Trends Pharmacol. Sci.,
21, 236–238.
Cotter, P.D., Ross, R.P., and Hill, C.
(2013) Bacteriocins - a viable alternative
to antibiotics? Nat. Rev. Microbiol., 11,
95–105.
Chongsiriwatana, N.P., Patch, J.A.,
Czyzewski, A.M., Dohm, M.T., Ivankin,
A., Gidalevitz, D., Zuckermann, R.N.,
and Barron, A.E. (2008) Peptoids that
mimic the structure, function, and
mechanism of helical antimicrobial peptides. Proc. Natl. Acad. Sci. U.S.A., 105,
2794–2799.
Epand, R.F., Raguse, T.L., Gellman, S.H.,
and Epand, R.M. (2004) Antimicrobial
14-helical beta-peptides: potent bilayer
disrupting agents. Biochemistry, 43,
9527–9535.
Choi, S., Isaacs, A., Clements, D., Liu,
D., Kim, H., Scott, R.W., Winkler, J.D.,
and DeGrado, W.F. (2009) De novo
design and in vivo activity of conformationally restrained antimicrobial
arylamide foldamers. Proc. Natl. Acad.
Sci. U.S.A., 106, 6968–6973.
Niu, Y., Wu, H., Li, Y., Hu, Y., Padhee,
S., Li, Q., Cao, C., and Cai, J. (2013)
AApeptides as a new class of antimicrobial agents. Org. Biomol. Chem., 11,
4283–4290.
Niu, Y., Wang, R.E., Wu, H., and Cai,
J. (2012) Recent development of small
antimicrobial peptidomimetics. Future
Med. Chem., 4, 1853–1862.
Goodman, C.M., Choi, S., Shandler, S.,
and DeGrado, W.F. (2007) Foldamers as
versatile frameworks for the design and
evolution of function. Nat. Chem. Biol.,
3, 252–262.
XXI
XXII
Preface
of antimicrobials. Biotechnol. Adv., 27,
76–83.
Hancock, R.E. (2006) Antibacterial peptides for therapeutic use: obstacles and
27. Harris, F. and Phoenix, D.A. (2006) Light
realistic outlook. Curr. Opin. Pharmacol.,
activated compounds as antimicrobial
6, 468–472.
agents – patently obvious? Recent Pat.
Hancock, R.E. and Sahl, H.G. (2006)
Antiinfect. Drug Discov., 1, 181–199.
Antimicrobial and host-defense peptides 28. Phoenix, D.A., Sayed, Z., Hussain, S.,
as new anti-infective therapeutic strateHarris, F., and Wainwright, M. (2003)
gies. Nat. Biotechnol., 24, 1551–1557.
The phototoxicity of phenothiazinium
Gellman, S. (2009) Structure and funcderivatives against Escherichia coli and
tion in peptidic foldamers. Biopolymers,
Staphylococcus aureus. FEMS Immunol.
92, 293.
Med. Microbiol., 39, 17–22.
Wu, Y.D. and Gellman, S. (2008) Pep29. Phoenix, D.A. and Harris, F. (2003)
tidomimetics. Acc. Chem. Res., 41,
Phenothiazinium-based photosensitizers:
1231–1232.
antibacterials of the future? Trends Mol.
Rai, M., Yadav, A., and Gade, A. (2009)
Med., 9, 283–285.
Silver nanoparticles as a new generation
22. Marr, A.K., Gooderham, W.J., and
23.
24.
25.
26.
1
1
The Problem of Microbial Drug Resistance
Iza Radecka, Claire Martin, and David Hill
1.1
Introduction
Microbial colonization, where it is not wanted, can lead to disease, disability,
and death. Therefore, control and/or destruction of pathogenic microorganisms is crucial for the prevention and treatment of disease. Modern medicine
is dependent on antimicrobial/chemotherapeutic agents such as antibiotics
(Greek anti, against, bios life). Antibiotics can either destroy pathogens or
inhibit their growth and avoid damage to the host. In the nineteenth century,
infections such as diarrhea, pneumonia, or post-surgical infections were the
main causes of death. Therefore, the discovery of antibiotics was of great
importance to society and impacted on the prevention and treatment of
disease. Antibiotics can be defined as compounds produced by microorganisms
that are effective against other microorganisms but nowadays also include
microbial compounds that have been synthetically altered. The classification
of antibiotics is based not only on the cellular components or systems they
affect but also on whether they inhibit cell growth (bacteriostatic drug) or kill
the cells (bactericidal drug) [1]. Other chemotherapeutic synthetic drugs, not
originating from microbes, such as sulfonamides, are also sometimes called
antibiotics [2].
1.2
History of the Origins, Development, and Use of Conventional Antibiotics
The modern era of antimicrobial agents began with the work of the German
scientist Paul Ehrlich (1854–1915), who, together with a Japanese scientist
Sahachiro Hata (1873–1938), discovered in 1909 the first sulfa drug called
arsphenamine – initially known as compound “606” (the 606th compound
tested). This new drug was available for treatment in 1910 under the trade name
Salvarsan. Arsphenamine, considered as a “magic bullet” with selective toxicity,
was used in the treatment of syphilis and sleeping sickness. Despite the fact that
Novel Antimicrobial Agents and Strategies, First Edition.
Edited by David A. Phoenix, Frederick Harris and Sarah R. Dennison.
© 2015 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2015 by Wiley-VCH Verlag GmbH & Co. KGaA.
2
1 The Problem of Microbial Drug Resistance
the mode of action of arsphenamine remained unclear, it was the most popular
antimicrobial drug successfully used until the 1940s [2, 3].
After Ehrlich’s success, many more compounds were tested for their possible
antimicrobial properties. In the 1930s, Gerard Domagk (1895–1964) tested a
number of leather, nontoxic (for animals) dyes for their antimicrobial activity.
His work led to the discovery of Prontosil Red (1932), the first sulfa antimicrobial
agent effective against pathogenic streptococci and staphylococci. This discovery
was so important that in 1939 he received the Nobel Prize for its discovery.
However, it was the discovery of the first antibiotic called penicillin that revolutionized the treatment of infectious diseases and initiated the new antibiotic
era. Although penicillin was first discovered by a French medical student Ernest
Duchesne in 1896, it was Alexander Fleming (1881–1955) who first observed
the lethal/antimicrobial activity of the substance, which he later named penicillin, against Staphylococcus aureus. He reported (1928) the inhibition of the
growth of pathogenic bacteria contaminated with Penicillium notatum spores.
Fleming published several papers about penicillin production and began efforts
to characterize penicillin. Unfortunately, he stopped his research with penicillin
at this stage as he was not able to demonstrate the stability of penicillin within
the body. In 1930, Fleming’s paper about penicillin produced by P. notatum was
again an object of great interest to Professor Howard Florey (1898–1968) and his
coworker Ernest Chain (1906–1979) who were investigating the antimicrobial
properties of many substances including Fleming’s penicillin. Crude penicillin
produced by P. notatum (Fleming’s strain) was purified and successfully tested
against staphylococci and streptococci. In March 1942, the first adult patient
was successfully treated with penicillin, which led to both scientists receiving
the Nobel Prize in 1945. In 1943, a new strain of Penicillium chrysogenum was
isolated from a moldy cantaloupe by Mary Hount from the Horthen Regional
Research Laboratory, Illinois, US, and the mass production of penicillin began [3].
In 1944, Selman Waksman, after screening about 10 000 strains of soil bacteria
and fungi, discovered a new antibiotic produced by Streptomyces griseus called
streptomycin. For his success, he received the Nobel Prize in 1952. By 1953,
production of chloramphenicol, neomycin, tobramycin, and tetracycline was also
possible [2].
Cephalosporins are the second class of antibiotics following penicillins. In
1945, Giuseppe Brotzu (1895–1955) isolated Cephalosporium acremonium from
sewage water in Sardina, Italy. Brotzu observed great antimicrobial activity
against some Gram-negative bacteria. Unable to proceed with his research,
Brotzu sent his cultures to Edward Abraham (Oxford University) who, together
with Guy Newton, isolated cephalosporin P, active only against Gram-positive
bacteria. Shortly after, cephalosporin N and cephalosporin C were discovered
(paper published in 1961). Cephalosporin N was later identified to be penicillin
N – active against both Gram-negative and Gram-positive bacteria.
Modern antibiotics used today are, or derive from, natural molecules isolated
during the “golden age” of antibiotic era (1940–1970) mostly from Streptomyces
species, a few from Gram-positive Bacillus species, and some from strains of
1.2
History of the Origins, Development, and Use of Conventional Antibiotics
3
Table 1.1 Examples of natural, semi-synthetic and synthetic antibiotics and their mode of
action [1, 3, 4, 6].
Group of
antibiotics
Mode of
action
Primary
target
Derivation
Organisms
β-lactams
Inhibition of cell
wall synthesis
Inhibition of cell
wall synthesis
Inhibition of RNA
synthesis
Penicillin binding
protein
Peptidoglycan
units
RNA polymerase
Natural and
semi-synthetic
Natural and
semi-synthetic
Natural and semi
synthetic
Gram-positive and
Gram-negative bacteria
Gram-positive bacteria
Inhibition of cell
wall synthesis
Inhibition of
protein synthesis
Cell membrane
Natural and semi
synthetic
Natural and semi
synthetic
Tetracyclines
Inhibition of
protein synthesis
30S ribosome
Natural and semi
synthetic
Macrolides
Inhibition of
protein synthesis
50S ribosome
Natural and semi
synthetic
Streptogramins
Inhibition of
protein synthesis
50S ribosome
Natural and semi
synthetic
Phenicols
Inhibition of
protein synthesis
50S ribosome
Natural and semi
synthetic
Trimethoprimsulfamethoxazole
Inhibition of DNA Inhibition of
synthesis
synthesis of
tetrahydrofolic
acid
Inhibition of DNA Topoisomerase II
synthesis
and IV
Glycopeptides and
glycolipopeptides
Rifamycins
Lipopeptides
Aminoglycosides
Fluoroquinolones
30S ribosome
Synthetic
Synthetic
Gram-positive and
Gram-negative bacteria,
M. tuberculosis
Gram-positive and
Gram-negative bacteria
Aerobic Gram-positive
and Gram-negative
bacteria, M. tuberculosis
Aerobic Gram-positive
and Gram-negative
bacteria
Aerobic and anaerobic
Gram-positive and
Gram-negative bacteria
Aerobic and anaerobic
Gram-positive and
Gram-negative bacteria
Some Gram-positive
and Gram-negative
bacteria
Gram-positive and
Gram-negative bacteria
Aerobic Gram-positive
and Gram-negative
bacteria; some
anaerobic
Gram-negative bacteria
and M. tuberculosis
Penicillium and Cephalosporium [4, 5]. Most bactericidal antibiotics kill the cell
by interfering with the essential cellular processes (Table 1.1). They inhibit DNA,
RNA, cell wall, or protein synthesis [1, 3, 4, 6].
Interestingly, it was also Fleming who, in his Nobel lecture, stated that bacteria can develop resistance to penicillin if exposed to low doses and that negligent
use could encourage resistance. Sadly, he was right, and soon after penicillin G
was introduced to hospitals (1940s) the problem of antibiotic-resistant bacteria
4
1 The Problem of Microbial Drug Resistance
emerged [7]. Only 3 years after his warning, 38% of S. aureus strains in only one
London hospital were penicillin resistant. Currently, around 90% of strains in the
United Kingdom and nearly all in the United States show penicillin resistance [8].
Antibiotic resistance (AR) is driven by the misuse of antibiotics due to selective
pressure. Moreover, unprecedented human air travel allows bacterial mobile resistance genes to be transported between continents. So the fact that bacteria and
their resistance genes can travel faster and further than ever before creates serious
risk to human health and development on a global scale [9, 10]. At the moment, in
Europe at least 25 000 patients die every year because of bacterial infections, which
cannot be treated with the available antibiotics [11]. Therefore, the development
of new antimicrobial drugs with new modes of action and the preservation of the
agents “in hand” are essential steps for the foreseeable future [7]. Great efforts
have also been made to understand the mechanisms by which currently available
antibiotics affect microbial cells. Antibiotic-facilitated cell death is very complex
and involves many genetic and biochemical pathways. It is essential to understand
the multilayered mechanisms by which currently available antibiotics kill bacteria,
and also create new alternative antimicrobial therapies [1].
1.3
Problems of Antibiotic Resistance
Unquestionably, the discovery of antibiotics was one of the most important
medical achievements in modern medicine and their introduction represents a
remarkable success story for society. However, the widespread use and misuse of
antibiotics for both clinical and nonclinical settings has resulted in the emergence
(selection) of a number of multiresistant bacteria called superbugs such as
methicillin-resistant Staphylococcus aureus (MRSA), vancomycin-intermediate
Staphylococcus aureus (VISA) [12], vancomycin-resistant Enterococcus spp.,
[10] carbapenem-resistant Mycobacterium tuberculosis [5], extended spectrum
β-lactamase-producing Escherichia coli, or the highly virulent antibiotic-resistant
Clostridium difficile [11, 13]. The emergence of antibiotic resistance in bacteria,
selected by negligent antibiotic usage, provides the most dramatic demonstration
of Darwinian selection as a result of a specific evolutionary pressure to adapt to
the presence of antimicrobials [14]. It has been reported that the consumption
of antimicrobials by food-producing animals around the world is also a powerful
driver of antibiotic multidrug resistance (AMR) in both humans and animals [8].
These activities also clearly create an ongoing explosion of antibiotic-resistant
infections generating a significant risk to public health on a global scale, as
there are very few or sometimes no effective antimicrobial agents available to
treat infections caused by both Gram-positive and Gram-negative pathogenic
bacteria [15, 16]. The problem of ever-increasing bacterial multiresistance is even
more alarming when we consider the diminishing number of new antimicrobials entering clinical practice [17, 18]. There is clearly an urgent need for the
development of new antibiotics or new alternatives to conventional antimicrobial
1.5
MDR Mechanisms of Major Pathogens
agents with novel mechanisms of antimicrobial action as even some common
infections are becoming increasingly difficult to treat. It is also very important
to stress that antimicrobial resistance is not only found in bacteria – that there
is a growing number of other pathogens such as viruses (that cause chronic
hepatitis B (CHB) or influenza), parasites (cause malaria), and fungi (Candida
infections) resistant to the antimicrobial agents [6, 19, 20]. Resistance to all
classes of antimalarial drugs has been well documented including artemisinin
derivatives and chloroquine. Moreover, resistance rates (10–20%) to anti-HIV
drug regimens have been reported in the United States and Europe. Many people
around the world suffer because of antimicrobial resistance.
1.4
Multiple Drug-Resistant (MDR), Extensively Drug-Resistant (XDR), and
Pan-Drug-Resistant (PDR) Organisms
There are many definitions in the medical literature used to characterize different patterns of bacterial multiresistance. International organizations such as the
European Centre for Disease Prevention and Control (ECDC), the Clinical Laboratory Standards Institute (CLSI), the European Committee and Antimicrobial
Susceptibility Testing (EUCAST), and the United States Food and Drug Administration (FDA) have made a combined effort to create standardized terminology that can be applied to all bacteria responsible for infections associated with
multidrug resistance [18, 21]. Consequently, “antimicrobial categories” were created (for each specific organism or group), each category containing the related
antimicrobial agents (Table 1.2). The term multiple drug resistance (MDR) refers
to organisms non-susceptible to at least one agent in three or more antimicrobial categories. Extensively (extreme) drug resistant (XRD) means the organism
shows non-susceptibility to at least one agent in all but two or fewer antimicrobial categories and pan-drug resistant (PDR) refers to an organism that shows
non-susceptibility against all (or nearly all) of the antimicrobial agents within the
antimicrobial categories.
1.5
MDR Mechanisms of Major Pathogens
At present, the treatment of bacterial infections is severely affected by the
emergence of antibiotic-resistant infections and epidemic increases of multidrug resistant (MDR), XRD, or increasingly PDR microorganisms [22] such
as vancomycin-resistant Enterococcus faecium (VRE), Enterobacter cloacae,
MRSA), XRD carbapenem-resistant Acinetobacter baumannii [8], third generation cephalosporin-resistant E. coli, third generation cephalosporin-resistant,
extended spectrum β-lactamase producing Klebsiella pneumonia (ESBL-KP),
carbapenem-resistant Klebsiella pneumoniae (CRKP) [8], carbapenem-resistant
5
6
1 The Problem of Microbial Drug Resistance
Table 1.2 Examples of antimicrobial categories and antimicrobial agents used to define
MDR, XDR and PDR [18].
Antimicrobial category
Antimicrobial agent
Carbapenems
Imipenen
Meropenem
Doripenem
Tetracycline
Doxycycline
Minocycline
Gentamicin
Tobramycin
Amikacin
Netilmicin
Colistin
Polymyxin B
Cefotaxime
Ceftriaxone
Ceftazidime
Vancomycin
Teicoplanin
Chloramphenicol
Streptomycin
Tetracyclines
Aminoglycosides
Polymyxins
Extended spectrum cephalosporins
third and fourth generation
Glycopeptides
Phenicols
Streptomycin
Pseudomonas aeruginosa, multidrug resistant Mycobacterium tuberculosis
(MDR-TB) [23], and C. difficile [6, 13, 15, 24–29].
Drug resistance can be caused by mobile genes or, in the absence of mobile
genetic elements, by sequential mutations in the microbial chromosome. Mobile
genes can be transferred between different bacteria by mobile genetic elements
such as plasmids, naked DNA, transposons, or bacteriophages. These genes code
for information against a particular antibiotic. In some microbes, multiple genes
can be present, resulting in MDR. Alternatively, resistance or MDR can also be
caused by sequential mutation in chromosomal DNA, which can result in mutation in the antibiotic target enzymes (topoisomerases) or/and in the overexpression of efflux pumps that expel structurally unrelated drugs [6, 30]. Chromosomal
genes can also be transferred. They can be acquired by one bacterium through
the uptake of naked DNA released from another microorganism by the process
called transformation (an introduction of an exogenous DNA into a cell, resulting
in a new phenotype). For example, emergence of high-level resistant S. aureus to
vancomycin, caused by a mobile element – transposon from enterococci – first
appeared in response to an intermediate dose of vancomycin. Bacteria are also
mobile and can easily travel from person to person, from continent to continent,
spreading the problem of microbial resistance [10].
1.5
MDR Mechanisms of Major Pathogens
Bacterial mechanisms of resistance vary. Active resistance can be achieved by
three major mechanisms: first, synthesis of specific enzymes that selectively target
and inactivate the drug (e.g., β-lactamases, macrolide esterases, epoxidases, or
several transferases); second, efflux of the antimicrobial agents from the cell via
membrane-associated efflux pumps; third, modification of the antibiotic target
sites (alteration of intracellular binding targets such as ribosomal RNA or DNA
gyrase involved in DNA replication, or even enzymes involved in the synthesis of
bacterial cell wall). The most important example of target change can be seen in
MRSA, where the acquisition and expression of mecA genes results in resistance
to methicillin and most of the β-lactam antibiotics [30–33]. All three mechanisms
can make the drug incapable of inhibiting microbial metabolic pathways that are
vital for microbial growth and survival [6, 31].
Antimicrobial resistance to a single antimicrobial agent is already problematic,
but the emerging multidrug resistance of Gram-negative bacteria is of serious
concern and it dramatically limits treatment options [25]. Gram-negative bacterial infections caused by MDR or PDR bacteria (such as E. coli, P. aeruginosa,
Klebsiella pneumoniae, and/or A. baumannii) can result in death. In 2013, they
were called the nightmare bacteria by the US Centres for Disease Control and
Prevention (CDC) and a new coming “red plaque” by Looke et al. [34]. The predominant cause of resistance of Gram-negative bacteria is related to one or more
β-lactamases, which can inactivate the β-lactam antibiotic by hydrolyzing the
amide bond of the β-lactam ring, leaving β-lactam antibiotics harmless to bacteria
[35–37]. In the 1980s, only cephalosporins (e.g., cefotaxime) were less susceptible
to β-lactamases; unfortunately, their repetitive use selected resistant strains able
to produce plasmid-mediated enzymes such as cefotaximasas (CTX-M) [35].
Research shows that ESBLs carried by E. coli and metallo-β-lactamases (SHV-1,
sulfhydryl variable) carried by K. pneumoniae and Enterobacter spp can easily
destroy the latest generation of penicillins or cephalosporins. They can even
inactivate carbapenems, which are often called the last available resort for
treatment of serious infections caused by Gram-negative bacteria [13, 35, 36].
Most of the species of E. coli, responsible for urinary tract infections and
Gram-negative bacteremia, are antibiotic sensitive, apart from being resistant
to ampicillin. However, research showed that up to 60% of E. coli isolates from
hospital and non-hospital environments are resistant to ampicillin because of
the production of plasmid-mediated TEM-1/2 (Temoniera from whom E. coli
TEM was isolated in 1963) β-lactamases [13, 35]. TEM-2 enzyme differs from
TEM-1 only by a single amino acid [13]. Microbes producing TEM-1 or TEM-2
are known to be resistant to ampicillin but still are susceptible to the third
generation of cephalosporins. However, it has been reported that mutations in
TEM-1 and TEM-2 can result in the production of new ESBLs (so far more than
100 of new TEM have been reported). Transferrable plasmids containing genes
encoding ESBLs are often associated with aminoglycoside resistance and other
resistances [13]. In 1990, the more virulent MDR CTX-M-producing E. coli has
emerged, replacing opportunistic hospital outbreaks with SHV- and TEM-type
ESBLs-KP. It was established that CTX-M enzymes are encoded in transferrable
7
8
1 The Problem of Microbial Drug Resistance
plasmids and transposons. These mobile elements have originated from other
bacteria such as Kluyvera spp. and have spread widely among enterobacteria
[13]. Highly transmittable CTX-M-producing E. coli can also be resistant to
the aminoglycosides and quinolones. As a result of this, MDR in E. coli is now
increasingly common in the hospital environments and community. It is also
known that phages can be involved in bacterial evolution and the creation of new
“super bugs” such as the deadly E. coli O157 : H7 strain [38].
K. pneumoniae, one of the most common clinical pathogen causing sepsis,
meningitis, pneumonia, and other diseases, is usually resistant to ampicillin by
production of a metallo-β-lactamases (SHV-1), similar to TEM-1 or TEM-2.
SHV-1 can be encoded by transferrable plasmid, integrons, or by chromosome mutations. Mutation of SHV-1 results in the production of one or
more ESBLs. MDR K. pneumoniae is becoming a serious concern worldwide.
Carbapenem-resistant organisms can produce several different carbapenemases.
Plasmid-mediated Klebsiella pneumoniae carbapenemase (KPC) isolates were
found to be responsible for many outbreaks worldwide and were associated with
a significant mortality rate [36, 39, 40].
Gram-negative P. aeruginosa, another well-known opportunistic pathogen [30],
is the third most common cause of hospital-acquired Gram-negative bacteremia
after E. coli and K. pneumonia. Some isolates of P. aeruginosa are inherently resistant to most penicillins, cephalosporins, and even to carbapenems [13, 28, 36, 41].
This multidrug resistance was caused by the overexpression of the chromosomally encoded efflux system, which is very common in Gram-negative bacteria such
as A. baumannii or P. aeruginosa [32]. In A. baumannii the efflux system is associated with the resistance-nodulation-cell division (RND) family of the transport
proteins. These multidrug efflux pumps consist of an efflux membrane transporter
(RND) that can interact with an outer membrane factor (OMF), which exports the
drug through both membranes [33]. A. baumannii shows an extraordinary ability
to develop multidrug resistance due to a high level of genomic plasticity and due
to mutation of endogenous genes. Alteration of these genes exhibits overexpression of the chromosomally encoded β-lactamases, loss of expression of porins,
mutation in gyrA and parC, and finally overexpression of efflux systems, which is
associated with increased drug resistance. There are three types of efflux systems
in A. baumannii: CraA (resistance to chloramphenicol); AbeM (extrudes several
antimicrobials); and AmvA (resistance to several detergents, dyes, disinfectants,
and erythromycin). There are also several tetracycline efflux pumps, for example,
TetA and TetB. TetA is associated with resistance to tetracycline, while TetB shows
resistance to tetracycline, doxycycline, and minocycline [30, 33, 42].
Gram-positive S. aureus has a great ability to develop multiple resistances [29].
Reports showed that it can be resistant to penicillin, tetracycline, erythromycin,
chloramphenicol, gentamicin, and methicillin. The MRSA, called the superbug,
emerged in 1961, only 2 years after methicillin was introduced and since then it
has become the most common multiple-antibiotic-resistant pathogen in many
parts of the world [27]. In 2011, it was estimated that MRSA was responsible
for 171,200 healthcare-associated infections (HAIs) in Europe per year, 5400
1.5
MDR Mechanisms of Major Pathogens
Table 1.3 Examples on antibiotics with activity against MRSA [15].
Antibiotic
Mode of action
Daptomycin
Causes a calcium ion-dependent disruption of bacterial cell
membrane, an efflux of potassium inhibits RNA, DNA, and
translation
Inhibits bacterial protein synthesis by binding to the domain V
regions of 23S rRNA
Bacteriostatic against most pathogens. Show broad spectrum of
antimicrobial activity. Inhibits bacterial protein synthesis by binding
to the 30S ribosomal sub-unit blocking binding of amino-acetyl
transfer RNA into acceptor side
Linezolid
Tigecycline
deaths, and more than a million extra days of hospitalization [29, 43]. Methicillin
resistance developed because of the acquisition of the mecA gene located on a
large genetic element called staphylococcal cassette chromosome mec (SCCmec)
integrated into the MRSA chromosome [27]. SCCmec has been possibly assimilated by horizontal transfer from an animal coagulase-negative pathogen
Staphylococcus sciuri. mecA gene encodes for the production of an abnormal
penicillin-binding protein PBP-2a (also called PBP-2′ ). PBPs are transpeptidases
necessary for cell wall peptidoglycan synthesis and are the target for penicillin.
PBP-2a is a transpeptidase that does not bind to penicillin so inhibition of
cell wall synthesis by penicillin does not occur [13, 31]. Many strains of MDR
MRSA remain susceptible only to vancomycin and teicoplanin (glycopeptides).
Unfortunately, in recent years some S. aureus isolates have also become glycopeptide tolerant, and even worse, several isolates now show glycopeptide
and carbapenems resistance. New antibiotics against MRSA infections such as
daptomycin (Cubicin®; Novartis), linezolid, and tigecycline (Tygacil®; Wyeth)
have been investigated (Table 1.3). However, a number of novel agents such as a
capsular polysaccharide-based vaccine, lipoglycopeptide ortivanacin, or the use
of signal molecule-based drugs (quorum sensing inhibitor) or cell wall-anchored
adhesions are in different stages of development [14, 29].
Hospital-acquired MRSA (HA-MRSA) have now been found outside the hospitals and spread to other healthcare facilities [27]. There is also massive spread
of community-acquired MRSA (CA-MRSA) infections. Some CA-MRSA isolates
can produce toxins called Panton-Valentine Leukocidin (PVL), which increases
its virulence. Expression of this virulence is controlled by complex staphylococcal
regulatory networks including the accessory gene regulator (agr) system. These
genes can vary between different strains [29, 43]. PVL is responsible for acute
skin infections and pneumonia. CA-MRSA can be easily transmitted from person
to person.
The development of antiviral drug resistance also represents serious complications. CHB virus is an example of antiviral drug resistance. The development of
resistance in hepatitis B virus (HBV) is related to the lack of proofreading function
in the DNA polymerase and its high replication rate [19], which, in the presence of
9
10
1 The Problem of Microbial Drug Resistance
the antiviral drug, result in specific DNA mutations (during replication process).
Clinically, antiviral drug resistance is first exhibited in higher levels of HBV DNA
(virological breakthrough), followed by increased levels of alanine aminotransferase (biochemical breakthrough). Although the DNA mutations developed
can affect the “fitness” of the viruses, they will also help the virus to survive the
presence of the drug and develop a high level of antiviral drug resistance. In
addition, compensatory additional DNA mutations help restore viral “fitness,”
leading to viral rebound. HBV shows a high level of resistance to antiviral drugs
such as lamivudine, telbivudine, and adefovir. Rapid development of antiviral drug
resistance has also been seen for influenza viruses A and B. There are two classes
of antiviral drugs approved in many countries: the adamantanes (active only
against influenza virus A) and the neuraminidase inhibitors (NAIs). However, due
to the rapid emergence of viral resistance, only NAIs are recommended by WHO
(since 2010) for the treatment or prophylaxis of influenza A and B infections. At
present, only two NAIs are licensed worldwide for therapeutic and prophylactic
uses: oseltamivir, commercially available as Tamiflu® (F. Hoffmann-La Roche),
and zanamivir, commercially available as Relenza® (GlaxoSmithKline). In 2009,
influenza pandemic patients with suspected or known influenza A (H1N1)pdm09
were treated with a new drug peramvir (BioCryst). The mechanism of resistance
is also linked to DNA mutations. Influenza viruses showing reduced sensitivity
to NAIs contain mutations, which directly or indirectly change the shape of
the influenza surface antigen–neuraminidase (NA) catalytic sites (made of 8
functional and 11 framework residues). The NA surface antigen exhibits two
important functions: first, it releases progeny virions, and second, it facilitates
viral spread. Any alterations to NA catalytic sites reduce the inhibitor binding of
the drug and therefore lower the efficiency of Tamiflu. In 2007, H1N1 influenza
strains in Europe and North America were reported resistant to NAI Tamiflu
owing to the H274Y mutation [44]. Rapid evolution of influenza surface genes
can create more worldwide dissemination of drug-resistant influenza infections
caused by A(H1N1) variants; therefore, the development of new antiviral drugs
and surveillance of viral infections is extremely important [45].
Pathogenic fungi such as MDR Candida spp. or MDR Candida krusei are known
to be responsible for life-threatening infections. They are called hidden killers
resulting in 46–75% mortality [46]. The multidrug resistance in Candida spp.
is related to low accumulation of drugs caused by genes encoding drug transporters. ATP-binding cassette (ABC) transporters are encoded by Candida drug
resistance (CDR1 and CDR2) and a major facilitator superfamily (MFS) transporter encoded by MDR1 genes. Overexpression of MDR1, which encodes the
MDR efflux pump of the MFS often, increases resistance to azole antifungal drugs.
Long-term therapies with fluconazole (antifungal drug) have led to the emergence
of fluconazole-resistant Candida albicans and C. krusei strains, which can also be
resistant to other drugs. C. krusei also showed decreased susceptibility to flucytosine and amphotericin [47]. Novel antimicrobial peptides that can target the
mitochondria and DNA of MDR Candida spp. are being developed in order to
fight microbial resistance [48].
1.6
Antimicrobial Stewardship Programs
1.6
Antimicrobial Stewardship Programs
Antimicrobial resistance has been recognized as a major global threat. Globalization of the world results in population movement, which favors the rapid spread
of new MRD organisms and infectious diseases [16]. The dramatic increase in
antibiotic-resistant infections leads to higher mortality, longer hospital stays,
and unavoidably increased treatment costs [49]. It can be said that the gene pool
for antimicrobial resistance has never been so big nor its selection pressure so
strong [1]. There was a time when antimicrobial agents were highly successful in
treating infections caused by pathogenic microbes; however, their unfettered use
in human clinical therapy, aquaculture, and food animal production has triggered
rapid development of antimicrobial resistance, especially in the developing world
[34, 38]. In recent years, the scientific community has raised serious concerns
about the fact that drug development will not be able to address the problem
posed by drug or multidrug resistance. So what do we do next? How do we
fight this multiresistance problem? Recognizing the serious global problem,
several nations, international health agencies, and many other organizations
worldwide have taken actions to counteract microbial resistance through the
application of novel strategies/initiatives. For example, the ARTEMIS Antifungal
Surveillance Program (2001–2005) was created to increase our understanding
and to monitor the spread of the uncommon but MDR fungal pathogen C. krusei
[47]. In 2001, a WHO global strategy was introduced in order to slow down and
reduce the spread of antimicrobial-resistant organisms. The strategy included
better access to appropriate antimicrobial agents, better use of antimicrobials,
better surveillance of antimicrobial resistance by strengthening health systems,
and enforcing of regulations and legislation. The strategy also included the
development of new drugs and vaccines [24]. In 2008, to avoid the spread of
resistance, virological monitoring of HIV patients was required. To facilitate this,
the WHO developed a Global Strategy for Prevention of HIV Drug Resistance
and established the HIVResNet network of experts and laboratories in order
to reduce the spread of HIV infection due to resistance to anti-HIV drugs
[18]. In 2011, during the World Health Day, the WHO urged the world for a
political commitment and the creation of a comprehensive plan that may help
fight antimicrobial resistance. Following advice, hospitals implemented novel
initiatives such as antimicrobial stewardships. The antimicrobial stewardship is
a combined set of strategies/guidelines created to reduce microbial resistance
[50–52]. The mission of antimicrobial stewardship program is to reduce inappropriate use of antimicrobial agents (dose, duration, route of administration) and
improve patient outcomes [53]. Antimicrobial stewardship also aims to reduce
the spread of infections and the development of antimicrobial resistance [52].
Several studies showed that antibiotic resistance can be reduced by shortening
the length of antibiotic courses [49, 54, 55]. An antimicrobial stewardship
program is multidisciplinary, and brings together key healthcare professionals
(nurses, general practitioners, pharmacists, clinical microbiologists, infectious
11
12
1 The Problem of Microbial Drug Resistance
diseases physicians) and hospital management. The program cannot be just
limited to large hospitals or academic centers only; it needs to include small
regional facilities in both developed and developing healthcare systems around
the world. Adopting novel stewardship strategies in the hospitals and community
can provide a systematic approach to the growing threat of antibiotic resistance.
1.7
Discussion
Multidrug resistance is a serious danger to the future of humans (and animals)
and could result in the development of untreatable diseases and death. Immediate
measures must be taken worldwide to safeguard the remaining antimicrobials and
to facilitate the development of new antimicrobial agents. Numerous papers have
been published about microbial multidrug-resistant pathogens created by intensive use of antibiotics in human and animal therapy, and food animal production.
As a consequence, AR and MDR bacteria have been found in hospitals, healthcare centers, community, and various food products and even in the environments
without a history of direct exposure to antibiotics [38]. Enteric bacteria such as
E. coli or Enterococcus spp. have been extensively investigated and the impact of
these commensal bacteria on the emergence of MDR gene pool has been recognized by the scientific community [38]. In addition, recent reports, released by
the WHO, have raised concerns about MDR tuberculosis (MDR)-TB. In January
2010, 58 countries reported cases of XDR-TB [56].
It can be said that the problem of bacterial resistance or multidrug resistance
in the ecosystem is a serious and complex issue. How should we manage and
prevent multidrug resistance? First, detailed knowledge of the nature of AR and
MDR pathogens is required in order to implement new successful strategies to
control the transmission of multidrug resistance within hospital/healthcare environments and the community. Investigations of historic strain and events that
have led to the origin of resistant strains such as mechanisms involved in the
emergence and dissemination are essential. Secondly, the rate of MDR microbes
can be reduced by the implementation of different intervention strategies relevant to the control of antibiotic use and control of hospital infections [18, 41, 57].
The control of antibiotic use focuses on factors such as choice of antibiotic or
combination of antibiotics for treatment, duration of the therapy, monitoring and
feedback on antimicrobial resistance, rational antibiotic usage, and regulations.
ESBL-producing E. coli was reported in 79% of surgical wound infections in India,
but only less than 5% in New Zealand [11, 57]. MRSA are highly dominant in the
United States (34%) while in the Netherlands the prevalence is ≤2%. These massive differences could be associated to the local differences in antibiotic policies
[57].
The second essential point is the control of hospital infection (to control
cross-infections within hospitals and within community) through rapid detection/diagnostic tests, prevention, and control of antimicrobial resistance, and also
References
plans for patients (admitted, re-admitted, discharged, or transferred) colonized
with resistant microbes. For example, to prevent further spread of MRSA the
“Search-and-Destroy” policy was implemented in Denmark in 1983, when
prevalence of MRSA was 30%; since then, the MRSA prevalence decreased to
less than 1% [27]. Nemeth et al. [57] reported that a substantial proportion of
the patients transferred to the hospital (Zurich University Hospital) from abroad
are colonized with MDR organisms. They concluded that the rigors of infection
control in the hospital are important and every patient admitted to the hospital
should be screened for MDR colonization. Reducing the spread of MDR by using
existing antibiotics and developing new antibiotics or novel alternative therapies
should be seen as a collective responsibility [8]. Worldwide surveillance systems
of resistance and reduction of transmission of resistant organisms are required to
generate valuable data important for research and development. Managing the
problem of bacterial resistance requires researchers, community, and politicians
working together to promote research and implement global strategies.
Acknowledgment
We thank the University of Wolverhampton for the support given in the preparation of this chapter.
References
1. Kohanski, M.A., Dwyer, D.J., and
2.
3.
4.
5.
6.
Collins, J.J. (2010) How antibiotics
kill bacteria: from targets to networks.
Nat. Rev. Microbiol., 8, 423–435.
Prescott, L.M., Harley, J.P., and Klein,
D.A. (2002) Microbiology, 5th edn, Mc
Grow-Hill Companies.
Zaffiri, L., Gardner, J., and
Toledo-Pereyea, L.H. (2012) History
of antibiotics. From salvarsan to
cephalosporins. J. Invest. Surg., 25,
67–77.
Fernebro, J. (2011) Fighting bacterial
infections – future treatment options.
Drug Resist. Updat., 14, 125–139.
Zucca, M., Scutera, S., and Savoia, D.
(2011) in Drug Development – A Case
Study Based Insight into Modern Strategies (ed. C. Rundfeldt), In Technologies,
pp. 123–162.
Levy, S.B. and Marshall, B. (2004)
Antimicrobial resistance worldwide
causes, challenges and responses. Nat.
Med. Suppl., 10, 122–129.
7. Hiramatsu, K., Igarashi, M., Morimoto,
Y., Baba, T., Umekita, M., and Akamatsu,
Y. (2012) Curing bacteria of antibiotic
resistance: reverse antibiotics, a novel
class of antibiotics in nature. Int. J.
Antimicrob. Agents, 39, 478–485.
8. Huttner, A., Harbarth, S., Carlet, J.,
Cosgrove, S., Goossens, H., Holmes,
A., Jarlier, V., Voss, A., and Pittet, D.
(2013) Antimicrobial resistance: a global
view from the 2013 World Healthcareassociated Infections Forum. Antimicrob.
Resistance Infect. Control, 2, 31.
9. Grundmann, H., Klugman, K.P., Walsh,
T., Pardo, P.R., Sigauque, B., Khan, W.,
Laxminarayan, L., Heddini, A., and
Stelling, J. (2011) A framework for global
surveillance of antibiotic resistance. Drug
Resist. Updat., 14, 79–87.
10. Kumarasamy, K.K., Toleman, M.A.,
Welsh, T.R., Bagaria, J., Butt, F.,
Bakakrishnan, R., Chaundhary, M.D.,
Giske, C.G., Irfan, S., Krishnan, P.,
Kumar, A.V., Maharjan, S., Mushtag, S.,
Noorie, T., Paterson, D.L., Pearson,
13
14
1 The Problem of Microbial Drug Resistance
11.
12.
13.
14.
15.
16.
17.
18.
19.
A., Perry, C., Pike, R., Rao, B., Ray, U.,
Sarma, J.B., Sharma, M., Sheridan, E.,
Thirunarayan, A., Turton, J., Upadhyay,
S., Warner, M., Welfere, W., Livermoer,
D.M., and Woodford, N. (2010) Emergence of a new antibiotic resistance
mechanism in India, Pakistan, and in
the UK: a molecular, biological, and
epidemiological study. Lancet Infect., 10,
597–602.
Cars, O., Hedin, A., and Heddini, A.
(2011) The global need for effective
antibiotics – moving towards concerted
action. Drug Resist. Updat., 14, 68–69.
Vergidis, P.I. and Falagas, M.E. (2008)
New antibiotics agents in bloodstream
infections. Int. J. Antimicrob. Agents, 32,
60–65.
Wright, G.D. (2005) Bacterial resistance
to antibiotics: enzymatic degradation
and modification. Adv. Drug Delivery
Rev., 57, 1451–1470.
French, G.L. (2010) The continuing
crisis in antibiotic resistance. Int. J.
Antimicrob. Agents, 36, 3–7.
Ippolito, G., Leone, S., Lauria, F.N.,
Nicastri, E., and Wenzel, R.P. (2010)
Methicillin – resistant Staphylococcus
aureus: the superbug. Int. J. Infect. Dis.,
14, 7–11.
Livermore, D. (2007) Introduction:
the challenge of multiresistance. Int. J.
Antimicrob. Agents, 29, 1–7.
Costelloe, C., Metcalfe, C., Lovering, A.,
Mant, D., and Hay, A.D. (2010) Effect
of antibiotic prescribing in primary
care on antimicrobial resistance in individual patients: systematic review and
meta-analysis. Br. Med. J., 340, 1–11.
Magiorakos, A.P., Srinivasan, A.,
Carey, R.B., Carmeli, Y., Falagas, M.E.,
Giske, C.G., Harbarth, S., Hindler,
J.F., Kahlmeter, G., Olsson-Liljequist,
B., Paterson, D.L., Rice, L.B., Stelling,
M.J., Vatopoulos, A., Weber, J.T., and
Monnet, D.L. (2012) Multidrug-resistant,
extensively drug-resistant and pandrugresistant bacteria: an international expert
proposal for interim standard definitions
for acquired resistance. Clin. Microbiol.
Infect., 18, 268–281.
WHO Antimicrobial Resistance. Fact
sheet N∘ 194, Updated May 2013,
http://www.who.int/mediacentre/
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
factsheets/fs194/en/ (accessed 3 February
2014).
Tana, M.M. and Ghany, M.G. (2013)
Hepatitis B virus treatment: management
of antiviral drug resistance. Clin. Liver
Dis., 2, 24–28.
Freire-Moran, L., Aronsson, B., Manz,
C., Gyssens, I.C., So, A.D., Monnet, D.L.,
and Cars, O. (2011) Critical shortage of
new antibiotics in development against
multidrug-resistant. Bacteria – Time to
react is now. Drug Resist. Updat., 14,
118–124.
Spellberg, B., Guidos, R., Gilbert, D.,
Bradley, J., Boucher, H.W., Scheld, W.M.,
Bartlett, J.G., and Edwards, J. (2008) The
epidemic of antibiotic-resistant infections: a call to action for the medical
community from the Infectious Diseases
Society of America. Clin. Infect. Dis., 46,
155–164.
Lynch, J.B. (2013) Multidrug-resistant
tuberculosis. Med. Clin. North Am., 97,
553–579.
World Health Organization (2012) The
Evolving Threat of Antimicrobial Resistance: Options for Action, World Health
Organization, Geneva, ISBN: 978 92 4
150318 1. www.who.int (accessed 9 April
2014).
Andersen, S.E. and Knudsen, J.D. (2013)
A managed multidisciplinary programme
on multi-resistant Klebsiella pneumoniae
in a Danish university hospital. BMJ
Qual. Saf., 22, 907–915.
Kallen, A.J., Hidron, A.I., Patel, J., and
Srinvasan, A. (2010) Multidrug resistance among Gram-negative pathogens
that caused healthcare – associated
infections reported to the National
Safety Network, 2006-2008. Infect.
Control Hosp. Epidemiol., 31, 528–531.
Deurenberg, R.H. and Stobberingh, E.E.
(2008) The evolution of Staphylococcus
aureus. Infect. Genet. Evol., 8, 747–763.
Arias, C.A. and Murray, B.E. (2009)
Antibiotic-resistant bugs in the 21st
century – a clinical super-challenge. N.
Engl. J. Med., 360, 439–443.
Gould, I.M., David, M.Z., Esposito,
S., Garau, J., Lina, G., Mazzei, T., and
Peters, G. (2012) New insights into
methicillin-resistant Staphylococcus
aureus (MRSA) pathogenesis, treatment
References
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
and resistance. Int. J. Antimicrob. Agents,
39, 96–104.
Nikaido, H. and Pagès, J.M. (2012)
Broad-specificity efflux pumps and their
role in multidrug resistance of Gramnegative bacteria. FEMS Microbiol. Rev.,
36, 340–363.
Lambert, P.A. (2005) Bacterial resistance
to antibiotics: modified target sites. Adv.
Drug Delivery Rev., 57, 1471–1485.
Kumar, A. and Schweizer, H.P. (2005)
Bacterial resistance to antibiotics: active
efflux and reduced uptake. Adv. Drug
Delivery Rev., 57, 1486–1513.
Coyne, S., Rosenfeld, N., Lambert, T.,
Courvalin, P., and Perichon, B. (2010)
Overexpression of resistance-nodulationcell division pump AdeFGH cofers
multidrug resistance in Acinetobacter baumannii. Antimicrob. Agents
Chemother., 54, 4389–4393.
Looke, D.F.M., Gottlieb, T., Jones, C.A.,
and Paterson, D.L. (2013) Gram-negative
resistance: can we combat the coming of
a new “Red Plague”? Med. J. Aust., 198,
243–244.
Bonnet, R. (2004) Growing group of
extended-spectrum β-lactamases: the
CTX-M enzymes. Antimicrob. Agents
Chemother., 48, 1–14.
Lahiri, S.D., Mangani, S., Durand-Reville,
T., Benvenuti, M., De Luca, F., Sanyal,
G., and Docquier, J.D. (2013) Structural
insight into potent broad-spectrum
inhibition with reversible recyclization
mechanism: avibactam in complex with
CTX-M-15 and Pseudomonas aeruginosa
AmpC β-Lactamases. Antimicrob. Agents
Chemother., 57, 2496–2505.
Munoz, P., Fernandez-Cruz, A.,
Rodriges- Creixems, M., and Bouza,
E. (2008) Gram – negative bloodstream
infections. Int. J. Antimicrob. Agents, 32,
10–14.
Wang, H.H. and Schaffner, D.W. (2011)
Antibiotic resistance: how much do
we now and where do we go from
here? Appl. Environ. Microbiol., 77,
7093–7095.
Saidel-Odes, L. and Borer, A. (2014)
Limiting and controlling carbapenemresistant Klebsiella pneumoniae. Infect.
Drug Resist., 7, 9–14.
40. Sho, T., Muratani, T., Hamasuna,
41.
42.
43.
44.
45.
46.
47.
R., Yakushiji, H., Fujimoto, N., and
Matsumoto, T. (2013) The mechanism
of high-level carbapenem resistance
in Klebsiella pneumoniae: underlying
Ompk36-deficient strains represent a threat of emerging high-level
carbapenem-resistant K. pneumoniae
with IMP-1 β-lactamase production
in Japan. Microb. Drug Resist., 19,
274–281.
French, G.L. (2005) Clinical impact and
relevance of antibiotic resistance. Adv.
Drug Delivery Rev., 57, 1514–1527.
Rumbo, C., Gato, E., López, M.,
Ruiz de Alegría, C., Fernández-Cuenca,
F., Martínez-Martínez, L., Vila,
J., Pachón, J., Cisneros, J.M.,
Rodríguez-Baño, J., Pascual, A., Bou,
G., and Tomás, M. (2013) Contribution
of efflux pumps, porins, and {beta}lactamases to multidrug resistance in
clinical isolates of Acinetobacter baumannii. Antimicrob. Agents Chemother.,
57, 5247–5257.
Gould, I.M., Cauda, R., Esposito, S., and
Gudiol, F. (2011) Management of serious
methicillin – resistant Staphylococcus
aureus infections: what are the limits?
Int. J. Antimicrob. Agents, 37, 202–209.
Samson, M., Pizzorno, A., Abed, Y.,
and Boivin, G. (2013) Influenza virus
resistance to neuraminidase inhibitors.
Antiviral Res., 98, 174–185.
Rahamat-Langendoen, J., Lokate, M.,
Friedrich, A., and Niesters, H. (2013)
The value of realtime sequence based
information in surveillance of nosocomial viral infections. Antimicrob. Resist.
Infect. Control, 13, P180.
Sun, N., Li, D., Fonzi, W., Li, X., Zhang,
X., and Calderone, R. (2013) Multidrugresistant transporter mdr1p-mediated
uptake of a novel antifungal compound.
Antimicrob. Agents Chemother., 57,
5931–5939.
Pfaller, M.A., Diekema, D.J., Gibbs,
D.L., Newell, V.A., Nagy, E., Dobiasova,
S., Rinaldi, M., Barton, R., Veselov,
A., and the Global Antifungal Surveillance Group (2008) Candida krusei, a
multidrug-resistant opportunistic fungal pathogen: geographic and temporal
15
16
1 The Problem of Microbial Drug Resistance
48.
49.
50.
51.
52.
53.
54.
Short-course empiric antibiotic therapy
trends from the ARTEMIS DISK antifor patients with pulmonary infiltrates
fungal surveillance program. J. Clin.
in the intensive care unit. A proposed
Microbiol., 46, 515–521.
solution for indiscriminate antibiotic
Mishra, B., Leishangthem, G.D., Gill, K.,
prescription. Am. J. Respir. Crit. Care
Singh, A.K., Das, S., Singh, K., Xess, I.,
Med., 162, 505–511.
Dinda, A., Kapil, A., Patro, I.K., and Dey,
D. (2013) A novel antimicrobial pep55. Deliberato, R.O., Marra, A.R.,
tide derived from modified N-terminal
Rodrigues Sanches, P., Martino, M.D.V.,
domain of bovine lactoferrin: design,
dos Santos Ferreira, E., Pasternak, J.,
synthesis, activity against multidrugTavares Paes, A., Pinto, L.M., dos Santos,
resistant bacteria and Candida. Biochim.
O.F.P., and Edmond, M.B. (2013) Clinical
Biophys. Acta, 1828, 677–686.
and economic impact of procalcitonin to
Hohn, A., Schroeder, S., Gehrt, A.,
shorten antimicrobial therapy in septic
Bernhardt, K., Bein, B., Wegscheider, K.,
patients with proven bacterial infecand Hochreiter, M. (2013) Procalcitonintion in an intensive care setting. Diagn.
guided algorithm to reduce length of
Microbiol. Infect. Dis., 76, 266–271.
antibiotic therapy in patients with severe 56. Streicher, E.M., Muller, B., Chihoda, V.,
sepsis and septic shock. BMC Infect.
Mlambo, C., Tait, M., Pillay, M., Torlipp,
Dis., 13, 158.
A., Hoek, K.G.P., Sirgel, F., van Pittius,
N.C.G., van Helden, P.D., Victor, T.C.,
Daly, C.G. (2013) Dental note. Antimiand Warren, R.M. (2012) Emergence ot
crobial stewardship. Aust. Prescriber, 36,
treatment of multidrug resistant (MDR)
123.
and extensively drug-resistant (XRD)
Gould, I.M. (2009) Antibiotic resistance:
tuberculosis in South Africa. Infect.
the perfect storm. Int. J. Antimicrob.
Genet. Evol., 12, 686–694.
Agents, 34, 2–5.
Charani, E. and Holmes, A.H. (2013)
57. Nemeth, J., Ledergerber, B., Preiswerk,
Antimicrobial stewardship programmes:
B., Nobile, A., Karrer, S., Ruef, C., and
the need for wider engagement. BMJ
Kuster, S.P. (2012) Multidrug-resistant
Qual. Saf., 22, 885–887.
bacteria in travellers hospitalized aboard:
prevalence, characteristics, and influence
Rawlins, M., McKenzie, D., and Mar,
on clinical outcome. J. Hosp. Infect., 82,
C. (2013) Antimicrobial stewardship:
254–259.
what’s it all about? Aust. Prescriber, 36,
116–120.
Singh, N., Rogers, P., Atwood, C.W.,
Wagener, M.M., and Yu, V.L. (2000)
17
2
Conventional Antibiotics – Revitalized by New Agents
Anthony Coates and Yanmin Hu
2.1
Introduction
Antibiotic resistance develops to all antibiotics [1–3]. Over several decades, this
has led to the need to replace old antibiotics with new ones. Unfortunately, the
world has not produced antibiotics fast enough to cope with the emergence of
antibiotic resistance, particularly for Gram-negative bacteria [4]. Between the
1940s and 1970s, the “Golden era,” about 20 new classes of antibiotics were
produced, which led to more than 200 analogs. Since then, there have only been
three new classes marketed, none of which are for Gram-negatives [5]. Can we
recreate the golden era? In other words, can we make 20 new classes of antibiotics
that are active against highly resistant bacteria? There is much debate about
this. While new antibiotics against Gram-positive bacteria have been marketed
in recent years, the main problem is that resistant Gram-negative bacteria are
poorly served, with no new class being marketed for 40 years [5]. Furthermore,
new antibiotics that are effective against the carbapenem-resistant bacteria [6],
which express, for example, NDM-1 [7], are not being introduced into the market
in good time, and we are playing “Catch-up.”
Is there a way forward? On the one hand, if enough money was provided
by governments, perhaps in a similar way to the Marshall plan [8, 9], or the
Public Health Emergency Medical Countermeasures Enterprise, [10] which is a
public–private partnership of multiple agencies of the US Federal Government,
many more antibiotics might reach the market. This would need to be accompanied by global efforts by nonprofit organizations such as the Bill & Melinda
Gates Foundation, the Drugs for Neglected Diseases Initiative (a research and
development organization that develops new treatments for neglected diseases),
and Medicines for Malaria Venture, which is a public–private partnership with
the aim of providing affordable antimalarial drug discovery and development. In
addition, there would need to be changes in regulation, and encouragements for
industry, for example, the Generating Antibiotics Incentives Now (GAIN) Act
and the proposed Antibiotic Development to Advance Patient Treatment Act,
both in the United States (“ADAPT Act”) for a limited population antimicrobial
Novel Antimicrobial Agents and Strategies, First Edition.
Edited by David A. Phoenix, Frederick Harris and Sarah R. Dennison.
© 2015 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2015 by Wiley-VCH Verlag GmbH & Co. KGaA.
18
2 Conventional Antibiotics – Revitalized by New Agents
drug pathway [8, 11]. On the other hand, in the long term, it may not be possible to
market enough antibiotics to keep up with the relentless emergence of antibiotic
resistance [12, 13]. While prevention will clearly play a greater role, this will not
substitute for new antibiotics. The existing strategy is to discover and develop
novel single antibiotic therapy [14]. Do we need to rethink this strategy? Considering costs alone, if we intend to discover and develop 200 new antibiotics, the
cost will be somewhere in excess of $1 billion per compound [15]. So this route
would be very expensive. Is it scientifically feasible to endlessly produce more and
more antibiotics? The past 40 years have shown that it is becoming more difficult
to bring new antibiotics to the market. The absence of new classes of antibiotics
for Gram-negative infections during this period is an important example [5]. Now
we have virtually untreatable carbapenem-resistant Gram-negative infections,
exemplified by those bacteria that express metallo-beta-lactamase (MBL) [7],
for which we are using, as a last resort, colistin [16], itself an old, relatively
toxic antibiotic. Colistin resistance is now emerging in MBL Enterobacteriaceae
[17]. Furthermore, the development of novel antibiotics against MBL-resistant
bacteria is still in early clinical development [14, 18]. In addition, we know that
antibiotic resistance arises to all antibiotics within a few years after entry into the
marketplace [19]. Therefore, a continuous flow of new antibiotics into the market
is needed. It seems unlikely that the world will be able to produce a limitless
number of antibiotics far into the future. If the supply of effective antibiotics dries
up, modern medicine is likely to suffer a devastating setback [13, 20].
We propose a new strategy. The world should revitalize conventional antibiotics
by combining them with antibiotic resistance breakers (ARBs). This approach
would mean that we could, potentially, continue to use conventional antibiotics.
This has the advantage of being a cheaper option than developing hundreds of
new antibiotics. For example, if each class of antibiotics could be resuscitated by a
single ARB, theoretically, most of the 200 existing antibiotics could become useful
again. Potentially, this could be achieved with fewer new compounds than would
be required for the replacement of the existing 200 compounds. There would be
substantial financial savings and this would transform the feasibility of prolonging
the antibiotic era. This chapter looks at the origins of combination antibiotic
therapy and examines whether it is possible to extend this concept, namely, the
combination of conventional antibiotics (see Table 2.1) with resistance breakers,
thereby revitalizing a wide range of different classes of antibiotics.
2.2
Conventional Antibiotics
The main classes of antibiotics that have been marketed, and many of their
analogs, are listed in Table 2.1. Resistance has occurred to all of them. The
β-lactams are degraded by bacterial β-lactamases, which can be neutralized by
combining the old antibiotic with a β-lactamase inhibitor such as clavulanic
2.2
Conventional Antibiotics
Table 2.1 Main classes of antibiotics.
Class
Examples
Aminoglycosides
Streptomycin, neomycin, kanamycin, paromycin,
gentamicin, tobramycin, amikacin, netilmicin,
spectinomycin, sisomicin, dibekalin, isepamicin
𝛃-Lactams
Penicillins
Cephalosporins
First generation
Second generation
Third generation
Fourth generation
Carbapenems
Monobactams
𝛃-lactamase
inhibitors
Glycopeptides
Macrolides
Metronidazole
Lincosamides
Lipopeptides
Oxazolidinones
Polymyxin
Quinolines
Quinolones
Rifamycins
Streptogramins
Sulfonamides
Tetracyclines
Trimethoprim
Penicillin G, penicillin V, methicillin, oxacillin, cloxacillin,
dicloxacillin, nafcillin, ampicillin, amoxicillin, carbenicillin,
ticarcillin, mezlocillin, piperacillin, azlocillin, temocillin
Cepalothin, cephapirin, cephradine, cephaloridine,
cefazolin
Cefamandole, cefuroxime, cephalexin, cefprozil, cefaclor,
loracarbef, cefoxitin, cefmetazole
Cefotaxime, ceftizoxime, ceftriaxone, cefoperazone,
ceftazidime, cefixime, cefpodoxime, ceftibuten, cefdinir
Cefpirome, cefepime
Imipenem, meropenem
Astreonam
Clavulanate, sulbactam, tazobactam
Vancomycin, teicoplanin
Erythromycin, azithromycin, clarithromycin
—
Lincomycin, clindamycin
Daptomycin
Linezolid
Polymyxin B, Polymyxin E (colistin)
Bedaquiline
Nalidixic acid, oxolinic acid, norfloxacin, pefloxacin,
enoxacin, ofloxacin/levofloxacin, ciprofloxacin,
temafloxacin, lomefloxacin, fleroxacin, grepafloxacin,
sparfloxacin, trovafloxacin, clinafloxacin, gatifloxacin,
moxifloxacin, sitafloxacin
Rifampicin (also called rifampin), rifapentine, rifabutin,
bezoxazinorifamycin, rifaximin
Quinupristin, daflopristin
Sulfanilamide, para-aminobenzoic acid, sulfadiazine,
sulfisoxazole, sulfamethoxazole, sulfathalidine
Tetracycline, chlortetracycline, demeclocycline,
minocycline, oxytetracycline, methacycline, doxycycline,
tigecycline
—
19
20
2 Conventional Antibiotics – Revitalized by New Agents
acid [21]. This is discussed in more detail in the next section, as are other
combinations. Potential combinations only exist for a minority of classes.
2.3
The Principles of Combination Antibiotic Therapy
In the clinic, combinations of antibiotics are often used. The main reasons for such
combinations are as follows:
1) Combinations break resistance and rejuvenate old antibiotics. The best
example of this approach is the combination of clavulanic acid and amoxicillin [21]. Clavulanic acid inhibits bacterial β-lactamase, which neutralizes
amoxicillin, thus allowing the latter to kill β-lactamase-producing bacteria.
Clavulanic acid alone has no antibacterial activity. This chapter primarily
deals with breaking resistance.
2) Combinations prevent the emergence of resistance during chemotherapy. It is
important to appreciate the limitations of this approach. While combinations
of antibiotics do prevent the emergence of resistance during tuberculosis
chemotherapy [22], it is unlikely that this will be effective in multispecies
environments such as the large intestine. In the case of Mycobacterium
tuberculosis, combinations are effective because mutations only arise in
the chromosome, and do not occur because of plasmid transfer from other
species of bacteria [23]. M. tuberculosis lives on its own in a relatively sterile
environment, for example, inside macrophages in the lung. So, there is little
opportunity for plasmid transfer. Resistance due to transfer of plasmids does
not occur in M. tuberculosis. In contrast, other bacteria, such as Escherichia
coli, live in the large intestine in a multispecies environment where resistance
is often transferred via plasmids [24]. Combinations such as sulfonamide and
trimethoprim (co-trimoxazole) already have high levels of resistance – for
example, over 95% of Gram-negative bacteria from babies in rural India [25]
in spite of early hopes that such a combination would prevent the emergence
of resistance [26]. A meta-analysis (including data from eight randomized
controlled trials) that compared aminoglycoside/β-lactam combination therapy with β-lactam monotherapy to observe the emergence of antimicrobial
resistance found that aminoglycoside/β-lactam combination therapy was not
associated with a reduced development of resistance when compared with
β-lactam therapy alone [27]. Nevertheless, for certain infections where chromosomal resistance is thought to be important, combinations of different
antibiotics may have the potential to prevent the emergence of resistance.
3) Combinations in which one antibiotic boosts the effect of a second antibiotic
and vice versa. This is called synergy. For instance, penicillin and gentamicin
are synergistic [28], and are used to treat bacterial endocarditis.
4) A combination of antibiotics is used by clinicians to broaden the number of
species of bacteria that are targeted. For example, if a seriously ill patient has
2.4
Antibiotic Resistance Breakers: Revitalize Conventional Antibiotics
suspected intra-abdominal infection with an unknown bacterium, an aminoglycoside and anti-anerobe agents can be used [29].
5) Sometimes, the clinician may be faced with an infection that harbors dormant bacteria as well as fast multiplying ones. Tuberculosis is well known as
an infection that persists owing to the presence of dormant bacteria that are
relatively tolerant to antibiotics. Combinations of antibiotics, typically containing four separate compounds (rifampicin, pyrazinamide, isoniazid, and
ethambutol), are used in the initial stages of tuberculosis therapy. Rifampicin
and pyrazinamide kill dormant bacteria and so are responsible for the shortening of the duration of chemotherapy from 12 to 6 months [22].
2.4
Antibiotic Resistance Breakers: Revitalize Conventional Antibiotics
The main threat to the effectiveness of a marketed antibiotic is the emergence of
widespread resistance among its bacterial targets. While prevention of resistance
is clearly the ultimate answer to this problem, the world is a long way from reversing this trend. Since resistance to an antibiotic is an inevitable consequence of
entry into the market, the main subject of this chapter is to examine the feasibility
of revitalizing conventional antibiotics by the addition of an ARB . The combination is active against resistant bacteria. In the large pyogenic bacterial field,
combination therapy has not been developed to the extent that it has in tuberculosis, although HIV and cancer therapy do use well-characterized combinations of
drugs. There are a number of ways that conventional antibiotics can be revitalized
by combining them with another agent.
2.4.1
𝛃-Lactamase Inhibitors
Bacteria can produce β-lactamases, which are enzymes that destroy the β-lactam
ring of β-lactam antibiotics, thereby reducing their effectiveness [30]. There are
over 1300 known β-lactamases. The concept of combining a β-lactam antibiotic
with a β-lactamase inhibitor in order to revitalize the antibiotic and to render
it active against β-lactamase-expressing bacteria, was first introduced into the
market by a combination of the β-lactamase inhibitor clavulanic acid, derived
from Streptomyces clavuligerus, with amoxicillin [31]. This combination is called
Augmentin (GlaxoSmithKline, Brentford, UK). In a clinical trial [32], patients with
non-bullous impetigo were treated with either amoxicillin alone or Augmentin.
The causative organism of impetigo, Staphylococcus aureus, was shown to be
present in lesions from all the patients. When tested for sensitivity to amoxicillin,
all the bacterial isolates were resistant, but were sensitive to Augmentin. Clinically, the Augmentin group of patients responded better than the amoxicillin
group. These data indicated that neutralization of bacterial β-lactamase can
revitalize amoxicillin.
21
22
2 Conventional Antibiotics – Revitalized by New Agents
Unfortunately, bacteria produce many β-lactamases that are not inhibited by
clavulanic acid. There has been a 100-fold increase in the number of known
β-lactamase inhibitors in the past 40 years [30]. The classification of bacterial
β-lactamases is complicated. We have used the Bush [30] system in this chapter,
bearing in mind that extended spectrum beta-lactamases (ESBLs, which include
TEM and SHV) and carbapenemases (such as NDM and KPC) in Gram-negatives
are thought to be of the greatest clinical importance because they are difficult to
treat and are relatively common in many countries [33, 34]. β-lactamases can be
divided into Ser- and MBLs, by their active sites. They are subdivided into molecular classes A–D, which have functional groups and major functional subgroups.
For example, the serine β-lactamases of molecular class C 1(1, 1e), which degrade
early cephalosporins, and expanded spectrum cephalosporins in the case of 1e,
Class A 2 (2a, 2b, 2be, 2br, 2f ), which degrade penicillins and others, and in the
case of 2f, penicillins, early and expanded spectrum cephalosporins, carbapenems
and monobactams, and Class D 2d (2de, 2df ), which destroy penicillins and in
the case of 2df, carbapenems. The MBLs B 3 (3a and 3b) target carbapenems, and
in the case of 3a, penicillins and early and expanded spectrum cephalosporins.
Enzymes that are expressed are C 1 (AmpC, CMY) and 1e (GC1), A 2a (PC1),
2b (TEM-1, SHV-1), 2be (CTX-M, ESBLs (TEM, SHV)), 2br (IRT, SHV-10), 2f
(KPC,SME), 2de (OXA-11, OXA-15), 2df (OXA-23, OXA-48), and B 3a (IMP,
VIM, NDM), 3b (CphA).
Clavulanic acid only neutralizes the serine β-lactamases A (2a, 2b, and 2be) and
has a partial effect on A (2f ), and D (2d). Clavulanic acid has also been combined
with ticarcillin (Timentin; GlaxoSmithKline, Brentford, UK). Other inhibitor
combinations include tazobactam with piperacillin (Zosyn; Pfizer, Philadelphia,
PA, USA), and sulbactam with ampicillin (Unasyn; Pfizer, Philadelphia, PA, USA).
Unfortunately, the current β-lactamase inhibitor combinations are not active
against bacteria that express AmpC or ESBLs. Even worse is that, so far, it is
proving difficult to develop MBL inhibitors that are effective against NDM [35].
Since the current marketed inhibitors are only active against class A enzymes
but lack effectiveness against class A KPC carbapenemases, new inhibitors are
under development, which broaden the β-lactamases that can be neutralized.
For example, avibactam, which is a bridged 1,6-diazabicyclo[3.2.1]ocatan-7-one
(DBO), is in clinical development. This compound is active against a wide range
of Class A and C serine β-lactamases [36], including ESBLs and class A carbapenemases. Although it neutralizes Class D OXA-48, it is inactive against other
D carbapenemases. This molecule also inhibits selected class D β-lactamases
including OXA-48, but not other class D carbapenemases or B MBLs. Avibactam
combinations with ceftaroline (Cereza-Forest) and cefdazidime (AstraZeneca
and Forest) are in clinical trials [35]. Another combination under development
(Cubist) is tazobactam and ceftolozane [37]. Tazobactam increases the activity of the combination against ESBL-producing Enterobacteriaceae, and can
partially neutralize AmpC and KPC carbapenemases. A new DBO (MK-7655
Merck) has been combined with imipenem and is in clinical trials [38]. This
combination is active against KPC-2-producing Klebsiella pneumoniae and
2.4
Antibiotic Resistance Breakers: Revitalize Conventional Antibiotics
AmpC-overexpressing isolates of Pseudomonas aeruginosa but not against those
that express metallo-carbapenemases [39].
2.4.2
Aminoglycoside-Modifying Enzyme Inhibitors
While these types of inhibitors have not yet reached the clinical trials phase of
development, some interesting in vitro experience has been achieved. In general,
inhibitors of aminoglycoside-modifying enzymes [40, 41] have struggled with
numerous different targets because bacteria may express multiple enzymes. However, inhibition of aminoglycoside phosphotransferases and acetyl transferases
has been shown by cationic antimicrobial peptides (AMPs) [40]. Indolicidin is
a bovine AMP. This peptide and its synthetic analogs inhibited both aminoglycoside phosphotransferase and aminoglycoside acetyltransferase classes. This is
the first description of broad-spectrum inhibitors of aminoglycoside resistance
enzymes. Crystallographic studies have shed light on the molecular structure
of aminoglycosidephosphotransferases or kinases (APHs). A review of APH
structures and inhibitors is covered by Shi and colleagues [42]. These data suggest
that the commercial development of a universal APH inhibitor may not be
feasible.
2.4.3
Antibiotic Efflux Pumps Inhibitors
Although there are numerous examples of antibiotic efflux pump inhibitors, none
are in clinical trials as yet. The main families of bacterial efflux pumps that are
chromosomally expressed and that are associated with multidrug resistance [43]
are the resistance nodulation division (RND) family (encodes AcrA/B-TolC), the
major facilitator superfamily (MFS) (encodes QacA), the staphylococcal multiresistance (SMR) (encodes QacC), the multidrug and toxic compound extrusion
(MATE) family (NorM), and the ATP binding cassette (ABC) (LmrA). Efflux pump
inhibitors include reserpine [44], which is too neurotoxic to be used at effective
concentrations in humans [45], berberine and palmatine [46], and other compounds (reviewed in [43]) including plant extracts, synthetic molecules, thioxanthenes, phenothiazenes, and arylpiperazines. While some inhibitors perform
well in vitro, problems with toxicity have not resulted in extensive clinical trials. In
addition, particularly in some Gram-negative bacteria, treatment with an inhibitor
may lead to compensatory upregulation of other efflux pumps. For example, [47],
RamA expression is induced by inhibition of efflux or inactivation of acrAB in
Salmonella typhimurium.
2.4.4
Synergy Associated with Bacterial Membrane Permeators
Synergy between non-antibiotics and antibiotics and between antibiotics themselves is well known. In some cases, this synergy is associated with one of the
23
24
2 Conventional Antibiotics – Revitalized by New Agents
pair in the combination being a bacterial membrane permeabilizer. Whether this
is responsible for the synergy is unknown in many cases, but it has been suggested [48] that permeabilization of the membrane may increase the intracellular concentration of the antibiotic in the combination, and this, in turn may
increase the antibacterial potency of the antibiotic. Some of these associations
are described here.
Gram-negative bacteria have two membranes. In the case of fluoroquinolones,
outer membrane proteins play a key part in helping these molecules to cross the
membrane [49, 50]. In contrast, passive diffusion is thought to be important for
translocation of the inner membrane of Gram-negatives and the single membrane
of Gram-positives [51–54]. In the 1960s [55, 56], improved penetration of fluoroquinolones was achieved by the addition of a 7-piperazine side chain and this
is thought to initiate translocation across the membrane [57]. This suggests that
adding side groups such as piperazine or membrane permeabilization compounds
in combinations could be a way of increasing the activity of current antibiotics.
One of the most serious problems in clinical practice in the world is the
emergence of carbapenem-resistant Gram-negative bacteria. Carbapenems are
often used as the antibiotics of last resort. Combinations of antibiotics are used
to treat patients with carbapenem-resistant MBL producing Gram-negative
infections such as K. pneumoniae [58], and these combinations often contain
colistin. This antibiotic, which is a polypeptide of the polymyxin group, increases
the permeability of Gram-negative membranes [59]. The polycationic regions of
colistin displace the bacterial counter ions in the lipopolysaccharide of the outer
membrane. The inner membrane is solubilized by the hydrophobic/hydrophilic
regions of colistin. While clinical data regarding the efficacy of different antibiotic
combinations is sparse, in vitro data [58] suggests that a combination of colistin,
rifampicin, and meropenem is effective against MBL-producing K. pneumoniae
(VIM; NDM-1).
AMPs can also increase the permeability of bacterial membranes, and can synergize with conventional antibiotics. For example, [60] AMPs have been created,
which synergize with conventional antibiotics such as cefotaxime, ciprofloxacin,
or erythromycin against highly resistant strains of the Gram-negative bacterium
Acinetobacter baumannii. There are three models of AMP membrane interaction
(reviewed in [61]): Barrel-stave pores, toroidal pores, and carpet mechanism, in
which peptides form a layer on the surface and dissolve the membrane [62]. AMPs
have numerous other effects on bacterial cells, and so synergy may not necessarily
be the most important as far as a bactericidal effect is concerned.
A recent development has been the observation of enhancement or synergy
between a compound that was developed against dormant S. aureus [63] and three
different classes of antimicrobials [64]. The compound (HT61; Helperby Therapeutics Ltd., London) depolarizes the bacterial cell membrane and is in clinical
trials. Another example is loperamide (Immodium; McNeil Consumer Healthcare, Fort Washington, PA, USA) [65], an opioid receptor agonist, which enhances
the activity of minocycline against E. coli, S. aureus, and P. aeruginosa. Loperamide
2.5
Discussion
interferes with the electrical component of the proton motive force of the bacterial membrane. This leads to an increase in the pH gradient, which enhances the
entry of tetracycline into the cell.
2.5
Discussion
Revitalizing old antibiotics by combination with a second compound means that
resistance to the old antibiotic is broken by the second compound, either directly
or indirectly. There is only one clear, clinically proven example of rejuvenation
of old antibiotics in this way, namely, the addition of β-lactamase inhibitors
to β-lactams. Arguably, the addition of the 7-piperazine ring to quinolones in
order to enhance the initiation of translocation could be regarded as another
example. Antibiotic–antibiotic combinations that are frequently used in clinical
practice, for example, in tuberculosis chemotherapy, do not break resistance as
such. Such antibiotic–antibiotic combinations (with the exception of those that
include colistin, and perhaps other membrane permeators) have other functions
such as preventing the emergence of resistance (tuberculosis chemotherapy)
or synergy (increasing efficacy). If resistance exists to the primary antibiotic, a
second antibiotic is added to which the organism is sensitive and this renders the
combination effective. Combinations can also broaden the spectrum of species
that are targeted. For example, in abdominal sepsis patients, two antibiotics such
as an aminoglycoside and anti-anaerobe agents are used together to cover as
many aerobic and anaerobic species of bacteria as possible before the results of
microbiological tests are available. Some combinations contain drugs that kill
dormant organisms (for instance, pyrazinamide and rifampicin in tuberculosis
chemotherapy), thus shortening the duration of therapy.
The advantages of revitalizing old antibiotics, such as β-lactams with a
β-lactamase inhibitor, is that the existing antibiotic can be used once again to
effectively treat a resistant bacterial infection that was previously untreatable
by that antibiotic. A further advantage of this approach is that it is relatively
low cost because one ARB can be used to rejuvenate several old antibiotics. In
addition, the risk that is associated with this approach is lower than that with
developing a novel antimicrobial because once the ARB has been shown to
be safe in clinical trials in combination with one compound it can be used to
rejuvenate other old antibiotics. Furthermore, instead of reproducing the golden
era of antibiotic discovery by creating 200 novel antibiotics, the world could,
potentially, rejuvenate existing antibiotics with 20 or less ARBs in combination
with 200 existing antibiotics.
Could ARBs prevent the emergence of resistance? While combinations of
drugs are used in tuberculosis, HIV, and cancer chemotherapy to reduce the
emergence of resistance, there are certain fundamental differences between these
combinations and ARBs for the treatment of pyogenic bacterial infections such
as urinary tract disease due to Gram-negative bacteria. The first difference is that
M. tuberculosis resistance is not transmitted by plasmids. It is chromosomally
25
26
2 Conventional Antibiotics – Revitalized by New Agents
mediated. This contrasts with resistance in pyogenic bacteria that is transmitted
by plasmids in some cases, is chromosomally mediated in others, and through
both mechanisms in some. It is unlikely that ARBs could reduce the emergence of
plasmid-mediated resistance, but they might be able to impact on chromosomal
resistance. A second important difference is that some ARBs, such as some
β-lactamase inhibitors, have no antibacterial activity by themselves. These ARBs
are unlikely to be able to prevent even chromosomal resistance because resistance
emergence is effectively appearing to the one old antibiotic alone. If, however,
the ARB has some antibacterial activity in its own right, such as HT61 [63],
mutants that arise to the old antibiotic can be killed by the ARB, and thus the
combination may be able to prevent the emergence of chromosomally mediated
resistance. Could ARBs be used to reduce the dose of old antibiotics against
sensitive bacterial strains, and so decrease the incidence of toxic side effects?
If the ARB can boost the effect of the old antibiotic against sensitive strains, it
may be possible to use a lower dose of the old antibiotic to achieve cure. Would
ARBs enhance activity against dormant bacteria? This depends on the ARB.
β-lactamase inhibitors have no action against dormant bacteria and so would not
increase a β-lactam’s activity against dormant bacteria. In contrast, other ARBs
such as HT61, which was selected for anti-dormancy activity [63, 64], boost the
activity of the combinations against dormant bacteria.
Historically, resistance has eventually emerged to every antibiotic after entry
into the market. Clearly, resistance will appear to ARB combinations. Experience
with β-lactamase inhibitors suggests that mutant bacteria emerge over time
that express β-lactamases, such as the B3a MBL NDM, that are resistant to, for
example, clavulanic acid [7]. Since bacteria produce over 1000 β-lactamases, it
seems likely that, when challenged with a new β-lactamase inhibitor, mutants
will emerge that can neutralize the inhibitor with a novel β-lactamase. Ways
need to be found to slow down the emergence of resistance. One possible route
could be to use ARBs that target the cell membrane, on the grounds that it may
take bacteria longer to develop resistance against combinations that act on the
bacterial membrane [66]. ARBs that can rescue old antibiotics from a wide range
of resistance challenges are needed, and those that can counteract MBLs are
urgently needed.
Acknowledgments
We would like to acknowledge financial support from the Burton Trustees and
Helperby Therapeutics Ltd.
References
1. Wellington, E.M., Boxall, A.B.,
Cross, P., Feil, E.J., Gaze, W.H.,
Hawkey, P.M., Johnson-Rollings, A.S.,
Jones, D.L., Lee, N.M., Otten, W.,
Thomas, C.M., and Williams, A.P.
(2013) The role of the natural
environment in the emergence of
antibiotic resistance in Gram-negative
bacteria. Lancet Infect. Dis., 13 (2),
155–165.
References
2. Arias, C.A. and Murray, B.E. (2009)
3.
4.
5.
6.
7.
8.
9.
10.
Antibiotic-resistant bugs in the 21st
century–a clinical super-challenge.
N. Engl. J. Med., 360, 439–443.
Livermore, D.M. (2009) Has the era
of untreatable infections arrived?
J. Antimicrob. Chemother., 64, 29–36.
Cantón, R., Akóva, M., Carmeli,
Y., Giske, C.G., Glupczynski, Y.,
Gniadkowski, M., Livermore, D.M.,
Miriagou, V., Naas, T., Rossolini, G.M.,
Samuelsen, Ø., Seifert, H., Woodford,
N., Nordmann, P., and European Network on Carbapenemases (2012) Rapid
evolution and spread of carbapenemases
among Enterobacteriaceae in Europe.
Clin. Microbiol. Infect., 18 (5), 413–431.
Coates, A.R., Halls, G., and Hu, Y. (2011)
Novel classes of antibiotics or more of
the same? Br. J. Pharmacol., 163 (1),
184–194.
Nordmann, P. (2013) Carbapenemaseproducing Enterobacteriaceae: overview
of a major public health challenge. Med.
Mal. Infect., pii, S0399-077X(13)00336-3,
doi: 10.1016/j.medmal.2013.11.007.
Bushnell, G., Mitrani-Gold, F., and
Mundy, L.M. (2013) Emergence of
New Delhi metallo-β-lactamase type
1-producing enterobacteriaceae and nonenterobacteriaceae: global case detection
and bacterial surveillance. Int. J. Infect.
Dis., 17 (5), e325–e333.
Laxminarayan, R., Duse, A., Wattal, C.,
Zaidi, A.K., Wertheim, H.F., Sumpradit,
N., Vlieghe, E., Hara, G.L., Gould, I.M.,
Goossens, H., Greko, C., So, A.D.,
Bigdeli, M., Tomson, G., Woodhouse,
W., Ombaka, E., Peralta, A.Q., Qamar,
F.N., Mir, F., Kariuki, S., Bhutta, Z.A.,
Coates, A., Bergstrom, R., Wright,
G.D., Brown, E.D., and Cars, O. (2013)
Antibiotic resistance-the need for global
solutions. Lancet Infect. Dis., 13 (12),
1057–1098.
The Marshall Plan (1948)
http://www.marshallfoundation.org/
TheMarshallPlan.htm (accessed 2 March
2014).
U.S. Department of Health and
Human Services (2010) The Public Health Emergency Medical
Countermeasures Enterprise 2010,
https://www.medicalcountermeasures.gov/
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
media/1138/mcmreviewfinalcover-508.pdf
(accessed 2 March 2014).
Antibiotic Development to Advance
Patient Treatment Act (2013)
http://usamobiledrugtesting.com/adaptact-would-create-a-new-limitedpopulation-approval-pathway-forantibiotics-and-antifungals (accessed
3 March 2014).
Grossi, P.A., Tebini, A., and
Dalla Gasperina, D. (2014) Novel
multi-drug resistant microorganisms in
critically ill: a potential threat. Minerva
Anestesiol. [Epub ahead of print].
Davies, S. (2013) The Drugs Don’t Work
(Penguin Special), Penguin.
Coates, A.R. and Halls, G. (2012) Antibiotics in phase II and III clinical trials.
Handb. Exp. Pharmacol., 211, 167–183.
Huttner, A., Harbarth, S., Carlet, J.,
Cosgrove, S., Goossens, H., Holmes,
A., Jarlier, V., Voss, A., and Pittet,
D. (2013) Antimicrobial resistance:
a global view from the 2013 World
Healthcare-Associated Infections Forum.
Antimicrob. Resist. Infect. Control, 2 (1),
31.
Tzouvelekis, L.S., Markogiannakis,
A., Psichogiou, M., Tassios, P.T., and
Daikos, G.L. (2012) Carbapenemases in
Klebsiella pneumoniae and other Enterobacteriaceae: an evolving crisis of global
dimensions. Clin. Microbiol. Rev., 25 (4),
682–707.
Tascini, C., Tagliaferri, E., Giani, T.,
Leonildi, A., Flammini, S., Casini, B.,
Lewis, R., Ferranti, S., Rossolini, G.M.,
and Menichetti, F. (2013) Synergistic
activity of colistin plus rifampin against
colistin-resistant KPC-producing Klebsiella pneumoniae. Antimicrob. Agents
Chemother., 57 (8), 3990–3993.
Chen, J., Shang, X., Hu, F., Lao, X., Gao,
X., Zheng, H., and Yao, W. (2013) βLactamase inhibitors: an update. Mini
Rev. Med. Chem., 13 (13), 1846–1861.
(1998) House of Lords inquiry into
antimicrobial resistance. Commun. Dis.
Rep. CDR Wkly., 8 (17), 147, 150.
So, A.D., Gupta, N., and Cars, O. (2010)
Tackling antibiotic resistance. Br. Med. J.,
340, c2071 (Editorial).
Watkins, R.R., Papp-Wallace, K.M.,
Drawz, S.M., and Bonomo, R.A. (2013)
27
28
2 Conventional Antibiotics – Revitalized by New Agents
22.
23.
24.
25.
26.
27.
28.
29.
30.
Novel β-lactamase inhibitors: a therapeutic hope against the scourge of
multidrug resistance. Front. Microbiol., 4,
392.
Mitchison, D. and Davies, G. (2012)
The chemotherapy of tuberculosis: past,
present and future. Int. J. Tuberc. Lung
Dis., 16 (6), 724–732.
Mitchison, D.A. (2012) Prevention
of drug resistance by combined drug
treatment of tuberculosis. Handb. Exp.
Pharmacol., 211, 87–98.
Dalhoff, A. (2012) Resistance
surveillance studies: a multifaceted
problem–the fluoroquinolone example.
Infection, 40 (3), 239–262.
Viswanathan, R., Singh, A.K., Ghosh, C.,
Dasgupta, S., Mukherjee, S., and Basu, S.
(2012) Profile of neonatal septicaemia at
a district-level sick newborn care unit.
J. Health Popul. Nutr., 30 (1), 41–48.
Grüneberg, R.N. (1979) The microbiological rationale for the combination
of sulphonamides with trimethoprim.
J. Antimicrob. Chemother., 5 (Suppl. B),
27–36.
Bliziotis, I.A., Samonis, G., Vardakas,
K.Z., Chrysanthopoulou, S., and Falagas,
M.E. (2005) Effect of aminoglycoside
and beta-lactam combination therapy
versus beta-lactam monotherapy on the
emergence of antimicrobial resistance: a
meta-analysis of randomized, controlled
trials. Clin. Infect. Dis., 41 (2), 149–158.
Indrelie, J.A., Wilson, W.R., Matsumoto,
J.Y., Geraci, J.E., and Washington, J.A.
II, (1984) Synergy of imipenem or penicillin G and aminoglycosides against
enterococci isolated from patients with
infective endocarditis. Antimicrob.
Agents Chemother., 26 (6), 909–912.
Investigators of the
Piperacillin/Tazobactam Intra-abdominal
Infection Study Group (1994) Results
of the North American trial of
piperacillin/tazobactam compared
with clindamycin and gentamicin in
the treatment of severe intra-abdominal
infections. Eur. J. Surg. Suppl., 573,
61–66.
Bush, K. (2013) Proliferation and significance of clinically relevant β-lactamases.
Ann. N. Y. Acad. Sci., 1277, 84–90.
31. Brogden, R.N., Carmine, A., Heel, R.C.,
32.
33.
34.
35.
36.
37.
38.
39.
40.
Morley, P.A., Speight, T.M., and Avery,
G.S. (1981) Amoxycillin/clavulanic acid:
a review of its antibacterial activity,
pharmacokinetics and therapeutic use.
Drugs, 22 (5), 337–362. Review.
Dagan, R. and Bar-David, Y. (1989)
Comparison of amoxicillin and clavulanic acid (augmentin) for the treatment
of nonbullous impetigo. Am. J. Dis.
Child., 143 (8), 916–918.
Savard, P. and Perl, T.M. (2012) A call
for action: managing the emergence of
multidrug-resistant Enterobacteriaceae
in the acute care settings. Curr. Opin.
Infect. Dis., 25, 371–377.
Lascols, C., Hackel, M., Hujer, A.M.,
Marshall, S.H., Bouchillon, S.K., Hoban,
D.J., Hawser, S.P., Badal, R.E., and
Bonomo, R.A. (2012) Using nucleic
acid microarraysto perform molecular epidemiology and detect novel
beta-lactamases: a snapshot of extendedspectrum betalactamases throughout the
world. J. Clin. Microbiol., 50, 1632–1639.
Buynak, J.D. (2013) Beta-Lactamase
inhibitors: a review of the patent literature 2010–2013. Expert Opin. Ther. Pat.,
23 (11), 1–13.
Drawz, S.M., Papp-Wallace, K.M., and
Bonomo, R.A. (2013) New β-lactamase
inhibitors: a therapeutic renaissance in
an “MDR world”. Antimicrob. Agents
Chemother., [Epub ahead of print].
Hong, M.C., Hsu, D.I., and Bounthavong,
M. (2013) Ceftolozane/tazobactam: a
novel antipseudomonal cephalosporin
and β-lactamase-inhibitor combination. Infect. Drug Resist., 6, 215–223.
(eCollection 2013. Review).
Shlaes, D.M. (2013) New β-lactam-βlactamase inhibitor combinations in
clinical development. Ann. N. Y. Acad.
Sci., 1277, 105–114.
Livermore, D.M., Warner, M., and
Mushtaq, S. (2013) Activity of MK7655 combined with imipenem against
Enterobacteriaceae and Pseudomonas
aeruginosa. J. Antimicrob. Chemother., 68
(10), 2286–2290.
Boehr, D.D., Draker, K.A., Koteva, K.,
Bains, M., Hancock, R.E., and Wright,
G.D. (2003) Broad spectrum peptide
inhibitors of aminoglycoside antibiotic
References
41.
42.
43.
44.
45.
46.
47.
48.
49.
resistance enzymes. Chem. Biol., 10 (2),
189–196.
Singh, H., Thangaraj, P., and Chakrabarti,
A. (2013) Acinetobacter baumannii: a
brief account of mechanisms of multidrug resistance and current and future
therapeutic management. J. Clin. Diagn.
Res., 7 (11), 2602–2605.
Shi, K., Caldwell, S.J., Fong, D.H., and
Berghuis, A.M. (2013) Prospects for
circumventing aminoglycoside kinase
mediated antibiotic resistance. Front.
Cell. Infect. Microbiol., 3, 22.
Piddock, L.J. (2006) Clinically relevant
chromosomally encoded multidrug resistance efflux pumps in bacteria. Clin.
Microbiol. Rev., 19 (2), 382–402.
Neyfakh, A.A., Bidnenko, V.E., and
Bo Chen, L. (1991) Efflux-mediated
multidrug resistance in Bacillus subtilis:
similarities and dissimilarities with the
mammalian system. Proc. Natl. Acad.
Sci. U.S.A., 88, 4781–4785.
Markham, P.N. and Neyfakh, A. (1996)
Inhibition of the multi-drug transporter
norA prevents emergence of norfloxacin
resistance in Staphylococcus aureus.
Antimicrob. Agents Chemother., 40,
2673–2674.
Hsieh, P.-C., Siegel, S.A., Rogers, B.,
Davis, D., and Lewis, K. (1998) Bacteria
lacking a multidrug pump: a sensitive
tool for drug discovery. Proc. Natl. Acad.
Sci. U.S.A., 95, 6602–6606.
Lawler, A.J., Ricci, V., Busby, S.J., and
Piddock, L.J. (2013) Genetic inactivation
of acrAB or inhibition of efflux induces
expression of ramA. J. Antimicrob.
Chemother., 68 (7), 1551–1557.
Buyck, J.M., Plésiat, P., Traore, H.,
Vanderbist, F., Tulkens, P.M., and
Van Bambeke, F. (2012) Increased susceptibility of Pseudomonas aeruginosa
to macrolides and ketolides in eukaryotic cell culture media and biological
fluids due to decreased expression of
oprM and increased outer-membrane
permeability. Clin. Infect. Dis., 55 (4),
534–542.
Masi, M. and Pagès, J.M. (2013) Structure, function and regulation of outer
membrane proteins involved in drug
transport in Enterobactericeae: the
50.
51.
52.
53.
54.
55.
56.
57.
58.
OmpF/C - TolC case. Open Microbiol. J.,
7, 22–33.
Tran, Q.T., Williams, S., Farid, R.,
Erdemli, G., and Pearlstein, R. (2013)
The translocation kinetics of antibiotics
through porin OmpC: insights from
structure-based solvation mapping using
WaterMap. Proteins, 81 (2), 291–299.
Bayer, A.S., Schneider, T., and Sahl, H.G.
(2013) Mechanisms of daptomycin resistance in Staphylococcus aureus: role of
the cell membrane and cell wall. Ann. N.
Y. Acad. Sci., 1277, 139–158.
Weingart, H., Petrescu, M., and
Winterhalter, M. (2008) Biophysical
characterization of in- and efflux in
Gram-negative bacteria. Curr. Drug
Targets, 9 (9), 789–796.
Nikaido, H. and Thanassi, D.G. (1993)
Penetration of lipophilic agents with
multiple protonation sites into bacterial
cells: tetracyclines and fluoroquinolones
as examples. Antimicrob. Agents
Chemother., 37, 1393–1399.
Vázquez, J.L., Merino, S., Domenech, O.,
Berlanga, M., Viñas, M., Montero, M.T.,
and Hernández-Borrell, J. (2001) Determination of the partition coefficients of
a homologous series of ciprofloxacin:
influence of the n − 4 piperazinyl alkylation on the antimicrobial activity. Int. J.
Pharm., 220, 53–62.
Domagala, J.M. (1994) Structure-activity
and structure-side-effect relationships for
the quinolone antibacterials. J. Antimicrob. Chemother., 33 (4), 685–706.
Sanchez, J.P., Domagala, J.M., Hagen,
S.E., Heifetz, C.L., Hutt, M.P., Nichols,
J.B. et al. (1988) Quinolone antibacterial
agents. Synthesis and structure-activity
relationships of 8-substituted quinolone3-carboxylic acids. J. Med. Chem., 31,
983–991.
Cramariuc, O., Rog, T., Javanainen, M.,
Monticelli, L., Polishchuk, A.V., and
Vattulainen, I. (2012) Mechanism for
translocation of fluoroquinolones across
lipid membranes. Biochim. Biophys.
Acta, 1818 (11), 2563–2571.
Tängdén, T., Hickman, R.A., Forsberg,
P., Lagerbäck, P., Giske, C.G., and Cars,
O. (2014) Evaluation of double and triple
antibiotic combinations for VIM- and
29
30
2 Conventional Antibiotics – Revitalized by New Agents
59.
60.
61.
62.
NDM-producing Klebsiella pneumoniae by in vitro time-kill experiments.
Antimicrob. Agents Chemother., 58 (3),
1757–1762.
Dhariwal, A.K. and Tullu, M.S. (2013)
Colistin: re-emergence of the ‘forgotten’
antimicrobial agent. J. Postgrad. Med., 59
(3), 208–215.
Gopal, R., Kim, Y.G., Lee, J.H., Lee,
S.K., Chae, J.D., Son, B.K., Seo, C.H.,
and Park, Y. (2013) Synergistic effects
and anti-biofilm properties of chimeric
peptides against MDR Acinetobacter
baumannii strains. Antimicrob. Agents
Chemother., [Epub ahead of print].
Tavares, L.S., Silva, C.S., de Souza, V.C.,
da Silva, V.L., Diniz, C.G., and Santos,
M.O. (2013) Strategies and molecular
tools to fight antimicrobial resistance:
resistome, transcriptome, and antimicrobial peptides. Front. Microbiol., 4, 412.
eCollection 2013.
Pietiäinen, M., François, P., Hyyryläinen,
H.L., Tangomo, M., Sass, V., Sahl, H.G.,
Schrenzel, J., and Kontinen, V.P. (2009)
Transcriptome analysis of the responses
of Staphylococcus aureus to antimicrobial peptides and characterization of
63.
64.
65.
66.
the roles of vraDE and vraSR in antimicrobial resistance. BMC Genomics, 10,
429.
Hu, Y., Shamaei-Tousi, A., Liu, Y., and
Coates, A. (2010) A new approach for
the discovery of antibiotics by targeting
non-multiplying bacteria: a novel topical
antibiotic for staphylococcal infections.
PLoS One, 5 (7), e11818.
Hu, Y. and Coates, A.R. (2013) Enhancement by novel anti-methicillin-resistant
Staphylococcus aureus compound HT61
of the activity of neomycin, gentamicin, mupirocin and chlorhexidine: in
vitro and in vivo studies. J. Antimicrob.
Chemother., 68 (2), 374–384.
Ejim, L., Farha, M.A., Falconer, S.B.,
Wildenhain, J., Coombes, B.K., Tyers,
M., Brown, E.D., and Wright, G.D.
(2011) Combinations of antibiotics and
nonantibiotic drugs enhance antimicrobial efficacy. Nat. Chem. Biol., 7,
348–350.
Hurdle, J.G., O’Neill, A.J., Chopra, I.,
and Lee, R.E. (2011) Targeting bacterial
membrane function: an underexploited
mechanism for treating persistent infections. Nat. Rev. Microbiol., 9 (1), 62–75.
31
3
Developing Novel Bacterial Targets: Carbonic Anhydrases as
Antibacterial Drug Targets
Clemente Capasso and Claudiu T. Supuran
3.1
Introduction
Metals play roles in approximately one-third of the known enzymes. Proteins having metals incorporated into their molecule are known as metalloenzymes [1, 2].
Many enzymes incorporate electrophilic or nucleophilic moieties, whose reactivity is enhanced by the presence of cations, explaining why metal ions are frequently
present in enzymes. In most metalloenzymes, the role of the cation is one of the
following three:
1) Within the active site of the protein, participating in the catalytic mechanism
of the enzyme
2) Into noncatalytic sites, stabilizing the structure of the enzyme and contributing to its tridimensional folding
3) Into the heme, within a prosthetic group as for hemoglobins and cytochromes
[1–4]. Metal removal from the molecule leads to the unfolding of the tridimensional structure and frequently to the inactivation of the enzyme [3, 4].
Intriguingly, metalloenzymes are incredibly different between them, regarding
the nature of the catalytic/structural cation present within the active site, the number of cations within the active site/the whole protein, the oligomeric state of
the enzyme, and so on, being involved in a multitude of important physiological processes [5]. Carbonic anhydrases (CAs), carboxypeptidases, hemoglobins,
cytochromes, phosphotransferases, alcohol dehydrogenase, arginase, ferredoxin,
and cytochrome oxidase are several examples of the most common metalloenzymes known to date [3, 4].
3.2
Carbonic Anhydrases
In this chapter, the reader’s attention is focused on the role of one of the crucial metalloenzymes in all life kingdoms, carbonic anhydrase (CA, EC 4.2.1.1).
Developing CA inhibitor-based antibiotics by inhibiting bacterial CAs present
Novel Antimicrobial Agents and Strategies, First Edition.
Edited by David A. Phoenix, Frederick Harris and Sarah R. Dennison.
© 2015 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2015 by Wiley-VCH Verlag GmbH & Co. KGaA.
32
3 Developing Novel Bacterial Targets: Carbonic Anhydrases as Antibacterial Drug Targets
in pathogenic species started to be investigated relatively recently as a new drug
design strategy for obtaining antibacterial agents [5–9].
CAs are ubiquitous metalloenzymes present in prokaryotes and eukaryotes. The
five genetically distinct classes known to date, the α-, β-, γ-, δ-, and ζ-CAs, are all
metalloenzymes, using Zn(II), Cd(II), or Fe(II) at their active sites [9–15]. The
α-CAs are present in vertebrates, bacteria, algae, and cytoplasm of green plants;
the β-CAs are predominantly found in bacteria, algae, and chloroplasts of both
mono and dicotyledons; the γ-CAs are present in Archaea, bacteria, and plants,
whereas the δ- and ζ CAs are present in marine diatoms and other eukaryotes
composing the plankton [16, 17]. These enzymes catalyze a very simple physiological reaction, the interconversion between carbon dioxide and the bicarbonate
ion when a proton is also formed, being involved in crucial physiological processes
connected with respiration and transport of CO2 /bicarbonate between metabolizing tissues and lungs (in vertebrates), pH and CO2 homeostasis, electrolyte
secretion in a variety of tissues/organs, biosynthetic reactions (such as gluconeogenesis, lipogenesis, and ureagenesis – again in vertebrates), bone resorption, calcification, and tumorigenicity [7, 12, 16, 18]. In prokaryotes, the existence of genes
encoding for CAs from at least three classes (α-, β-, and γ-class) suggests that
these enzymes play an important role in prokaryotic physiology. CAs, in fact, are
involved in the transport of CO2 or HCO3 − , in supplying CO2 or HCO3 − for
the biosynthetic reactions (and thus metabolisms); in pH regulation and also in
xenobiotic degradation (for example, cyanate, by Escherichia coli), as well as in
the survival of intracellular pathogens within their host [7, 12, 16, 18].
At present, infectious diseases are the second leading cause of death in the world
and the emergence of antibiotic-resistant bacteria is an inevitable and dangerous
phenomenon, inherent to most new drugs [19, 20]. The possibility of developing
new antibacterial agents raised much interest recently. The main classes of antibiotics clinically used nowadays act toward the inhibition of four classical targets: (i)
cell wall biosynthesis; (ii) protein biosynthesis; (iii) DNA and RNA biosynthesis;
and (iv) folate biosynthesis [21]. CAs started to be investigated in detail recently
in pathogenic bacteria, in the search for antibiotics with a novel mechanism of
action, as it has been demonstrated that in many bacteria CAs are essential for
the life cycle of the organism and that their inhibition leads to growth impairment
or growth defects of the pathogen [22–26].
3.3
CA Inhibitors
Several classes of CA inhibitors (CAIs) are known to date: the metal complexing
anions, and the unsubstituted sulfonamides, which bind to the Zn(II) ion of
the enzyme either by substituting the non-protein zinc ligand or by adding to
the metal coordination sphere, generating trigonal-bipyramidal species are the
classical, most frequently investigated ones [26–29]. The primary sulfonamides
were the first antimicrobial drugs, discovered in 1935 by Domagk, and they paved
3.4
Classes of CAs Present in Bacteria
the way for the antibiotic revolution in medicine [30]. The first sulfonamide
showing effective antibacterial activity, prontosil, was a prodrug, with the real
antibacterial agent being sulfanilamide, a compound isosteric/isostructural
with 4-aminobenzoic acid (pABA). Sulfanilamide is generated by the in vivo
reduction of prontosil. In the years following the discovery of sulfanilamide
as a bacteriostatic agent, a range of analogs have entered into clinical use
(constituting the so-called sulfa drug class of antibacterials), and many of these
compounds are still widely used [7]. Sulfonamides, such as the clinically used
derivatives acetazolamide (AAZ), methazolamide (MZA), ethoxzolamide (EZA),
dichlorophenamide, dorzolamide (DZA), and brinzolamide (BRZ), bind in a
tetrahedral geometry to the Zn(II) ion in a deprotonated state, with the nitrogen
atom of the sulfonamide moiety coordinated to Zn(II) and an extended network of
hydrogen bonds, involving amino acid residues Thr199 and Glu106 (numbering
system used for the human CA, isoform I), also participating in the anchoring of
the inhibitor molecule to the metal ion [7]. The aromatic/heterocyclic part of the
inhibitor interacts with the hydrophilic and hydrophobic residues of the cavity.
Anions, such as the inorganic metal-complexing ones, or more complicated
species such as the carboxylates, are also known to bind to the CAs, but generally
with less efficiency compared to the sulfonamides [8, 24, 27]. Anions may bind
either with the tetrahedral geometry of the metal ion or as trigonal-bipyramidal
adducts. Enzymes found in vertebrates, arthropods, corals, fungi, bacteria,
diatoms, and Archaea have been investigated for their inhibition with simple
inorganic anions. Anion inhibitors are important both for understanding the
inhibition/catalytic mechanisms of these enzymes fundamental for many physiologic processes, and for designing novel types of inhibitors that may have clinical
applications for the management of a variety of disorders in which CAs are
involved [9–14, 17, 26].
3.4
Classes of CAs Present in Bacteria
The bacterial genomes encode CAs belonging to at least three classes (the α-, β-,
and γ-CA class). The metal coordination pattern in the α and γ-CAs involves three
histidine residues and a water molecule, while the amino acid residues involved
in the catalytic cycle of β-CAs are two cysteines and one histidine, the fourth
metal ion ligand being again a water molecule/hydroxide ion. The existence of
five CA classes (α, β, γ, δ, and ζ) that have evolved independently has been well
documented in the literature, these enzymes being an excellent example of convergent evolution at the molecular level [13, 31–36]. Hence, we present here the
most parsimonious phylogenetic tree of the CAs, in order to elucidate the phylogenetic relationships existing among the three classes found in bacteria, the
α, β, and γ-CAs. Phylogenetic analysis was carried out using α-, β-, and γ-CAs
from different prokaryotic and eukaryotic species (Figure 3.1). From the dendrogram shown in Figure 3.1, the α-CAs (bacterial and non-bacterial) appear closely
33
34
3 Developing Novel Bacterial Targets: Carbonic Anhydrases as Antibacterial Drug Targets
HumCAII_alpha
HumCAI_alpha
HpyICA_alpha
0.115
SspCA_alpha
0.158
VchCA_alpha
0.723
SsalCA_alpha
0.856
NgonCA_alpha
0.925 PseCA_gamma
0.996
BglCA_gamma
0.253
CAM_gamma
0.468
CreCA_gamma
AthCA_gamma
PgiCA_gamma
PgiCA_beta
0.877
0.875
MinCA_beta
AbaCA_beta
0.936
0.99
0.889
0.885
CA class
Organism
Alpha
Helicobacter pylori J99
Homo sapiens, isoform II
Homo sapiens, isoform I
Sulfurihydrogenibium yellowstonense YO3AOP1
Streptococcus salivarius PS4
Vibrio cholerae
Neisseria gonorrhoeae
NP_223829.1
AAH11949.1
NP_001158302.1
ACD66216.1
EIC81445.1
AFC59768.1
CAA72038.1
HpyICA_alpha
HumCAII_alpha
humCAI_alpha
SspCA_alpha
SsalCA_alpha
VchCA_alpha
NgonCA_alpha
Beta
Schizosaccharomyces pombe
Brucella suis 1330
Burkholderia thailandensis Bt4
Coccomyxa sp.
Chlamydomonas reinhardtii
Acinetobacter baumannii
Porphyromonas gingivalis
Myroides injenensis
Zea mays
Vigna radiata
Flaveria bidentis, isoform I
Arabidopsis thaliana
Helicobacter pylori
Legionella pneumophila 2300/99
Escherichia coli
Methanobacterium thermoautotrophicum
Saccharomyces cerevisiae
Dekkera bruxellensis AWRI1499
CAA21790
NP_699962.1
ZP_02386321
AAC33484.1
XP_001699151.1
YP_002326524
YP_001929649.1
ZP_10784819
NP_001147846.1
AAD27876
AAA86939.2
AAA50156
BAF34127.1
YP_003619232
ACI70660
GI:13786688
GAA26059
EIF49256
SpoCA_beta
BsuCA_beta
BthCA_beta
CspCA_beta
CreCA_beta
AbaCA_beta
PgiCA_beta
MinCA_beta
ZmaCA_beta
VraCA_beta
FbiCA_beta
AthCA_beta
HpyCA_beta
LpnCA_beta
EcoCa_beta
Cab_beta
SceCA_beta
DbrCA_beta
Gamma
Pseudomonas sp. PAMC 25886
Burkholderia gladioli BSR3
Methanosarcina thermophila
Chlamydomonas reinhardtii
Arabidopsis thaliana
Porphyromonas gingivalis
ZP_10427314.1
YP_004359911.1
ACQ57353.1
XP_001703237.1
NP_564091.1
YP_001929649.1
PseCA_gamma
BgICA_gamma
CAM_gamma
CreCA_gamma
AthCA_gamma
PgiCA_gamma
Cab_beta
0.889
ZmaCA_beta
VraCA_beta
FbiCA_beta
AthCA_beta
HpyCA_beta
LpnCA_beta
EcoCA_beta
SpoCA_beta
BsuCA_beta
0.965
BthCA_beta
0.858
CspCA_beta
0.968
CreCA_beta
SceCA_beta
DbrCA_beta
0.997
0.968
0.507
0.248
0.861
0.496
0
0.963
0.955
0.133
Accession number Cryptonym
Figure 3.1 Phylogenetic analysis carried out on the α, β, and γ-CAs identified in the genome of different prokaryotic and eukaryotic organisms.
All the CA classes, microorganisms, accession numbers, and cryptonyms have been indicated in the table included in this figure.
3.6
α-CAs in Pathogenic Bacteria
related to the γ-CAs clustering in a branch distinct from that of the β-CAs. These
results probably indicate that there was a gene duplication event occurring in the
ancestral lineage of CAs separating the α- and γ-class enzymes from the β-CAs.
3.5
Pathogenic Bacterial CAs
Microbes express their pathogenicity by means of their virulence, which determines the ability of pathogens to enter a host, evade host defenses, grow in the host
environment, counteract its immune responses, assimilate iron or other nutrients
from the host, or sense environmental changes [37–39]. Numerous enzymes are
responsible for the virulence, acting against host components and contributing to
the damage of host tissues.
Cloning of the genomes of many pathogenic microorganisms offered the possibility of exploring alternative pathways for inhibiting virulence factors or proteins essential for their life cycle, and such an approach was applied systematically
for CAs from pathogenic bacteria in the last decade. As described above, bacterial genomes are characterized by the presence of genes encoding for α-, β-, and
γ-CAs. Hence, the investigations of these three classes of CAs may reveal novel
aspects of microbial virulence. CAs started to be investigated in detail recently
in pathogenic bacteria, in the search for antibiotics with a novel mechanism of
action, as it has been demonstrated that in many of them, these enzymes are essential for the life cycle of the pathogen [7, 16].
3.6
𝛂-CAs in Pathogenic Bacteria
The α-CAs are present in vertebrates, protozoa, algae, and cytoplasm of green
plants and in bacteria [7, 16]. The first bacterial α-CA was characterized from
the pathogen Neisseria gonorrhoeae [40], a Gram-negative coffee-bean-shaped
diplococcic bacteria responsible for the sexually transmitted infection gonorrhea.
The α-CA identified in N. gonorrhoeae (NgCA) However, the three-dimensional
structures of these two contains 252 amino acid residues, has a molecular mass
of 28 kDa and, presumably, is localized in the periplasm of the bacterial cell. Its
amino acid sequence is only about 35% identical with that of human (h) hCA
II. enzymes are quite similar, although several surface loops are considerably
shorter in the bacterial enzyme than in the hCA II [40]. One important difference
between these two α-CAs is the presence of a disulfide bond in bacterial enzyme,
while the cytoplasmic hCA II contains only one Cys residue. The bacterial enzyme
showed a high CO2 hydrase activity, with a k cat of 1.1 × 106 s−1 and K m of 20 mM
(at pH 9 and 25 ∘ C). The enzyme also showed esterase activity for the hydrolysis of
4-nitrophenyl acetate, similarly to the mammalian isoforms hCA I and II. NgCA
crystallized as a dimer. Recently, our groups resolved the three-dimensional
35
36
3 Developing Novel Bacterial Targets: Carbonic Anhydrases as Antibacterial Drug Targets
structure of the first thermostable α-CA identified in the thermophilic bacteria
Sulfurihydrogenibium yellowstonense. The extremo-α-CA was named SspCA and,
intriguingly, this molecule had an asymmetric crystallographic unit formed by a
dimer with two independent moieties. This is a very interesting result considering
that α-CAs were long considered to be monomeric enzymes [9, 11, 36, 41].
The best studied bacterial α-CA is that from the gastric pathogen provoking
ulcer and gastric cancer, Helicobacter pylori, named hpαCA. This CA has a
periplasmic localization and was shown to be catalytically efficient with almost
identical activity to that of the human isoform hCA I, for the CO2 hydration
reaction, and highly inhibited by many sulfonamides/sulfamates, including AAZ,
EZA, topiramate (TPM), and sulpiride (SLP), all clinically used drugs [42–44].
Furthermore, some CAIs, such as AAZ and MZA, were shown to inhibit the
bacterial growth in cell cultures. Since the genome of H. pylori encodes also for
a cytoplasmic β-CA and the efficacy of H. pylori eradication therapies currently
employed has been decreasing because of drug resistance and side effects of the
commonly used drugs, the dual inhibition of α- and/or β-CAs of H. pylori could
be applied as an alternative therapy in patients with H. pylori infection or for the
prevention of gastroduodenal diseases due to this widespread pathogen [42–46].
In 2005, in fact, Shahidzadeh and coworkers [47] showed the efficacy of AAZ in
the treatment of gastric ulcer. This compound (as well as EZA) was in fact widely
used as an antiulcer agent in the 1970s and 1980s [47].
Recently, our group cloned, purified, and characterized an α-CA from the
human pathogenic bacterium Vibrio cholerae, VchCA [31, 48, 49]. This enzyme
showed significant catalytic activity for CO2 hydration reaction. VchCA has
kinetic parameters quite similar to those of the human isoform hCA I, with
a k cat of 8.23 × 105 s−1 and a K m of 11.7 mM, which leads to a k cat /K m of
7.0 × 107 M−1 s−1 (compared to 5.0 × 107 M−1 s−1 for hCA I) [31]. VchCA is thus
slightly more active than hCA I but also four times more active compared to
the bacterial enzyme, hpαCA. VchCA was about half as active as hCA II, one
of the best catalysts known in nature (which has a k cat /K m of 1.5 × 108 M−1 s−1 )
[31]. The inhibition study with sulfonamides and sulfamates led to the detection
of a large number of low nanomolar inhibitors, among which are MZA, AAZ,
EZA, DZA, BRZ, benzolamide (BZA), and indisulam (IND) (K I values in the
range 0.69–8.1 nM). For example, VchCA was strongly inhibited by AAZ, a
sulfonamide in clinical use, with a K I of 6.8 nM, and was almost twice more
sensitive to this inhibitor than hCA II and three times more sensitive than the
H. pylori enzyme. Since bicarbonate was demonstrated to be a virulence factor
of this bacterium and since EZA was shown to inhibit the in vivo virulence, we
proposed that VchCA may be a target for antibiotic development, exploiting a
mechanism of action rarely considered until now [31]. Many inorganic anions
and several small molecules were also investigated as VchCA inhibitors [49].
Inorganic anions such as cyanate, cyanide, hydrogen sulfide, hydrogen sulfite,
and trithiocarbonate were effective VchCA inhibitors with inhibition constants in
the range of 33–88 μM. Other effective inhibitors were diethyldithiocarbamate,
sulfamide, sulfamate, phenylboronic acid, and phenylarsonic acid, with K I s of
3.7
β-CAs in Pathogenic Bacteria
7–43 μM. Halides (bromide, iodide), bicarbonate, and carbonate were much
less effective VchCA inhibitors, with K Is in the range of 4.64–28.0 mM. The
resistance of VchCA to bicarbonate inhibition may represent an evolutionary
adaptation of this enzyme to living in an environment rich in this ion, such as the
gastrointestinal tract, as bicarbonate is a virulence enhancer of this bacterium,
as mentioned above. These findings are thus of extreme importance because
they may help in engineering highly active α-CAs such as hCA II in order to
obtain mutated enzymes with enhanced thermostability for use under the harsh
conditions of the biotechnological processes. For example, the temperature of
combustion gases or liquids into which the CO2 are dissolved may easily exceed
the optimal temperature for the enzyme used in CO2 capture process.
3.7
𝛃-CAs in Pathogenic Bacteria
β-CAs are found in bacteria, algae, and chloroplasts of both mono- and dicotyledons, and also in many fungi and some Archaea. CAs belonging to the β-class were
cloned, purified, and characterized in many pathogenic bacteria, such as E. coli,
Mycobacterium tuberculosis, Salmonella enterica, H. pylori, Haemophilus influenzae, and Streptococcus pneumoniae [16, 29]. A list of the β-CAs isolated from
pathogens is presented in Table 3.1. The main difference between these enzymes
and the α-CAs discussed above consists in the fact that usually the β-CAs are
oligomers, generally formed of two to six monomers of molecular weight (of the
monomer) of 25–30 kDa [35, 45, 50, 51]. The tridimensional structures of these
enzymes are similar to that of the S. enterica isoform stCA 1, characterized by a
long channel at the bottom of which the catalytic zinc ion is found, tetrahedrally
coordinated by Cys42, Asp44, His98, and Cys101 (numbering system of stCA1).
This is called closed active site as these enzymes are not catalytically active (at pH
Table 3.1 β-CAs from bacteria cloned and characterized so far.
Pathogen
Helicobacter pylori
Escherichia coli
Haemophilus influenzae
Mycobacterium tuberculosis
Brucella suis
Streptococcus pneumoniae
Salmonella enterica
Enzyme abbreviation
hp𝛃CA
EcoCA
HICA
mtCA1
mtCA2
mtCA3
mtCA4
bsCA1
bsCA2
PCA
stCA1
stCA2
37
38
3 Developing Novel Bacterial Targets: Carbonic Anhydrases as Antibacterial Drug Targets
values <8.3). At pH values >8.3, the “closed active site” is converted to the “open
active site,” associated with a movement of the Asp residue from the catalytic
Zn(II) ion, with the concomitant coordination of an incoming water molecule
approaching the metal ion [35, 45, 50, 51]. This water molecule (as hydroxide ion)
is in fact responsible for the catalytic activity, as for the α-CAs investigated in much
greater detail. Many of these CAs reported in Table 3.1 displayed excellent activity
for the CO2 hydration but lacked esterase activity, similar to the β-class enzymes
isolated from other organisms (plants, arthropods, etc.). For some of these CAs
in vitro and in vivo inhibition studies with various classes of inhibitors, such as
anions, sulfonamides, and sulfamates, have been reported and are discussed in
the following paragraphs.
As mentioned above, the genome of H. pylori encodes for two different classes of
CAs, with different subcellular localization: a periplasmic α-class CA (hpαCA) and
a cytoplasmic β-class CA (hpβCA). These two CAs were shown to be catalytically
efficient with almost identical activity to that of the human isoform hCA I, for the
CO2 hydration reaction, and highly inhibited by many sulfonamides/sulfamates,
including clinically used drugs [42, 43, 45]. In vivo, it was possible to observe inhibition of the bacterial growth for H. pylori, S. pneumoniae, Brucella suis, and
M. tuberculosis. In fact, most of the sulfonamides possess a rather polar nature
and may have problems passing through the cell membranes of many pathogenic
bacteria. This was the case for M. tuberculosis possessing four β-CAs, which have
good catalytic activity and are inhibited in the low nanomolar range by many
sulfonamides [52, 53]. All in vivo inhibition studies with such CAIs only gave negative results (fort the inhibition of growth of the pathogen), but recently, Colina’s
group showed that a class of phenol inhibitors possesses antimycobacterial activity in vitro and in vivo [52]. Unlike the sulfonamides, these derivatives were more
lipophilic and probably are able to better penetrate through the bacterial cell walls.
Thus, many of these β-CAs are in fact validated drug targets.
Our group obtained in vitro inhibition data for several of these enzymes
with sulfonamide/sulfamates, which represent one of the main classes of CAIs.
These compounds are clinically used drugs, for example, AAZ, MZA, EZA,
dichorophenamide DCP, DZA, BRZ, BZA, TPM, zonisamide ZNS, SLP, IND,
celecoxib CLX, valdecoxib VLX, as diuretics, antiepileptics, antiglaucoma,
and anti-inflammatory agents. It was observed that most CAs from bacterial
pathogenic organisms are inhibited in the micro-nanomolar range by many
such sulfonamide/sulfamate drugs [6, 35, 50, 54]. It is interesting to note that
the research in this area may lead to highly effective and bacterial CA-selective
compounds, which may validate these enzymes as antibacterial drug targets.
Recently, Ferry and coworkers [55, 56] purified and characterized kinetically
a β-CA in the pathogen Clostridium perfringens (named CpeCA). C. perfringens
is the most common bacterial agent for gas gangrene, which is necrosis, putrefaction of tissues, and gas production. The gas gangrene is caused primarily by
C. perfringens alpha toxin. Besides, C. perfringens is also the third most common
cause of food poisoning in the United Kingdom and the United States though it
can sometimes be ingested and cause no harm. The study of CAs in C. perfringens
3.8
γ-CAs from Pathogenic Bacteria
provides the basis for developing better clostridial enzyme inhibitors with potential as anti-infectives with a new mechanism of action. Vullo et al. [55] described
the first inhibition study of this β-CA. CpeCA was poorly inhibited by iodide
and bromide, and was inhibited with KIs in the range of 1–10 mM by a range
of anions such as (thio)cyanate, azide, bicarbonate, nitrate, nitrite, hydrogensulfite, hydrogensulfide, stannate, tellurate, pyrophosphate, divanadate, tetraborate,
peroxydisulfate, sulfate, iminodisulfonate, and fluorosulfonate. Better inhibition,
with K Is of 0.36–1.0 mM, was observed for cyanide, carbonate, selenate, selenocyanide, trithiocarbonate, and diethyldithiocarbamate, whereas the best CpeCA
inhibitors were sulfamate, sulfamide, phenylboronic acid, and phenylarsonic acid,
which had K Is in the range of 7–75 μM.
3.8
𝛄-CAs from Pathogenic Bacteria
The γ-CAs were found in Archaea, bacteria, and plants. The prototype of
the γ-class CAs, “Cam,” has been isolated from the methanogenic archaeon
Methanosarcina thermophila [57–59]. The crystal structures of zinc-containing
and cobalt-substituted Cam were reported in the unbound form and cocrystallized with sulfate or bicarbonate. Cam has several features that differentiate
it from the α- and β-CAs. Thus, the protein fold is composed of a left-handed
β-helix motif interrupted by three protruding loops and followed by short
and long α-helices. Cam monomer self-associates in a homotrimer with the
approximate molecular weight of 70 kDa. The Zn(II) ion within the active site
is coordinated by three histidine residues, as in α-CAs, but relative to the
tetrahedral coordination geometry seen at the active site of α-CAs; the active site
of this γ-CA contains additional metal-bound water ligands, so that the overall
coordination geometry is trigonal-bipyramidal for the zinc-containing Cam and
octahedral for the cobalt-substituted enzyme [60, 61]. Two of the His residues
coordinating the metal ion belong to one monomer (monomer A), whereas the
third one is from the adjacent monomer (monomer B). Thus, the three active
sites are located at the interface between pairs of monomers. The catalytic
mechanism of γ-CAs was proposed to be similar to the one presented for the
α-class enzymes. Still, the finding that Zn(II) is not tetracoordinated as originally
reported but pentacoordinated, with two water molecules bound to the metal
ion, demonstrates that much is still to be understood regarding these enzymes. At
this moment, the zinc hydroxide mechanism is accepted as being valid for γ-CAs,
as it is probable that an equilibrium exists between the trigonal-bipyramidal
and the tetrahedral species of the metal ion from the active site of the enzyme
[60, 61].
Our groups identified a γ-CA (denominated PgiCA) in the genome of the
pathogenic bacterium Porphyromonas gingivalis, which is a Gram-negative oral
anaerobe involved in the pathogenesis of periodontitis and arthritis [62, 63].
39
40
3 Developing Novel Bacterial Targets: Carbonic Anhydrases as Antibacterial Drug Targets
These are inflammatory diseases, which destroy the tissues supporting the
tooth/bone, eventually leading to tooth loss as well as other serious conditions.
More than 700 bacterial species colonize the oral cavity, but P. gingivalis is
the species mainly associated with the chronic form of periodontitis. The open
reading frame of the P. gingivalis gene encodes a 192 amino acid polypeptide
chain, which displays 33 and 30% identity when compared with the prototypical
γ-CAs CAM and CAMH, respectively. CAM and CAMH both belong to the
γ-CA class and were isolated from the archaeon M. thermophila; they have
been thoroughly characterized by Ferry’s group, as discussed above. PgiCA was
shown to possess a significant catalytic activity for the reaction that converts
the CO2 to bicarbonate and protons, with a k cat of 4.1 × 105 s−1 and a k cat /K m of
5.4 × 107 M−1 s−1 , thus being 62 times more effective as a catalyst compared to
CAM (k cat /K m of 8.7 × 105 M−1 s−1 ) [62, 63]. Like most enzymes belonging to
the CA superfamily, PgiCA was also inhibited by AAZ with inhibition constants
of 324 nM whereas CAM was inhibited in the low nanomolar range (K I s of
63 nM). We have also investigated the inhibition profile of the new enzyme with
a range of inorganic anions such as thiocyanate, cyanide, azide, hydrogen sulfide,
sulfamate, and trithiocarbonate. These anions were effective PgiCA inhibitors
with inhibition constants in the range of 41–97 μM [62, 63]. Other effective
inhibitors were diethyldithiocarbamate, sulfamide, and phenylboronic acid, with
K Is of 4.0–9.8 μM [62, 63]. The role of this enzyme as a possible virulence factor
of P. gingivalis is poorly understood at the moment but its good catalytic activity
and the possibility to be inhibited by a large number of compounds may lead to
interesting developments in the field.
3.9
Conclusions
The extensive use of antibiotics has led to serious public health problems due to the
emergence of multiresistant bacterial pathogens worldwide. The recent cloning
and characterization of many CAs in pathogenic bacteria and the proof of concept studies showing that these are potential drug targets and can lead to growth
inhibition in these bacteria offer interesting new alternatives that were not yet
fully exploited clinically. In many pathogenic bacteria, α-, β-, and γ-CAs have been
cloned and characterized in detail in the last years. For some of these enzymes,
the X-ray crystal structures were determined at high resolution, allowing a good
understanding of the catalytic/inhibition mechanisms. Here, we have highlighted
the existence of a great number of low nanomolar/micromolar CAIs targeting
some of them, and belonging to various chemical classes (sulfonamides, sulfamates, sulfamides, carboxylates, phenols, etc.). The inhibitors targeting the α-, β-,
and γ-CAs from many pathogenic bacteria may provide opportunities to identify
novel antibacterial targets for the development of alternative classes of antibiotics
and to design more potent antimicrobial compounds derived from the existing
References
antibiotics in clinical use for decades. This represents a fascinating research field
that can lead to interesting developments in the antibacterial drug research.
References
1. McCall, K.A., Huang, C., and Fierke,
2.
3.
4.
5.
6.
7.
8.
9.
C.A. (2000) Function and mechanism
of zinc metalloenzymes. J. Nutr., 130,
1437S–1446S.
Reetz, M.T. (2012) Artificial metalloenzymes as catalysts in stereoselective
Diels-Alder reactions. Chem. Rec., 12,
391–406.
Vallee, B.L. and Auld, D.S. (1992) Active
zinc binding sites of zinc metalloenzymes. Matrix, 1, 5–19.
Feinberg, H., Greenblatt, H.M., and
Shoham, G. (1993) Structural studies
of the role of the active site metal in
metalloenzymes. J. Chem. Inf. Comput.
Sci., 33, 501–516.
Lopez, M., Kohler, S., and Winum,
J.Y. (2012) Zinc metalloenzymes as
new targets against the bacterial
pathogen Brucella. J. Inorg. Biochem.,
111, 138–145.
Nishimori, I., Minakuchi, T., Vullo,
D., Scozzafava, A., and Supuran, C.T.
(2011) Inhibition studies of the betacarbonic anhydrases from the bacterial
pathogen Salmonella enterica serovar
Typhimurium with sulfonamides and
sulfamates. Bioorg. Med. Chem., 19,
5023–5030.
Supuran, C.T. (2011) Bacterial carbonic
anhydrases as drug targets: toward novel
antibiotics? Front. Pharmacol., 2, 34.
Vullo, D., Nishimori, I., Minakuchi,
T., Scozzafava, A., and Supuran, C.T.
(2011) Inhibition studies with anions
and small molecules of two novel betacarbonic anhydrases from the bacterial
pathogen Salmonella enterica serovar
Typhimurium. Bioorg. Med. Chem. Lett.,
21, 3591–3595.
Vullo, D., De Luca, V., Scozzafava, A.,
Carginale, V., Rossi, M., Supuran, C.T.,
and Capasso, C. (2012) The first activation study of a bacterial carbonic anhydrase (CA). The thermostable alpha-CA
from Sulfurihydrogenibium yellowstonense YO3AOP1 is highly activated by
10.
11.
12.
13.
14.
15.
amino acids and amines. Bioorg. Med.
Chem. Lett., 22, 6324–6327.
Akdemir, A., Vullo, D., Luca, V.D.,
Scozzafava, A., Carginale, V., Rossi, M.,
Supuran, C.T., and Capasso, C. (2013)
The extremo-alpha-carbonic anhydrase (CA) from Sulfurihydrogenibium
azorense, the fastest CA known, is
highly activated by amino acids and
amines. Bioorg. Med. Chem. Lett., 23,
1087–1090.
De Luca, V., Vullo, D., Scozzafava, A.,
Carginale, V., Rossi, M., Supuran, C.T.,
and Capasso, C. (2012) Anion inhibition
studies of an alpha-carbonic anhydrase from the thermophilic bacterium
Sulfurihydrogenibium yellowstonense
YO3AOP1. Bioorg. Med. Chem. Lett., 22,
5630–5634.
Luca, V.D., Vullo, D., Scozzafava, A.,
Carginale, V., Rossi, M., Supuran,
C.T., and Capasso, C. (2013) An
alpha-carbonic anhydrase from the
thermophilic bacterium Sulphurihydrogenibium azorense is the fastest enzyme
known for the CO(2) hydration reaction.
Bioorg. Med. Chem., 21, 1534–1538.
Vullo, D., De Luca, V., Scozzafava, A.,
Carginale, V., Rossi, M., Supuran, C.T.,
and Capasso, C. (2012) Anion inhibition
studies of the fastest carbonic anhydrase
(CA) known, the extremo-CA from
the bacterium Sulfurihydrogenibium
azorense. Bioorg. Med. Chem. Lett., 22,
7142–7145.
Vullo, D., Luca, V.D., Scozzafava, A.,
Carginale, V., Rossi, M., Supuran,
C.T., and Capasso, C. (2013) The
alpha-carbonic anhydrase from the
thermophilic bacterium Sulfurihydrogenibium yellowstonense YO3AOP1
is highly susceptible to inhibition by
sulfonamides. Bioorg. Med. Chem., 21,
1465–14699.
Nishimori, I., Vullo, D., Minakuchi,
T., Scozzafava, A., Capasso, C., and
Supuran, C.T. (2013) Restoring catalytic
41
42
3 Developing Novel Bacterial Targets: Carbonic Anhydrases as Antibacterial Drug Targets
16.
17.
18.
19.
20.
21.
22.
23.
activity to the human carbonic anhydrase (CA) related proteins VIII, X and
XI affords isoforms with high catalytic
efficiency and susceptibility to anion
inhibition. Bioorg. Med. Chem. Lett., 23,
256–260.
Supuran, C.T. (2008) Carbonic anhydrases: novel therapeutic applications for
inhibitors and activators. Nat. Rev. Drug
Discovery, 7, 168–181.
Del Prete, S., Vullo, D., Scozzafava, A.,
Capasso, C., and Supuran, C.T. (2014)
Cloning, characterization and anion
inhibition study of the delta-class carbonic anhydrase (TweCA) from the
marine diatom Thalassiosira weissflogii.
Bioorg. Med. Chem., 22, 531–536.
Nishimori, I., Minakuchi, T., Onishi,
S., Vullo, D., Scozzafava, A., and
Supuran, C.T. (2007) Carbonic anhydrase
inhibitors. DNA cloning, characterization, and inhibition studies of the human
secretory isoform VI, a new target for
sulfonamide and sulfamate inhibitors.
J. Med. Chem., 50, 381–388.
Sacarlal, J., Nhacolo, A.Q., Sigauque, B.,
Nhalungo, D.A., Abacassamo, F., Sacoor,
C.N., Aide, P., Machevo, S., Nhampossa,
T., Macete, E.V., Bassat, Q., David, C.,
Bardaji, A., Letang, E., Saute, F., Aponte,
J.J., Thompson, R., and Alonso, P.L.
(2009) A 10 year study of the cause
of death in children under 15 years in
Manhica, Mozambique. BMC Public
Health, 9, 67.
Nour, N.M. (2012) Premature delivery
and the millennium development goal.
Rev. Obstet. Gynecol., 5, 100–105.
Gaynor, M. and Mankin, A.S. (2003)
Macrolide antibiotics: binding site,
mechanism of action, resistance. Curr.
Top. Med. Chem., 3, 949–961.
Rusconi, S., Scozzafava, A.,
Mastrolorenzo, A., and Supuran, C.T.
(2004) New advances in HIV entry
inhibitors development. Curr. Drug
Targets Infect. Disord., 4, 339–355.
Innocenti, A., Hall, R.A., Schlicker,
C., Scozzafava, A., Steegborn, C.,
Muhlschlegel, F.A., and Supuran, C.T.
(2009) Carbonic anhydrase inhibitors.
Inhibition and homology modeling
studies of the fungal beta-carbonic
anhydrase from Candida albicans with
24.
25.
26.
27.
28.
29.
30.
31.
32.
sulfonamides. Bioorg. Med. Chem., 17,
4503–4509.
Maresca, A., Scozzafava, A., Kohler, S.,
Winum, J.Y., and Supuran, C.T. (2012)
Inhibition of beta-carbonic anhydrases
from the bacterial pathogen Brucella suis
with inorganic anions. J. Inorg. Biochem.,
110, 36–39.
Supuran, C.T. (2012) Inhibition of bacterial carbonic anhydrases and zinc
proteases: from orphan targets to innovative new antibiotic drugs. Curr. Med.
Chem., 19, 831–844.
Capasso, C. and Supuran, C.T. (2013)
Anti-infective carbonic anhydrase
inhibitors: a patent and literature review.
Expert Opin. Ther. Pat., 23, 693–704.
Bertucci, A., Innocenti, A., Scozzafava,
A., Tambutte, S., Zoccola, D., and
Supuran, C.T. (2011) Carbonic anhydrase inhibitors. Inhibition studies
with anions and sulfonamides of a new
cytosolic enzyme from the scleractinian
coral Stylophora pistillata. Bioorg. Med.
Chem. Lett., 21, 710–714.
Temperini, C., Scozzafava, A., and
Supuran, C.T. (2010) Carbonic anhydrase
inhibitors. X-ray crystal studies of the
carbonic anhydrase II-trithiocarbonate
adduct--an inhibitor mimicking the
sulfonamide and urea binding to the
enzyme. Bioorg. Med. Chem. Lett., 20,
474–478.
Supuran, C.T. (2010) Carbonic anhydrase
inhibitors. Bioorg. Med. Chem. Lett., 20,
3467–3474.
Sulek, K. (1968) Nobel prize for Gerhard
Domagk in 1939 for discovery of the
antibacterial activity of prontosil. Wiad.
Lek., 21, 1089.
Del Prete, S., Isik, S., Vullo, D., De Luca,
V., Carginale, V., Scozzafava, A.,
Supuran, C.T., and Capasso, C. (2012)
DNA cloning, characterization, and
inhibition studies of an alpha-carbonic
anhydrase from the pathogenic bacterium Vibrio cholerae. J. Med. Chem.,
55, 10742–10748.
Luca, V.D., Vullo, D., Scozzafava, A.,
Carginale, V., Rossi, M., Supuran,
C.T., and Capasso, C. (2013) An
alpha-carbonic anhydrase from the
thermophilic bacterium Sulphurihydrogenibium azorense is the fastest enzyme
References
33.
34.
35.
36.
37.
38.
39.
known for the CO2 hydration reaction.
Bioorg. Med. Chem., 21, 1465–1469.
Monti, S.M., De Simone, G., Dathan,
N.A., Ludwig, M., Vullo, D., Scozzafava,
A., Capasso, C., and Supuran, C.T.
(2013) Kinetic and anion inhibition
studies of a beta-carbonic anhydrase
(FbiCA 1) from the C4 plant Flaveria
bidentis. Bioorg. Med. Chem. Lett., 23,
1626–1630.
Pan, P., Vermelho, A.B.,
Capaci Rodrigues, G., Scozzafava, A.,
Tolvanen, M.E., Parkkila, S., Capasso, C.,
and Supuran, C.T. (2013) Cloning, characterization, and sulfonamide and thiol
inhibition studies of an alpha-carbonic
anhydrase from Trypanosoma cruzi,
the causative agent of Chagas disease.
J. Med. Chem., 56, 1761–1771.
Vullo, D., Leewattanapasuk, W.,
Muhlschlegel, F.A., Mastrolorenzo, A.,
Capasso, C., and Supuran, C.T. (2013)
Carbonic anhydrase inhibitors: inhibition of the beta-class enzyme from
the pathogenic yeast Candida glabrata
with sulfonamides, sulfamates and sulfamides. Bioorg. Med. Chem. Lett., 23,
2647–2652.
Vullo, D., Luca, V.D., Scozzafava, A.,
Carginale, V., Rossi, M., Supuran,
C.T., and Capasso, C. (2013) The
alpha-carbonic anhydrase from the
thermophilic bacterium Sulfurihydrogenibium yellowstonense YO3AOP1
is highly susceptible to inhibition by
sulfonamides. Bioorg. Med. Chem., 21,
1534–1538.
Cardoso, T., Ribeiro, O., Aragao, I.C.,
Costa-Pereira, A., and Sarmento, A.E.
(2012) Additional risk factors for infection by multidrug-resistant pathogens in
healthcare-associated infection: a large
cohort study. BMC Infect. Dis., 12, 375.
Cox, G.M., McDade, H.C., Chen,
S.C., Tucker, S.C., Gottfredsson, M.,
Wright, L.C., Sorrell, T.C., Leidich, S.D.,
Casadevall, A., Ghannoum, M.A., and
Perfect, J.R. (2001) Extracellular phospholipase activity is a virulence factor
for Cryptococcus neoformans. Mol.
Microbiol., 39, 166–175.
Cox, G.M., Mukherjee, J., Cole, G.T.,
Casadevall, A., and Perfect, J.R. (2000)
40.
41.
42.
43.
44.
45.
46.
Urease as a virulence factor in experimental cryptococcosis. Infect. Immun.,
68, 443–448.
Chirica, L.C., Elleby, B., Jonsson, B.H.,
and Lindskog, S. (1997) The complete
sequence, expression in Escherichia
coli, purification and some properties
of carbonic anhydrase from Neisseria
gonorrhoeae. Eur. J. Biochem./FEBS, 244,
755–760.
Di Fiore, A., Capasso, C., De Luca, V.,
Monti, S.M., Carginale, V., Supuran, C.T.,
Scozzafava, A., Pedone, C., Rossi, M.,
and De Simone, G. (2013) X-ray structure of the first ‘extremo-alpha-carbonic
anhydrase’, a dimeric enzyme from the
thermophilic bacterium Sulfurihydrogenibium yellowstonense YO3AOP1.
Acta Crystallogr. D Biol. Crystallogr., 69,
1150–1159.
Morishita, S., Nishimori, I., Minakuchi,
T., Onishi, S., Takeuchi, H., Sugiura, T.,
Vullo, D., Scozzafava, A., and Supuran,
C.T. (2008) Cloning, polymorphism, and
inhibition of beta-carbonic anhydrase of
Helicobacter pylori. J. Gastroenterol., 43,
849–857.
Nishimori, I., Minakuchi, T., Kohsaki,
T., Onishi, S., Takeuchi, H., Vullo, D.,
Scozzafava, A., and Supuran, C.T. (2007)
Carbonic anhydrase inhibitors: the betacarbonic anhydrase from Helicobacter
pylori is a new target for sulfonamide
and sulfamate inhibitors. Bioorg. Med.
Chem. Lett., 17, 3585–3594.
Nishimori, I., Minakuchi, T., Morimoto,
K., Sano, S., Onishi, S., Takeuchi,
H., Vullo, D., Scozzafava, A., and
Supuran, C.T. (2006) Carbonic anhydrase
inhibitors: DNA cloning and inhibition
studies of the alpha-carbonic anhydrase
from Helicobacter pylori, a new target
for developing sulfonamide and sulfamate gastric drugs. J. Med. Chem., 49,
2117–2126.
Nishimori, I., Onishi, S., Takeuchi, H.,
and Supuran, C.T. (2008) The alpha and
beta classes carbonic anhydrases from
Helicobacter pylori as novel drug targets.
Curr. Pharm. Des., 14, 622–630.
Nishimori, I., Vullo, D., Minakuchi, T.,
Morimoto, K., Onishi, S., Scozzafava,
A., and Supuran, C.T. (2006) Carbonic anhydrase inhibitors: cloning
43
44
3 Developing Novel Bacterial Targets: Carbonic Anhydrases as Antibacterial Drug Targets
47.
48.
49.
50.
51.
52.
53.
and sulfonamide inhibition studies of
a carboxyterminal truncated alphacarbonic anhydrase from Helicobacter
pylori. Bioorg. Med. Chem. Lett., 16,
2182–2188.
Shahidzadeh, R., Opekun, A., Shiotani,
A., and Graham, D.Y. (2005) Effect
of the carbonic anhydrase inhibitor,
acetazolamide, on Helicobacter pylori
infection in vivo: a pilot study. Helicobacter, 10, 136–138.
Del Prete, S., De Luca, V., Scozzafava,
A., Carginale, V., Supuran, C.T., and
Capasso, C. (2014) Biochemical properties of a new alpha-carbonic anhydrase
from the human pathogenic bacterium,
Vibrio cholerae. J. Enzyme Inhib. Med.
Chem., 29, 23–27.
Vullo, D., Isik, S., Del Prete, S., De Luca,
V., Carginale, V., Scozzafava, A.,
Supuran, C.T., and Capasso, C. (2013)
Anion inhibition studies of the alphacarbonic anhydrase from the pathogenic
bacterium Vibrio cholerae. Bioorg. Med.
Chem. Lett., 23, 1636–1638.
Joseph, P., Ouahrani-Bettache, S.,
Montero, J.L., Nishimori, I., Minakuchi,
T., Vullo, D., Scozzafava, A., Winum,
J.Y., Kohler, S., and Supuran, C.T. (2011)
A new beta-carbonic anhydrase from
Brucella suis, its cloning, characterization, and inhibition with sulfonamides
and sulfamates, leading to impaired
pathogen growth. Bioorg. Med. Chem.,
19, 1172–1178.
Vullo, D., Nishimori, I., Scozzafava, A.,
Kohler, S., Winum, J.Y., and Supuran,
C.T. (2010) Inhibition studies of a betacarbonic anhydrase from Brucella suis
with a series of water soluble glycosyl
sulfanilamides. Bioorg. Med. Chem. Lett.,
20, 2178–2182.
Buchieri, M.V., Riafrecha, L.E.,
Rodriguez, O.M., Vullo, D., Morbidoni,
H.R., Supuran, C.T., and Colinas, P.A.
(2013) Inhibition of the beta-carbonic
anhydrases from Mycobacterium tuberculosis with C-cinnamoyl glycosides:
identification of the first inhibitor with
anti-mycobacterial activity. Bioorg. Med.
Chem. Lett., 23, 740–743.
Maresca, A., Vullo, D., Scozzafava, A.,
Manole, G., and Supuran, C.T. (2013)
Inhibition of the beta-class carbonic
54.
55.
56.
57.
58.
59.
60.
61.
anhydrases from Mycobacterium tuberculosis with carboxylic acids. J. Enzyme
Inhib. Med. Chem., 28, 392–396.
Guzel, O., Maresca, A., Hall, R.A.,
Scozzafava, A., Mastrolorenzo, A.,
Muhlschlegel, F.A., and Supuran, C.T.
(2010) Carbonic anhydrase inhibitors.
The beta-carbonic anhydrases from the
fungal pathogens Cryptococcus neoformans and Candida albicans are strongly
inhibited by substituted-phenyl-1Hindole-5-sulfonamides. Bioorg. Med.
Chem. Lett., 20, 2508–2511.
Vullo, D., Sai Kumar, R.S., Scozzafava,
A., Capasso, C., Ferry, J.G., and Supuran,
C.T. (2013) Anion inhibition studies of a
beta-carbonic anhydrase from Clostridium perfringens. Bioorg. Med. Chem.
Lett., 23, 6706–6710.
Kumar, R.S., Hendrick, W., Correll, J.B.,
Patterson, A.D., Melville, S.B., and Ferry,
J.G. (2013) Biochemistry and physiology
of the beta class carbonic anhydrase
(Cpb) from Clostridium perfringens
strain 13. J. Bacteriol., 195, 2262–2269.
Innocenti, A., Zimmerman, S., Ferry,
J.G., Scozzafava, A., and Supuran,
C.T. (2004) Carbonic anhydrase
inhibitors. Inhibition of the zinc and
cobalt gamma-class enzyme from the
archaeon Methanosarcina thermophila
with anions. Bioorg. Med. Chem. Lett.,
14, 3327–3331.
Zimmerman, S., Domsic, J.F., Tu, C.,
Robbins, A.H., McKenna, R., Silverman,
D.N., and Ferry, J.G. (2013) Role of
Trp19 and Tyr200 in catalysis by the
gamma-class carbonic anhydrase from
Methanosarcina thermophila. Arch.
Biochem. Biophys., 529, 11–17.
Zimmerman, S.A., Tomb, J.F., and Ferry,
J.G. (2010) Characterization of CamH
from Methanosarcina thermophila,
founding member of a subclass of the
{gamma} class of carbonic anhydrases.
J. Bacteriol., 192, 1353–1360.
Ingram-Smith, C., Gorrell, A., Lawrence,
S.H., Iyer, P., Smith, K., and Ferry, J.G.
(2005) Characterization of the acetate
binding pocket in the Methanosarcina
thermophila acetate kinase. J. Bacteriol.,
187, 2386–2394.
Lawrence, S.H., Luther, K.B., Schindelin,
H., and Ferry, J.G. (2006) Structural and
References
functional studies suggest a catalytic
63. Del Prete, S., Vullo, D., De Luca, V.,
mechanism for the phosphotransacetyCarginale, V., Scozzafava, A., Supuran,
lase from Methanosarcina thermophila.
C.T., and Capasso, C. (2013) A highly
J. Bacteriol., 188, 1143–1154.
catalytically active gamma-carbonic
anhydrase from the pathogenic anaer62. Del Prete, S., De Luca, V., Vullo, D.,
obe Porphyromonas gingivalis and its
Scozzafava, A., Carginale, V., Supuran,
inhibition profile with anions and small
C.T., and Capasso, C. (2013) Biochemical
molecules. Bioorg. Med. Chem. Lett., 23,
characterization of the gamma-carbonic
4067–4071.
anhydrase from the oral pathogen Porphyromonas gingivalis, PgiCA. J. Enzyme
Inhib. Med. Chem., in press.
45
47
4
Magainins – A Model for Development of Eukaryotic
Antimicrobial Peptides (AMPs)
Sarah R. Dennison, Frederick Harris, and David A. Phoenix
4.1
Introduction
Microbial infection is a threat constantly faced by all living creatures and it
has long been known that the adaptive immune system provides protection
against this threat. However, it has also long been known that plants and insects,
which lack an adaptive immune system, remain free from these infections for
most of the time. The reasons for this immunity from infection were largely
a mystery until the answer was provided, in part, by the discovery of host
defense peptides (HDPs) in both plants [1] and insects [2]. It is now known
that HDPs are produced by all eukaryotes [3] and are critical effectors of both
adaptive and innate immunity. Indeed, they serve a wide range of functions,
including direct antimicrobial activity [4, 5] and a host of immune-modulatory
effects [6, 7]. There is some debate as to what the discovery of HDPs should be
attributed but it is generally accepted that two landmark studies in the 1980s
played a major role in focusing research on these peptides [8]. In 1980, two
peptides with antimicrobial activity, P9A and P9B, were isolated from the pupae of
the silk moth, Hyalophora cecropia [2] and were characterized and renamed as the
now familiar HDPs, “cecropins” [9]. Around 6 years later, peptides with antimicrobial activity were independently identified in skin secretions of the African clawed
frog, Xenopus laevis, by Williams and colleagues [10] in the United Kingdom and
by Zasloff [11] in the United States. The latter reported two closely related peptides, which were derived from a common, larger protein and non-hemolytic while
possessing potent activity against numerous protozoa, fungi, and bacteria [11,
12]. On the basis of these results, Zasloff [11] observed that “ … these peptides
may be responsible for the extraordinary freedom from infection characteristic
of wound healing in this animal and appear to constitute a previously unrecognized vertebrate antimicrobial host defence system.” To recognize their role as
HDPs in the defense systems of X. laevis, these peptides were named magainins 1 and 2 after the Hebrew word for “Shield” [11]. Since this seminal study,
it has become clear that the host defense role of magainins is not restricted to
X. laevis with homologs of these peptides identified in the skin secretions of other
Novel Antimicrobial Agents and Strategies, First Edition.
Edited by David A. Phoenix, Frederick Harris and Sarah R. Dennison.
© 2015 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2015 by Wiley-VCH Verlag GmbH & Co. KGaA.
48
4 Magainins – A Model for Development of Eukaryotic Antimicrobial Peptides (AMPs)
Table 4.1 Magainins from Xenopodinae.
Host frog
Magainins
Sequence of magainins
References
X. laevis
X. laevis
X. borealis
X. borealis
X. clivii
X. clivii
X. muelleri
X. muelleri
X. muelleri West
X. petersii
X. petersii
X. amieti
X. amieti
X. andrei
X. andrei
X. pygmaeus
X. pygmaeus
X. lenduensis
X. lenduensis
X. laevis × X.
muelleri hybrid
S. tropicalis
S. epitropicalis
Magainin-1
Magainin-2
Magainin-B1
Magainin-B2
Magainin-C1
Magainin-C2
Magainin-M1
Magainin-M2
Magainin-MW1
Magainin-P1
Magainin-P2
Magainin-AM1
Magainin-AM2
Magainin-AN1
Magainin-AN2
Magainin-PG1
Magainin-PG2
Magainin-L1
Magainin-L2
Magainin-LM1
GIGKFLHSAGKFGKAFVGEIMKS
GIGKFLHSAKKFGKAFVGEIMNS
GKFLHSAGKFGKAFLGEVMIG
GIGKFLHSAGKFGKAFLGEVMKS
GVGKFLHSAKKFGQALASEIMKS
GVGKFLHSAKKFGQALVSEIMKS
GIGKFLHSAGKFGKAFIGEIMKS
GFKQFVHSLGKFGKAFVGEMIKPK
GIGKFLHSAGKFGKAFLGEVMKS
GIGKFLHSAGKFGKAFVGEIMKS
GIGQFLHSAKKFGKAFVGEIMKS
GIKEFAHSLGKFGKAFVGGILNQ
GVSKILHSAGKFGKAFLGEIMKS
GIKEFAHSLGKFGKAFVGGILNQ
GVSKILHSAGKFGKAFLGEIMKS
GVGKFLHAAGKFGKALMGEMMKS
GVSQFLHSASKFGKALMGEIMKS
GIGKFLHSAKKFGKAFVGEVMKS
GISQFLHSAKKFGKAFAGEIMKS
GIGKFLHSAKKFAKAFVGEIMNS
[13]
[13]
[13]
[13]
[13]
[13]
[13]
[13]
[13]
[22]
[22]
[13]
[13]
[13]
[13]
[22]
[22]
[22]
[22]
[23]
Magainin-ST1
Magainin-SE1
GLKEVAHSAKKFAKGFISGLTGS
GLKEVLHSTKKFAKGFITGLTGQ
[24]
[25]
This table shows magainins identified in various species of Xenopodinae. These peptides show
considerable levels of sequence homology and phylogenetic analyzes have suggested that they may
have evolved from a common ancestral gene by a series of duplication events [13].
Table 4.2 Major analogs of magainins.
Magainin analog
Peptide sequence
Magainin-2a
F5W-magainin-2
Magainin-A
Magainin-H1
Magainin-H2
MSI-78 (Pexiganan)
MSI-594
MSI-99
MSI-843
GIGKFLHSAKKFGKAFVGEIMNS-NH2
GIGKWLHSAKKFGKAFVGEIMNS
(βA)IGKFLHAAKKFAKAFVAEIMNS-NH2
GIKKFLHIIWKFIKAFVGEIMNS
IIKKFLHSIWKFGKAFVGEIMNI
GIGKFLKKAKKFGKAFVKILKK-NH2
GIGKFLKKAKKGIGAVLKVLTTGL-NH2
GIGKFLKSAKKFGKAFVKILNS-NH2
Oct–OOLLOOLOOL–NH2
References
[26]
[27]
[19]
[28]
[28]
[29]
[30]
[31]
[32]
4.2
Magainins and Their Antimicrobial Action
species belonging to the genera Xenopus and Silurana, which together comprise
the Xenopodinae (Table 4.1) [13]. Not long after their discovery, the therapeutic
potential of magainins was recognized [14, 15] and in response, numerous analogs
of these peptides have been developed to maximize broad spectrum antimicrobial
activity and minimize cytotoxicity (Table 4.2) [16–21]. Currently, magainins and
their derivatives are under development for novel usage in a number of medical and biotechnological applications. In this chapter, we present an overview of
recent progress in major examples of these applications.
4.2
Magainins and Their Antimicrobial Action
As one of the first families of HDPs to be discovered, there has been intense
investigation into the structure/function relationships underpinning the antimicrobial action of magainins and their analogs [33, 34]. On the basis of these
studies, it has been established that the ability of magainins and most other
HDPs to kill microbes is generally related to their direct interactions with the
membranes of these organisms rather than through the use of specific, chiral
receptors [8, 35–37], although a few exceptions are known [8, 38–41]. To initiate
these interactions, magainins, along with the vast majority of HDPs, carry a
net positive charge, thereby allowing them to show increased selectivity for the
negatively charged target microbial membranes over mammalian membranes,
which exhibit no overall charge [8, 35]. Once in the interfacial environment
of the targeted microbial membrane, magainins, which are unstructured in
aqueous solution, undergo a conformational change to adopt high levels of
α-helical secondary structure. These structures exhibit the spatial segregation of
hydrophobic and hydrophilic residues [42], which allows magainins to orientate
parallel to the microbial membrane surface such that their hydrophilic residues
associate with the bilayer lipid head-group region while their hydrophobic
residues interact with its apolar lipid acyl chains [33, 34]. It is generally accepted
that this mode of bilayer partitioning represents an early step in the mechanisms
Table 4.3 Major membrane disruptive mechanisms for magainins and analogs.
Mechanisms
Magainins and analogs
Toroidal pore
Disordered toroidal pore
Aggregate model
Interfacial activity model
Chaotic or non-stoichiometric model
Carpet mechanism
Membrane thinning/thickening
Charged lipid clustering
Magainin-2
Magainin-H2
Magainin-2 and analogs
Magainin-2
Magainin-2
F5W-magainin-2
Magainin-2
MSI-78 and analogs
References
[48]
[62]
[49]
[63]
[64]
[65]
[66]
[67]
49
50
4 Magainins – A Model for Development of Eukaryotic Antimicrobial Peptides (AMPs)
used by magainins to disrupt microbial membranes, although a number of
other models have been presented to describe these mechanisms (Table 4.3)
[8, 43–47]. The most commonly cited model was first presented in the 1990s
and proposes that these peptides induce the death of microbial cells through
toroidal pore formation [48, 49]. Essentially, this model involves the association
of magainins with membrane lipids to form a supramolecular arrangement of
high curvature, thereby transiently forming pores that can induce cell death
through membrane dysfunction and the leakage of essential cellular contents
(Figure 4.1) [34, 44]. Variants of this model have been proposed by more recent
studies (Table 4.3) [8, 50–53]. However, regardless of the specific antimicrobial
Lys 11
Lys 14
Lys 10
Lys 4
N
(a)
C
Phe 12
Phe 5
Phe 16
+
+
+
+
+ +
+
(b)
Figure 4.1 Magainins and membrane interaction. Figure (a) shows magainin-2, which
is generally accepted as the prototype of
α-helical HDPs [8]. The α-helical architecture of the peptide shows a segregation of
polar and apolar amino acid residues, giving the molecule amphiphilic characteristics. For clarity, the polar face of the α-helix
is indicated by the peptide’s lysine residues
and its apolar face residues by phenylaniline
residues. Figure (b) shows magainin-2 permeabilizing the membrane using a toroidal
pore mechanism. Initially, the peptide is orientated parallel to the bilayer surface with
its apolar residues buried in the hydrophobic membrane core and its polar residues
associating with the lipid head-group region.
+
+
+
However, the aggregation of magainin-2
molecules on the membrane surface imposes
a positive curvature strain by increasing
the distance between membrane lipid head
groups. When the aggregation of these peptide molecules reaches a critical concentration, they realign perpendicular to the
bilayer, causing the membrane surface to
cavitate inwards and ultimately form a pore.
In this pore, magainin-2 molecules remain in
close association with membrane lipid head
groups such that these pores are lined by
polar lipid head groups and hydrophilic surfaces of peptide molecules. These pores can
then induce cell death through membrane
dysfunction and the leakage of essential cellular contents [8].
4.3
Magainins as Antibiotics
mechanisms ascribed to magainins, these mechanisms are relatively nonspecific
and generally involve attack on multiple hydrophobic/polyanionic targets as
opposed to the limited number of targets used by conventional antibiotics [54].
On the basis of these observations, it is generally believed that it is difficult
for microbes to acquire resistance to the action of magainins and other HDPs
[5, 55, 56]. For example, using experimental evolution, multiple passages at
sub-inhibitory concentrations of magainins were required to produce bacterial
strains that were nonsusceptible to these peptides [57, 58]. These observations
clearly give magainins and other HDPs a major advantage over conventional
antibiotics and make them attractive propositions for development as potential
next generation antimicrobials [54, 59–61].
4.3
Magainins as Antibiotics
Efforts to develop magainins as clinically relevant antibiotics stem from studies
in the 1990s, which focused on the effect of changing a variety of structural and
physiochemical properties on the antimicrobial action of these peptides [49,
68] including amino acid composition, sequence length, charge, and α-helicity
[18, 19, 21, 69, 70]. The insight gained from these studies coupled with that
gained from further structure/function studies by Zasloff and colleagues led to
the development of MSI-78 or pexiganan [71], which is a more cationic analog
of magainin 2 that contains five additional lysine residues and an α-amidated
C-terminus (Table 4.2). Since these studies, MSI-78 has been extensively characterized and shown to possess activity against a broad range of Gram-positive
and Gram-negative bacteria that involves membrane disruption via a toroidal
pore-type mechanism, which appears to be generally the case for all magainins
and their analogs so far tested [42, 72, 73].
The therapeutic potential of MSI-78 as an antibiotic for systemic administration was suggested by in vitro studies, which showed that the peptide was able
to synergize the action of a number of β-lactams against bacteria responsible for
bloodstream infections in neutropenic patients, including Staphylococci, which
are currently the primary cause of nosocomial bacteremia [74]. A similar potential was suggested for magainin-2 when it was found that the peptide synergized
the activity of a range of conventional antibiotics against both Gram-positive and
Gram-negative organisms. In particular, the peptide was able to strongly enhance
the action of β-lactams such as ceftriaxone and meropenem against strains of
oxacillin-resistant Staphylococcus aureus, which were resistant to these antibiotics
when acting alone [8, 75]. On the basis of these observations, it was suggested
that magainins and their analogs may be able to synergize the systemic administration of conventional antibiotics to combat microbial pathogens [14]. Strongly
supporting this suggestion, the administration of magainins was found to synergize various β-lactams such as cefepime in the treatment of Escherichia coli
51
52
4 Magainins – A Model for Development of Eukaryotic Antimicrobial Peptides (AMPs)
infections in both mice and mice with induced neutropenia [14, 76, 77]. More
recently, using a rat model of septic shock, several studies showed that the administration of magainin 1, magainin 2, magainin 2a, and MSI-78, either alone or in
combination with β-lactams, strongly inhibited the growth of Gram-negative bacteria. The highest antibacterial efficacy obtained in these studies was observed
following the co-administration of these peptides with piperacillin and iminipem.
This was accompanied by strong reductions in the plasma endotoxin levels normally associated with β-lactam activity and attributed to the ability of magainins to
bind lipopolysaccharide (LPS). On the basis of these observations, it was suggested
that β-lactams and/or magainins and their analogs may act as lead compounds
for the investigation of novel agents with activity against septic shock induced
by Gram-negative pathogens [78, 79]. A recently developed strategy to synergize
the activity of HDPs and conventional antibiotics has been to conjugate the two
molecules [80]. For example, magainin 2 and a truncated analog were conjugated
to vancomycin and it was found that the conjugated magainin analog showed
increased antibacterial activity against vancomycin-resistant Enterococci as compared to vancomycin alone [81].
Currently, the major therapeutic potential shown by MSI-78 is for topical application as an antibacterial agent, which developed from studies on the peptide’s
ability to treat a variety of conditions such as impetigo and its ability to support wound healing associated with various infections [71, 72]. These studies led
to MSI-78 becoming the first of the HDPs to undergo commercial development
[54, 82] when the peptide entered clinical trials for the topical treatment of diabetic foot ulcers [59, 71, 72] whose infections are the most common causes of
hospitalizations and amputations in diabetic patients [83]. In phase III trials, MSI78 achieved clinical cure or improvement in about 90% of diabetic patients with
infected foot ulcers and the peptide was well tolerated by these individuals. However, the US Food and Drug Administration (FDA) did not approve marketing
of the peptide after concluding that it offered no improvement over the conventional treatment of foot ulcers with ofloxacin, a fluoroquinolone antibiotic [59,
72, 84]. However, this conclusion has been questioned [84, 85] and there is hope
that MSI-78 may yet achieve commercial success with advances in areas such as
peptide manufacturing and clinical trial design [59, 72, 84]. In response, there
has been continuing research into mechanisms by which the peptide exerts its
membrane interactions and antibacterial activity [47, 86–99]. In general, these
studies support the toroidal pore mechanism proposed for microbial membrane
disruption by MSI-78 and have provided further insight into the dynamics of the
peptide–lipid interactions involved in the use of this mechanism [91, 96, 99].
These studies have also emphasized the importance of membrane composition
to the selectivity of MSI-78 for bacterial cells over eukaryotic cells [95] and the
ability of the peptide to inactivate bacteria, particularly the role of anionic lipid
[67]. The presence of anionic lipid in bacterial membranes is key to the ability of
MSI-78 to target and disrupt these membranes but the presence of high levels of
this lipid in the bilayer was shown to inhibit this disruptive ability [86]. Consistent
with these results, the binding of strongly cationic MSI-78 to LPS, which is the
4.3
Magainins as Antibiotics
major anionic component of the outer membrane, was found to inhibit the ability
of the peptide to traverse these membranes and thereby its potential to induce the
death of some Gram-negative bacteria via disruption of the inner membrane [89].
A number of studies have shown that MSI-78 is able to induce the segregation
of anionic lipids from zwitterionic lipids within the inner membranes of Gramnegative bacteria [47] and different anionic lipids from one another within the
inner membranes of Gram-positive bacteria [97]. In both cases, it was proposed
that this anionic lipid clustering effect could play an important role in the antibacterial mechanism of MSI-78 such as by contributing to membrane disruption and
impairing the function of bilayer proteins [47, 97].
The insight provided by the continuing structure/function studies on MSI-78
has helped not only to provide a clearer picture of the antibacterial action used
by the peptide but has also made an important contribution to efforts aimed at
producing novel analogs and structural mimics of MSI-78 with improved stability
and antimicrobial efficacy. A comprehensive review of these novel compounds is
beyond the scope of this chapter, but a survey of the literature over the last decade
showed that these compounds include de novo peptides [100], acylated analogs
[32, 101, 102], fluorinated analogs [99, 103–106], polyethylene glycol (PEG)
ylated analogs [107], β-peptides [108], peptoids [109, 110], oligourea polymers
[111, 112], aryl-based compounds [113, 114], poly(amidoamine) dendrimers
[115], polynorbornenes [116], meta-phenylene ethynylenes [117, 118], and other
“MSI” compounds [16, 30, 32, 47, 87, 92–95, 97, 98, 119–135]. Many of these
molecules are membrane interactive and show antibacterial activity, which is
comparable or superior to that of MSI-78, but in general, they are not well studied
with the best characterized belonging to the “MSI” group of compounds [16, 30,
32, 47, 87, 92–95, 97, 98, 119–135]. For example, MSI-843 is a 10 residue lipopeptide, which contains 6 ornithine residues and a conjugated C-terminal octanyl
moiety (Table 4.2), and was originally developed as a potential treatment for
Pseudomonas aeruginosa infection in cystic fibrosis patients [136]. More recent
studies have shown that MSI-843 has activity against other Gram-negative organisms, including E. coli as well as Gram-positive bacteria such as S. aureus [32, 47,
67]. On the basis of this broad range antibacterial activity, the lipopeptide has
recently been included in a patent, along with MSI-78 and MSI-594, for use in the
treatment of infections associated with disorders of the skin [137] and sinonasal
cavity [138]. In other studies, MSI-751 (N-amidino-phenylalanyl-dioctylamide)
and MSI-774 (1,12-[Di-(N-amidino-arginine-phenylalanyl)]diaminododecane)
were found to show potent and rapid antibacterial activity when directed against
a range of oral pathogens such as Porphyromonas gingivalis, Fusobacterium
nucleatum, and Prevotella spp., which led to the suggestion that these peptides
may be suitable for development as agents in the prevention and treatment of
periodontal diseases [119].
On the basis of the progress made in the development of magainins and
their analogs from X. laevis, over the last few years a number of studies have
investigated the antibacterial capabilities of magainins recently identified in other
species of the Xenopodinae and hybrids of these species (Table 4.4) [139, 140].
53
54
4 Magainins – A Model for Development of Eukaryotic Antimicrobial Peptides (AMPs)
Table 4.4 The antimicrobial and hemolytic abilities of representative magainins.
Magainin
S. aureus
Magainin-1
Magainin-2
Magainin-B1
Magainin-B2
Magainin-C1
Magainin-C2
Magainin-P1
Magainin-P2
Magainin-M1
Magainin-MW1
Magainin-AM1
Magainin-AM2
Magainin-AN2
Magainin-PG1
Magainin-PG2
Magainin-L1
Magainin-L2
Magainin-ST1
Magainin-SE1
< 40
< 40
> 100
50
> 100
> 100
> 50
> 50
50
> 200
> 200
50
> 100
> 50
> 50
> 50
> 50
> 256
> 160
E. coli
Candida albicans
< 40
< 30
> 100
> 100
100
50
> 50
> 50
12.5
50
> 200
> 200
100
> 50
> 50
> 50
> 50
64
> 160
—
35
—
100
—
—
—
—
25
100
—
—
—
—
—
—
—
—
—
EC50
430
430
—
> 200
> 200
> 200
> 50
> 50
180
> 200
—
—
—
> 50
> 50
> 50
> 50
> 256
> 160
References
[58, 69, 75, 149]
[11, 69, 75, 149, 150]
[151]
[151]
[152]
[152]
[22]
[22]
[141]
[141]
[143]
[143]
[142]
[22]
[22]
[22]
[22]
[24]
[25]
Interestingly, hybrids resulting from a cross between frogs from X. laevis and
Xenopus muelleri expressed a novel peptide, magainin-LM1, which was absent
from both parental species, although its antimicrobial activity has not yet been
investigated [23]. These more recently described magainins are as yet not well
characterized but showed similarities to their counterparts in X. laevis, including
strong cationicity, significant levels of sequence homology and very low levels of
hemolysis (Table 4.4) [13]. However, only magainin-M1 appeared to have broad
spectrum antibacterial activity (Table 4.4) comparable to that of magainins 1 and
2 and may therefore be suitable for development as a clinically relevant antibacterial. The remaining recently described magainins showed only weak antibacterial
activity with differing specificities (Table 4.4) and it was suggested that in vivo,
these peptides may serve to synergize the activity of other HDPs and thereby
offer better defense to the host frog against infections [23, 24, 139, 141, 142, 143]
as has been demonstrated for magainins from X. laevis [42]. Indeed, it has been
suggested that this antimicrobial synergism may play a role in defending Xenopus
and other species against Batrachochytrium dendrobatidis [13, 144], whose
lethal skin infections are currently responsible for a devastating global decline in
amphibian populations [145]. Clearly, weak antibacterial activity may limit the
potential of some individual magainins for development as useful anti-infectives.
However, in cases such as magainin-AM1 (Table 4.4), it has been suggested that
these weakly antibacterial peptides may serve some other primary biological
activity, allowing the possibility of development for other therapeutic applications
4.4
Other Antimicrobial Uses of Magainins
[143] or use in combination with conventional antibiotics. Consistent with this
suggestion, a very recent study showed that both magainin-AM1 and magaininAM2 possessed a potent ability to stimulate secretion of glucagon-like peptide 1
and exert direct effects on insulin secretion [146]. Insulin-releasing actions have
been demonstrated for HDPs from a number of frogs [147], including Xenopus
and Silurana [148], and it was suggested that magainin-AM1 and magainin-AM2
may possess the potential to act as therapeutic agents for the treatment of type 2
diabetes [146].
4.4
Other Antimicrobial Uses of Magainins
In comparison with their antibacterial and antifungal activities, there have been
relatively few studies on the antiviral capabilities of magainins. However, it has
been shown that magainin-2, magainin-1, and some “MSI” compounds, such
as MSI-594 (Table 4.2), are able to inhibit herpes simplex virus (HSV) [16, 122,
153–157], which causes chronic, recurrent genital infections and, globally, is
the most frequent cause of genital ulceration [158]. In the case of each of these
magainins, their antiviral activity appeared to involve disruption of the viral
envelope and it was proposed that MSI-594 and several of its analogs may be
suitable for development as therapeutically useful antiviral agents [16, 122,
153–157]. Moreover, taken with the fact that magainin-A inhibits the growth of
bacterial pathogens such as Neisseria gonorrhoeae [159], which cause sexually
transmitted infections (STIs) [160–162], these combined results led to the
suggestion that magainins may also have the potential to act as broad spectrum
agents in the prevention of these infections [16, 61, 163, 164]. Strongly supporting
this suggestion, a number of “MSI” compounds, such as MSI-420 and MSI-591
[16, 122], have been patented for topical application in the treatment of STIs
[165].
Magainins represent one of the few groups of HDPs that have been investigated
for their potential to act as contraceptive agents [61, 163, 164, 166]. Earlier studies showed that magainin-2a, magainin-G, and magainin-A possessed spermicidal
activity based on their ability to arrest sperm mobility in the case of hamsters [167]
and humans [168–170]. This effect appeared to be related to the ability of these
peptides to disrupt the outer plasma membrane of sperm cells and was enhanced
by the removal of cholesterol from these membranes, which occurs during capacitation in vivo [168–170]. It is well established that this sterol is able to modulate
the interaction of magainins with mammalian membranes and thereby plays a
major role in the preference of these peptides for prokaryotic membranes [95,
171, 172]. Further studies on magainin-A showed that the intravaginal administration of the peptide immediately prior to insemination prevented conception
through the spermicidal activity of the peptide in the cases of rats [169], rabbits
[173], and the monkey, Macaca radiata [159]. However, the intravaginal administration of another magainin analog, (Ala8,13,18)-magainin-2a, after successful
55
56
4 Magainins – A Model for Development of Eukaryotic Antimicrobial Peptides (AMPs)
insemination prevented conception in the monkey Macaca mulatta through the
peptide’s ability to inhibit the pre-implantation of embryos [174]. Magainin-2a
was also shown to exhibit toxicity to pre-implantation embryos in in vitro studies on mice [175, 176] and while the mechanisms underpinning the embryotoxic
activities of this peptide and (Ala8,13,18)-magainin-2a remained unclear, it was
suggested that they may involve the interaction of these cationic peptides with
anionic surface components in the cytoplasmic membranes of early placental trophoblasts, thereby adversely affecting their growth, differentiation, and function
[177, 178]. These various magainins and analogs were found to show low levels of
hemolysis and cytotoxicity in vaginal cells, which led to the suggestion that they
may be suitable for development as dual function microbicides, capable of reducing the risk of STIs as well as providing fertility control [61, 159, 169, 173–176].
A role in crop protection was suggested for magainins when in vitro studies
showed that magainin-2, and MSI-99, which is a close analog of MSI-78 (Table 4.2)
[179], exhibited activity against Gram-positive bacteria such as Clavibacter michiganensis, Gram-negative organisms such as Pseudomonas syringae, and fungi such
as Penicillium digitatum [31], which are phytopathogens responsible for a variety
of plant diseases [180–182]. These results led a very recent study to design de novo
analogs of magainin-2, which showed potent activity against a broad range of phytopathogens, including C. michiganensis and P. syringae, but exhibited no toxicity
toward human cells or plant protoplasts [100]. Similarly to naturally occurring
magainins [42], these analogs appeared to utilize membrane disruptive mechanisms to facilitate their antimicrobial action, and it was suggested that they may
serve in crop protection either as components in microbicidal sprays for external application or as transgene products to bolster plant immune defense systems
[100]. The results of this latter study and other work [31, 100] led to the suggestion that MSI-99 could also be used as a transgene product to endow plants with
resistance to microbial infections and diseases [183]. In response, the peptide was
expressed via the chloroplast genome in transgenic tobacco plants to give high
levels of production, which was found to provide protection against pathogenic
microbes without harmful effects to either the host plants or their chloroplasts
[125]. Since these studies, it has been shown that similar protection is provided
by transgenic expression of MSI-99 in a variety of plants and crops, including other
tobaccos [124, 126], soybeans [135], grapes [128, 130–133], bananas [124], tomatoes [123], potatoes [127, 134], and rape seed [184].
A number of investigations have suggested that magainins and analogs have
the potential to act as biocides for the treatment of microbial biofilms [59, 117,
185, 186], which are becoming increasingly problematic in areas such as the
food chain and the medical arena because of the development of resistance
to established antimicrobial strategies [187–190]. To form biofilms, microbes
attach to a surface and proliferate under favorable conditions, which leads to the
formation of a polysaccharide matrix with embedded microbial cells that are up
to a thousand times less susceptible to antibiotics and other biocides than their
planktonic counterparts [186]. In response, the anti-biofilm potential of magainins immobilized on to surfaces was investigated, which was strongly supported
4.5
Future Prospects for Magainins
when magainin-2 and several of its analogs were covalently bound to polymeric
beads and found to retain their antimicrobial activity [191, 192]. Magainin-1 has
also been shown to retain its antimicrobial activity when immobilized onto a
number of other surfaces, including titanium oxide [193], stainless steel [194],
and gold [195]. In a very recent study, the peptide was linked to polymer brushes
[196], which are assemblies of macromolecules tethered at one end to a substrate
[197]. This coating was used to immobilize magainin-1 onto a silica surface, which
then showed antimicrobial activity, and it was suggested that these functionalized
antimicrobial brushes possessed the potential for generic application to surfaces
[196]. In each of these studies, magainin-1 was found to possess potent activity
against a variety of bacterial pathogens, including Listeria ivanovii and Bacillus
cereus, [191–196], which are Gram-positive bacteria involved in food-borne
diseases that exhibit strong abilities to form biofilms [198, 199]. In combination,
these studies clearly suggest that immobilized magainins and their analogs
have the potential to inhibit or prevent microbial colonization and growth on
surfaces and thereby endow substrates with antimicrobial functionality against
biofilms [191–196]. Immobilized magainins have also been used as recognition
molecules to capture target bacteria and thereby serve as selective probes for
pathogenic organisms in a series of recently developed methodologies. In two
cases, magainin-1 was immobilized onto glass slides and the target bacteria
detected either by a labeled fluorescent dye [200] or by a target-specific antibody
[201]. In another case, this peptide was immobilized onto gold microelectrodes
to form part of an electronic biosensor, which facilitated the selective detection of
pathogenic bacteria via impedance spectroscopy [202]. Most recently, magainin2 was immobilized onto a nitrocellulose membrane and used as a recognition
molecule for bacteria in a lateral flow assay, which uses target-specific antibodies
as a detection technique [203]. In each of these studies, it was found that
immobilized magainins exhibited high sensitivity in the detection of bacterial
pathogens such as E. coli O157:H7 and Salmonella typhimurium [200–203], two
of the most prolific causes of food-borne illness worldwide [204, 205].
4.5
Future Prospects for Magainins
Magainins were one of the first families of HDPs to be discovered and are now
universally taken as the prototypes of HDPs [8]. The impact of magainins on
research into the development of HDPs is unquestionable, reflected in the fact
that a database search on magainins shows that nearly 1500 manuscripts have
been published on them over the last 25 years, with production increasing annually. The structure/function relationships gleaned from these peptides has led to
the production of a multitude of analogs and structural mimics of magainins,
many of which have antimicrobial activity and the potential for future therapeutic
development and thereby, commercialization. Indeed, although generally beyond
the scope of this chapter, it has become generally accepted that another property
57
58
4 Magainins – A Model for Development of Eukaryotic Antimicrobial Peptides (AMPs)
of many HDPs is anticancer activity [206] and many magainin-derived molecules
are being developed in this capacity [206, 207]. Studies on magainins have also
led to the identification of a number of previously unknown magainins, which
form a family of closely related, naturally occurring HDPs that can be used to
relate structure to function and thereby lead to the design of more efficient
antimicrobial agents. Interestingly, the majority of these newer magainins are far
less potent than those from X. laevis and it is a tantalizing thought as to how the
future of HDPs might have progressed if Michael Zasloff had studied a different
frog from the Xenopodinae!
Magainins also blazed the trail for the commercialization of HDPs in the form
MSI-78 and there are now over a dozen of these peptides at various stages of clinical trials [54]. The refusal of the FDA to grant marketing approval was described
by Zasloff as “a total miscarriage of the approval process” and it has since been
observed that this verdict was based on, at least, questionable ethics [84]. However, under pressure from the pharmaceutical industry, there have been recent
changes in FDA regulations governing the approval for marketing of new drugs
and antibiotics for human use [208]. Since then, it has been observed that these
changes could bring MSI-78 through to new clinical trials to achieve its full potential [209]. Indeed, it has recently been observed that HDPs are the new frontier in
the therapy of infections [60] and as shown in this chapter, magainins are on the
front line.
References
1. Fernandez de Caleya, R., Gonzalez-
2.
3.
4.
5.
Pascual, B., Garcia-Olmedo, F., and
Carbonero, P. (1972) Susceptibility of
phytopathogenic bacteria to wheat
purothionins in vitro. Appl. Microbiol.,
23, 998–1000.
Hultmark, D., Steiner, H., Rasmuson,
T., and Boman, H.G. (1980) Insect
immunity – purification and properties
of 3 inducible bactericidal proteins
from hemolymph of immunized
pupae of hyalophora-cecropia. Eur.
J. Biochem., 106, 7–16.
Wang, G., Li, X., and Wang, Z. (2009)
APD2: the updated antimicrobial peptide database and its application in
peptide design. Nucleic Acids Res., 37,
D933–D937.
Li, Y., Xiang, Q., Zhang, Q., Huang,
Y., and Su, Z. (2012) Overview on the
recent study of antimicrobial peptides:
origins, functions, relative mechanisms
and application. Peptides, 37, 207–215.
Pasupuleti, M., Schmidtchen, A., and
Malmsten, M. (2012) Antimicrobial
6.
7.
8.
9.
10.
peptides: key components of the innate
immune system. Crit. Rev. Biotechnol.,
32, 143–171.
Choi, K.-Y., Chow, L.N.Y., and
Mookherjee, N. (2012) Cationic host
defence peptides: multifaceted role in
immune modulation and inflammation.
J. Innate Immun., 4, 361–370.
Steinstraesser, L., Kraneburg, U.,
Jacobsen, F., and Al-Benna, S. (2011)
Host defense peptides and their
antimicrobial-immunomodulatory
duality. Immunobiology, 216, 322–333.
Phoenix, D.A., Dennison, S.R., and
Harris, F. (2013) Antimicrobial Peptides,
Wiley-VCH Verlag GmbH, Weinheim.
Steiner, H., Hultmark, D., Engstrom, A.,
Bennich, H., and Boman, H.G. (1981)
Sequence and specificity of 2 antibacterial proteins involved in insect
immunity. Nature, 292, 246–248.
Giovannini, M.G., Poulter, L., Gibson,
B.W., and Williams, D.H. (1987) Biosynthesis and degradation of peptides
References
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
derived from Xenopus-laevis prohormones. Biochem. J., 243, 113–120.
Zasloff, M. (1987) Magainins, a class
of antimicrobial peptides from xenopus skin – isolation, characterization
of 2 active forms, and partial cDNA
sequence of a precursor. Proc. Natl.
Acad. Sci. U.S.A., 84, 5449–5453.
Terry, A.S., Poulter, L., Williams, D.H.,
Nutkins, J.C., Giovannini, M.G., Moore,
C.H., and Gibson, B.W. (1988) The
cDNA sequence coding for prepro-PGS
(prepro-magainins) and aspects of the
processing of this prepro-polypeptide.
J. Biol. Chem., 263, 5745–5751.
Conlon, J.M., Mechkarska, M., and
King, J.D. (2012) Host-defense peptides
in skin secretions of African clawed
frogs (Xenopodinae, Pipidae). Gen.
Comp. Endocrinol., 176, 513–518.
Jacob, L. and Zasloff, M. (1994) in
Antimicrobial Peptides (eds J. Marsh
and J.A. Goode), John Wiley & Sons,
Ltd, Chichester, pp. 197–216.
Chopra, I. (1993) The magainins – antimicrobial peptides with
potential for topical application.
J. Antimicrob. Chemother., 32, 351–353.
Zasloff, M. (2002) Amphibian antimicrobial peptides, in Peptide Antibiotics
(eds C.J. Dutton, M.A. Haxell, H.A.I.
McArthur, and R.G. Wax), Marcel
Dekker Inc.
Iwahori, A., Hirota, Y., Sampe, R.,
Miyano, S., Takahashi, N., Sasatsu, M.,
Kondo, I., and Numao, N. (1997) On
the antibacterial activity of normal and
reversed magainin 2 analogs against
Helicobacter pylori. Biol. Pharm. Bull.,
20, 805–808.
Bessalle, R., Haas, H., Goria, A.,
Shalit, I., and Fridkin, M. (1992)
Augmentation of the antibacterial
activity of magainin by positive-charge
chain extension. Antimicrob. Agents
Chemother., 36, 313–317.
Chen, H.C., Brown, J.H., Morell, J.L.,
and Huang, C.M. (1988) Synthetic
magainin analogs with improved
antimicrobial activity. FEBS Lett.,
236, 462–466.
Dathe, M., Nikolenko, H., Meyer, J.,
Beyermann, M., and Bienert, M. (2001)
Optimization of the antimicrobial
21.
22.
23.
24.
25.
26.
27.
activity of magainin peptides by modification of charge. FEBS Lett., 501,
146–150.
Cuervo, J.H., Rodriguez, B., and
Houghten, R.A. (1988) The magainins:
sequence factors relevant to increased
antimicrobial activity and decreased
hemolytic activity. Pept. Res., 1, 81–86.
King, J.D., Mechkarska, M., Coquet, L.,
Leprince, J., Jouenne, T., Vaudry, H.,
Takada, K., and Conlon, J.M. (2012)
Host-defense peptides from skin secretions of the tetraploid frogs Xenopus
petersii and Xenopus pygmaeus, and
the octoploid frog Xenopus lenduensis
(Pipidae). Peptides, 33, 35–43.
Mechkarska, M., Meetani, M.,
Michalak, P., Vaksman, Z., Takada, K.,
and Conlon, J.M. (2012) Hybridization between the African clawed
frogs Xenopus laevis and Xenopus
muelleri (Pipidae) increases the multiplicity of antimicrobial peptides in
skin secretions of female offspring.
Comp. Biochem. Physiol. D: Genomics
Proteomics, 7, 285–291.
Roelants, K., Fry, B.G., Ye, L.M.,
Stijlemans, B., Brys, L., Kok, P.,
Clynen, E., Schoofs, L., Cornelis, P.,
and Bossuyt, F. (2013) Origin and functional diversification of an amphibian
defense peptide arsenal. PLoS Genet., 9,
e1003662.
Conlon, J.M., Mechkarska, M., Prajeep,
M., Sonnevend, A., Coquet, L.,
Leprince, J., Jouenne, T., Vaudry, H.,
and King, J.D. (2012) Host-defense peptides in skin secretions of the tetraploid
frog Silurana epitropicalis with potent
activity against methicillin-resistant
Staphylococcus aureus (MRSA). Peptides, 37, 113–119.
Soravia, E., Martini, G., and Zasloff,
M. (1988) Antimicrobial properties of
peptides from xenopus granular gland
secretions. FEBS Lett., 228, 337–340.
Matsuzaki, K., Murase, O., Tokuda, H.,
Funakoshi, S., Fujii, N., and Miyajima,
K. (1994) Orientational and aggregational states of Magainin 2 in
phospholipid bilayers. Biochemistry,
33, 3342–3349.
59
60
4 Magainins – A Model for Development of Eukaryotic Antimicrobial Peptides (AMPs)
28. Tachi, T., Epand, R.F., Epand, R.M.,
29.
30.
31.
32.
33.
34.
35.
36.
37.
and Matsuzaki, K. (2002) Positiondependent hydrophobicity of the
antimicrobial magainin peptide affects
the mode of peptide−lipid interactions
and selective toxicity†. Biochemistry,
41, 10723–10731.
Hallock, K.J., Lee, D.K., and
Ramamoorthy, A. (2003) MSI-78, an
analogue of the magainin antimicrobial
peptides, disrupts lipid bilayer structure
via positive curvature strain. Biophys. J.,
84, 3052–3060.
Domadia, P.N., Bhunia, A.,
Ramamoorthy, A., and Bhattacharjya,
S. (2010) Structure, interactions, and
antibacterial activities of MSI-594
derived mutant peptide MSI-594F5A in
lipopolysaccharide micelles: role of the
helical hairpin conformation in outermembrane permeabilization. J. Am.
Chem. Soc., 132, 18417–18428.
Alan, A.R. and Earle, E.D. (2002) Sensitivity of bacterial and fungal plant
pathogens to the lytic peptides, MSI-99,
magainin II, and cecropin B. Mol. Plant
Microbe Interact., 15, 701–708.
Thennarasu, S., Lee, D.K., Tan, A.,
Kari, U.P., and Ramamoorthy, A. (2005)
Antimicrobial activity and membrane
selective interactions of a synthetic
lipopeptide MSI-843. Biochim. Biophys.
Acta, Biomembr., 1711, 49–58.
Bechinger, B. (2011) Insights into the
mechanisms of action of host defence
peptides from biophysical and structural investigations. J. Pept. Sci., 17,
306–314.
Bechinger, B. and Aisenbrey, C. (2012)
The polymorphic nature of membraneactive peptides from biophysical and
structural investigations. Curr. Protein
Pept. Sci., 13, 602–610.
Dennison, S.R., Wallace, J., Harris, F.,
and Phoenix, D.A. (2005) Amphiphilic
alpha-helical antimicrobial peptides and
their structure/function relationships.
Protein Pept. Lett., 12, 31–39.
Soares, J.W. and Mello, C.M. (2004)
Antimicrobial peptides: A review of
how peptide structure impacts antimicrobial Proc. SPIE, 5271, 20–27.
Chen, Y.X., Vasil, A.I., Rehaume, L.,
Mant, C.T., Burns, J.L., Vasil, M.L.,
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
Hancock, R.E.W., and Hodges, R.S.
(2006) Comparison of biophysical and
biologic properties of alpha-helical
enantiomeric antimicrobial peptides.
Chem. Biol. Drug Des., 67, 162–173.
Yount, N.Y. and Yeaman, M.R. (2013)
Peptide antimicrobials: cell wall as a
bacterial target. Ann. N. Y. Acad. Sci.,
1277, 127–138.
Wilmes, M., Cammue, B.P.A., Sahl,
H.-G., and Thevissen, K. (2011) Antibiotic activities of host defense peptides:
more to it than lipid bilayer perturbation. Nat. Prod. Rep., 28, 1350–1358.
Chang, T.-W., Lin, Y.-M., Wang, C.-F.,
and Liao, Y.-D. (2012) Outer membrane
lipoprotein Lpp is Gram-negative bacterial cell surface receptor for cationic
antimicrobial peptides. J. Biol. Chem.,
287, 418–428.
Pushpanathan, M., Gunasekaran, P.,
and Rajendhran, J. (2013) Antimicrobial
peptides: versatile biological properties.
Int. J. Pept., 2013, 15.
Haney, E.F., Hunter, H.N., Matsuzaki,
K., and Vogel, H.J. (2009) Solution
NMR studies of amphibian antimicrobial peptides: linking structure
to function? Biochim. Biophys. Acta,
Biomembr., 1788, 1639–1655.
Brandenburg, L.-O., Merres, J.,
Albrecht, L.-J., Varoga, D., and Pufe,
T. (2012) Antimicrobial peptides:
multifunctional drugs for different
applications. Polymers, 4, 539–560.
Brogden, K.A. (2005) Antimicrobial
peptides: pore formers or metabolic
inhibitors in bacteria? Nat. Rev. Microbiol., 3, 238–250.
Sengupta, D., Leontiadou, H., Mark,
A.E., and Marrink, S.-J. (2008) Toroidal
pores formed by antimicrobial peptides
show significant disorder. Biochim.
Biophys. Acta, Biomembr., 1778,
2308–2317.
Harris, F., Dennison, S.R., and Phoenix,
D.A. (2012) Aberrant action of amyloidogenic host defense peptides: a
new paradigm to investigate neurodegenerative disorders? FASEB J., 26,
1776–1781.
Epand, R.F., Maloy, W.L.,
Ramamoorthy, A., and Epand, R.M.
(2010) Probing the “charge cluster
References
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
mechanism” in amphipathic helical cationic antimicrobial peptides.
Biochemistry, 49, 4076–4084.
Ludtke, S.J., He, K., Heller, W.T.,
Harroun, T.A., Yang, L., and Huang,
H.W. (1996) Membrane pores induced
by magainin. Biochemistry, 35,
13723–13728.
Matsuzaki, K. (1998) Magainins as
paradigm for the mode of action of
pore forming polypeptides. Biochim.
Biophys. Acta, Rev. Biomembr., 1376,
391–400.
Mihajlovic, M. and Lazaridis, T. (2012)
Charge distribution and imperfect
amphipathicity affect pore formation
by antimicrobial peptides. Biochim.
Biophys. Acta, Biomembr., 1818,
1274–1283.
He, Y. and Lazaridis, T. (2013) Activity
determinants of helical antimicrobial
peptides: a large-scale computational
study. PLoS One, 8, e66440.
He, Y., Prieto, L., and Lazaridis, T.
(2013) Modeling peptide binding to
anionic membrane pores. J. Comput.
Chem., 34, 1463–1475.
Woo, H.J. and Wallqvist, A. (2011)
Spontaneous buckling of lipid bilayer
and vesicle budding induced by
antimicrobial peptide magainin 2: a
coarse-grained simulation study. J. Phys.
Chem. B, 115, 8122–8129.
Fjell, C.D., Hiss, J.A., Hancock, R.E.W.,
and Schneider, G. (2012) Designing
antimicrobial peptides: form follows
function. Nat. Rev. Drug Discovery, 11,
37–51.
Peschel, A. and Sahl, H.G. (2006) The
co-evolution of host cationic antimicrobial peptides and microbial resistance.
Nat. Rev. Microbiol., 4, 529–536.
Anaya-Lopez, J.L., Lopez-Meza, J.E.,
and Ochoa-Zarzosa, A. (2013) Bacterial
resistance to cationic antimicrobial
peptides. Crit. Rev. Microbiol., 39,
180–195.
Shireen, T., Singh, M., Das, T., and
Mukhopadhyay, K. (2013) Differential
adaptive responses of staphylococcus
aureus to in vitro selection with different antimicrobial peptides. Antimicrob.
Agents Chemother., 57, 5134–5137.
58. Maria-Neto, S., Candido, E.d.S.,
59.
60.
61.
62.
63.
64.
65.
66.
67.
Rodrigues, D.R., de Sousa, D.A., da
Silva, E.M., Pepe de Moraes, L.M., de
Jesus Otero-Gonzalez, A., Magalhaes,
B.S., Dias, S.C., and Franco, O.L. (2012)
Deciphering the magainin resistance
process of escherichia coli strains
in light of the cytosolic proteome.
Antimicrob. Agents Chemother., 56,
1714–1724.
Conlon, J.M. and Sonnevend, A. (2011)
Clinical applications of amphibian
antimicrobial peptides. J. Med. Sci., 4,
62–72.
Zucca, M., Scutera, S., and Savoia, D.
(2011) Antimicrobial peptides: new
frontiers in the therapy of infections,
in Drug Development – A Case Study
Based Insight into Modern Strategies
(ed. C. Rundfeldt), InTech.
Zairi, A., Tangy, F., Bouassida, K., and
Hani, K. (2009) Dermaseptins and
magainins: antimicrobial peptides from
frogs’ skin-new sources for a promising spermicides microbicides-a mini
review. J. Biomed. Biotechnol., 2009,
452567.
Leontiadou, H., Mark, A.E., and
Marrink, S.J. (2006) Antimicrobial
peptides in action. J. Am. Chem. Soc.,
128, 12156–12161.
Wimley, W.C. (2010) Describing the
mechanism of antimicrobial peptide
action with the interfacial activity
model. ACS Chem. Biol., 5, 905–917.
Gregory, S.M., Pokorny, A., and
Almeida, P.F.F. (2009) Magainin 2 revisited: a test of the quantitative model
for the all-or-none permeabilization of
phospholipid vesicles. Biophys. J., 96,
116–131.
Imura, Y., Choda, N., and Matsuzaki, K.
(2008) Magainin 2 in action: distinct
modes of membrane permeabilization
in living bacterial and mammalian cells.
Biophys. J., 95, 5757–5765.
Ludtke, S., He, K., and Huang, H.
(1995) Membrane thinning caused
by magainin 2. Biochemistry, 34,
16764–16769.
Epand, R.M. and Epand, R.F. (2011)
Bacterial membrane lipids in the action
of antimicrobial agents. J. Pept. Sci., 17,
298–305.
61
62
4 Magainins – A Model for Development of Eukaryotic Antimicrobial Peptides (AMPs)
68. Matsuzaki, K. (1999) Why and how
69.
70.
71.
72.
73.
74.
75.
76.
are peptide–lipid interactions utilized
for self-defense? Magainins and tachyplesins as archetypes. Biochim. Biophys.
Acta, Biomembr., 1462, 1–10.
Zasloff, M., Martin, B., and Chen, H.C.
(1988) Antimicrobial activity of synthetic magainin peptides and several
analogs. Proc. Natl. Acad. Sci. U.S.A.,
85, 910–913.
Matsuzaki, K., Sugishita, K., Harada,
M., Fujii, N., and Miyajima, K. (1997)
Interactions of an antimicrobial peptide, magainin 2, with outer and inner
membranes of Gram-negative bacteria.
Biochim. Biophys. Acta, Biomembr.,
1327, 119–130.
Islam, K. and Hawser, S.P. (1998) MSI78 magainin pharmaceuticals. IDrugs,
1, 605–609.
Gottler, L.M. and Ramamoorthy,
A. (2009) Structure, membrane orientation, mechanism, and function
of pexiganan – A highly potent
antimicrobial peptide designed from
magainin. Biochim. Biophys. Acta,
Biomembr., 1788, 1680–1686.
Tamba, Y., Ariyama, H., Levadny, V.,
and Yamazaki, M. (2010) Kinetic pathway of antimicrobial peptide magainin
2-induced pore formation in lipid
membranes. J. Phys. Chem. B, 114,
12018–12026.
Giacometti, A., Cirioni, O., Kamysz, W.,
D’Amato, G., Silvestri, C., Licci, A.,
Nadolski, P., Riva, A., Łukasiak, J., and
Scalise, G. (2005) In vitro activity of
MSI-78 alone and in combination with
antibiotics against bacteria responsible
for bloodstream infections in neutropenic patients. Int. J. Antimicrob.
Agents, 26, 235–240.
Giacometti, A., Cirioni, O., Del Prete,
M.S., Paggi, A.M., D’Errico, M.M., and
Scalise, G. (2000) Combination studies between polycationic peptides and
clinically used antibiotics against Grampositive and Gram-negative bacteria.
Peptides, 21, 1155–1160.
Jacob, L. and Zasloff, M. (2007)
Ciba Foundation Symposium
186 – Antimicrobial Peptides, John
Wiley & Sons, Ltd, pp. 197–223.
77. Darveau, R.P., Cunningham, M.D.,
78.
79.
80.
81.
82.
83.
84.
85.
86.
Seachord, C.L., Cassianoclough, L.,
Cosand, W.L., Blake, J., and Watkins,
C.S. (1991) Beta-lactam antibiotics
potentiate magainin-2 antimicrobial
activity in vitro and in vivo. Antimicrob.
Agents Chemother., 35, 1153–1159.
Cirioni, O., Giacometti, A., Ghiselli, R.,
Mocchegiani, F., Fineo, A., Orlando,
F., Del Prete, M.S., Rocchi, M., Saba,
V., and Scalise, G. (2002) Single-dose
intraperitoneal magainins improve
survival in a Gram-negative-pathogen
septic shock rat model. Antimicrob.
Agents Chemother., 46, 101–104.
Giacometti, A., Cirioni, O., Ghiselli, R.,
Orlando, F., Kamysz, W., Rocchi, M.,
D’Amato, G., Mocchegiani, F., Silvestri,
C., Łukasiak, J., Saba, V., and Scalise, G.
(2005) Effects of pexiganan alone and
combined with betalactams in experimental endotoxic shock. Peptides, 26,
207–216.
Devocelle, M. (2012) Targeted antimicrobial peptides. Front. Immunol., 3,
309.
Arnusch, C.J., Pieters, R.J., and
Breukink, E. (2012) Enhanced membrane pore formation through
high-affinity targeted antimicrobial
peptides. PLoS One, 7, e39768.
Giuliani, A., Pirri, G., and Nicoletto,
S.F. (2007) Antimicrobial peptides:
an overview of a promising class of
therapeutics. Cent. Eur. J. Biol., 2,
1–33.
Hobizal, K.B. and Wukich, D.K. (2012)
Diabetic foot infections: current concept review. Diabet. Foot Ankle., 3,
18409.
Moore, A. (2003) The big and small
of drug discovery – Biotech versus
pharma: advantages and drawbacks
in drug development. EMBO Rep., 4,
114–117.
Nelson, E.A., O’Meara, S., Golder, S.,
Dalton, J., Craig, D., and Iglesias, C. on
behalf of the, DSG (2006) Systematic
review of antimicrobial treatments for
diabetic foot ulcers. Diabet. Med., 23,
348–359.
Lee, D.-K., Brender, J.R., Sciacca,
M.F.M., Krishnamoorthy, J., Yu, C.,
and Ramamoorthy, A. (2013) Lipid
References
87.
88.
89.
90.
91.
92.
93.
94.
composition-dependent membrane
membrane-disrupting mechanism of
fragmentation and pore-forming mechantimicrobial peptides MSI-78 and
anisms of membrane disruption by
MSI-594 derived from magainin 2 and
pexiganan (MSI-78). Biochemistry, 52,
melittin. Biophys. J., 91, 206–216.
3254–3263.
95. McHenry, A.J., Sciacca, M.F.M.,
Porcelli, F., Buck-Koehntop, B.A.,
Brender, J.R., and Ramamoorthy,
Thennarasu, S., Ramamoorthy, A.,
A. (2012) Does cholesterol suppress
and Veglia, G. (2006) Structures of
the antimicrobial peptide induced
the dimeric and monomeric variants
disruption of lipid raft containing
of magainin antimicrobial peptides
membranes? Biochim. Biophys. Acta,
(MSI-78 and MSI-594) in micelles and
Biomembr., 1818, 3019–3024.
bilayers, determined by NMR spec96. Buer, B.C., Chugh, J., Al-Hashimi,
troscopy. Biochemistry, 45, 5793–5799.
H.M., and Marsh, E.N.G. (2010) Using
Ramamoorthy, A. and Xu, J.D. (2013)
fluorine nuclear magnetic resonance
2D H-1/H-1 RFDR and NOESY NMR
to probe the interaction of membraneexperiments on a membrane-bound
active peptides with the lipid bilayer.
antimicrobial peptide under magic
Biochemistry, 49, 5760–5765.
angle spinning. J. Phys. Chem. B, 117,
97. Epand, R.F., Maloy, L., Ramamoorthy,
6693–6700.
A., and Epand, R.M. (2010) AmphiPius, J., Morrow, M.R., and Booth, V.
pathic helical cationic antimicrobial
(2012) H-2 solid-state nuclear magpeptides promote rapid formation of
netic resonance investigation of whole
crystalline states in the presence of
escherichia coli interacting with antimiphosphatidylglycerol: lipid clustering
crobial peptide MSI-78. Biochemistry,
in anionic membranes. Biophys. J., 98,
51, 118–125.
2564–2573.
Yamamoto, K., Vivekanandan, S., and
98. Epand, R.M. and Epand, R.F. (2010)
Ramamoorthy, A. (2011) Fast NMR
Biophysical analysis of membranedata acquisition from bicelles containtargeting antimicrobial peptides:
ing a membrane-associated peptide at
membrane properties and the design
natural-abundance. J. Phys. Chem. B,
of peptides specifically targeting gram115, 12448–12455.
negative bacteria, in Antimicrobial
Yang, P., Ramamoorthy, A., and Chen,
Peptides: Discovery, Design and Novel
Z. (2011) Membrane orientation of
Therapeutic Strategies, CABI.
MSI-78 measured by sum frequency
99. Suzuki, Y., Buer, B.C., Al-Hashimi,
generation vibrational spectroscopy.
H.M., and Marsh, E.N.G. (2011) Using
Langmuir, 27, 7760–7767.
fluorine nuclear magnetic resonance
Zimmerman, L.B., Worley, B.V.,
to probe changes in the structure and
Palermo, E.F., Brender, J.R., Lee, K.D.,
dynamics of membrane-active pepKuroda, K., Ramamoorthy, A., and
tides interacting with lipid bilayers.
Meyerhoff, M.E. (2011) AbsorbanceBiochemistry, 50, 5979–5987.
based assay for membrane disruption
by antimicrobial peptides and synthetic 100. Zeitler, B., Diaz, A.H., Dangel, A.,
Thellmann, M., Meyer, H., Sattler, M.,
copolymers using pyrroloquinoline
and Lindermayr, C. (2013) De-novo
quinone-loaded liposomes. Anal.
design of antimicrobial peptides for
Biochem., 411, 194–199.
plant protection. PLoS One, 8, e71687.
Bhattacharjya, S. and Ramamoorthy,
A. (2009) Multifunctional host defense 101. Maloy, W.L. and Kari, U.P. (1995)
Structure-activity studies on magainpeptides: functional and mechanistic
ins and other host-defense peptides.
insights from NMR structures of potent
Biopolymers, 37, 105–122.
antimicrobial peptides. FEBS J., 276,
102. Radzishevsky, I.S., Rotem, S., Zaknoon,
6465–6473.
F., Gaidukov, L., Dagan, A., and Mor,
Ramamoorthy, A., Thennarasu, S., Lee,
A. (2005) Effects of acyl versus aminoaD.K., Tan, A.M., and Maloy, L. (2006)
Solid-state NMR investigation of the
cyl conjugation on the properties of
63
64
4 Magainins – A Model for Development of Eukaryotic Antimicrobial Peptides (AMPs)
103.
104.
105.
106.
107.
108.
109.
110.
111.
112.
antimicrobial peptides. Antimicrob.
Agents Chemother., 49, 2412–2420.
Meng, H. and Kumar, K. (2007)
Antimicrobial activity and protease
stability of peptides containing fluorinated amino acids. J. Am. Chem. Soc.,
129, 15615–15622.
Gottler, L.M., Lee, H.-Y., Shelburne,
C.E., Ramamoorthy, A., and Marsh,
E.N.G. (2008) Using fluorous amino
acids to modulate the biological activity
of an antimicrobial peptide. ChemBioChem, 9, 370–373.
Buer, B.C. and Marsh, E.N.G. (2012)
Fluorine: a new element in protein
design. Protein Sci., 21, 453–462.
Buer, B.C., Levin, B.J., and Marsh,
E.N.G. (2013) Perfluoro-tert-butylhomoserine as a sensitive F-19 NMR
reporter for peptide-membrane interactions in solution. J. Pept. Sci., 19,
308–314.
Imura, Y., Nishida, M., and Matsuzaki,
K. (2007) Action mechanism of PEGylated magainin 2 analogue peptide.
Biochim. Biophys. Acta, Biomembr.,
1768, 2578–2585.
Porter, E.A., Weisblum, B., and
Gellman, S.H. (2002) Mimicry of
host-defense peptides by unnatural
oligomers: antimicrobial beta-peptides.
J. Am. Chem. Soc., 124, 7324–7330.
Chongsiriwatana, N.P., Patch, J.A.,
Czyzewski, A.M., Dohm, M.T., Ivankin,
A., Gidalevitz, D., Zuckermann, R.N.,
and Barron, A.E. (2008) Peptoids that
mimic the structure, function, and
mechanism of helical antimicrobial
peptides. Proc. Natl. Acad. Sci. U.S.A.,
105, 2794–2799.
Dohm, M.T., Kapoor, R., and Barron,
A.E. (2011) Peptoids: bio-inspired polymers as potential pharmaceuticals.
Curr. Pharm. Des., 17, 2732–2747.
Tang, H.Z., Doerksen, R.J., and Tew,
G.N. (2005) Synthesis of urea oligomers
and their antibacterial activity. Chem.
Commun., 1537–1539.
Violette, A., Fournel, S., Lamour, K.,
Chaloin, O., Frisch, B., Briand, J.P.,
Monteil, H., and Guichard, G. (2006)
Mimicking helical antibacterial peptides
with nonpeptidic folding oligomers.
Chem. Biol., 13, 531–538.
113. Thaker, H.D., Som, A., Ayaz, F., Lui,
114.
115.
116.
117.
118.
119.
120.
121.
D., Pan, W., Scott, R.W., Anguita, J.,
and Tew, G.N. (2012) Synthetic mimics of antimicrobial peptides with
immunomodulatory responses. J. Am.
Chem. Soc., 134, 11088–11091.
Thaker, H.D., Sgolastra, F., Clements,
D., Scott, R.W., and Tew, G.N. (2011)
Synthetic mimics of antimicrobial peptides from triaryl scaffolds. J. Med.
Chem., 54, 2241–2254.
Majoros, I.J., Williams, C.R., Becker,
A.C., and Baker, J.R. Jr., (2009)
Surface interaction and behavior
of poly(amidoamine) dendrimers:
deformability and lipid bilayer disruption. J. Comput. Theor. Nanosci., 6,
1430–1436.
Eren, T., Som, A., Rennie, J.R., Nelson,
C.F., Urgina, Y., Nusslein, K., Coughlin,
E.B., and Tew, G.N. (2008) Antibacterial and hemolytic activities of
quaternary pyridinium functionalized
polynorbornenes. Macromol. Chem.
Phys., 209, 516–524.
Beckloff, N., Laube, D., Castro, T.,
Furgang, D., Park, S., Perlin, D.,
Clements, D., Tang, H., Scott, R.W.,
Tew, G.N., and Diamond, G. (2007)
Activity of an antimicrobial peptide
mimetic against planktonic and biofilm
cultures of oral pathogens. Antimicrob.
Agents Chemother., 51, 4125–4132.
Tew, G.N., Clements, D., Tang, H.,
Arnt, L., and Scott, R.W. (2006)
Antimicrobial activity of an abiotic host defense peptide mimic.
Biochim. Biophys. Acta, Biomembr.,
1758, 1387–1392.
Genco, C.A., Maloy, W.L., Kari, U.P.,
and Motley, M. (2003) Antimicrobial activity of magainin analogues
against anaerobic oral pathogens. Int. J.
Antimicrob. Agents, 21, 75–78.
Bhunia, A., Ramamoorthy, A., and
Bhattacharjya, S. (2009) Helical hairpin
structure of a potent antimicrobial
peptide MSI-594 in lipopolysaccharide
micelles by NMR spectroscopy. Chem.
Eur. J., 15, 2036–2040.
Chen, X. and Chen, Z. (2006) SFG
studies on interactions between antimicrobial peptides and supported lipid
References
122.
123.
124.
125.
126.
127.
128.
129.
bilayers. Biochim. Biophys. Acta,
Biomembr., 1758, 1257–1273.
Egal, M., Conrad, M., MacDonald,
D.L., Maloy, W.L., Motley, M., and
Genco, C.A. (1999) Antiviral effects of
synthetic membrane-active peptides
on herpes simplex virus, type 1. Int. J.
Antimicrob. Agents, 13, 57–60.
Alan, A.R., Blowers, A., and Earle, E.D.
(2004) Expression of a magainin-type
antimicrobial peptide gene (MSI-99) in
tomato enhances resistance to bacterial speck disease. Plant Cell Rep., 22,
388–396.
Chakrabarti, A., Ganapathi, T.R.,
Mukherjee, P.K., and Bapat, V.A. (2003)
MSI-99, a magainin analogue, imparts
enhanced disease resistance in transgenic tobacco and banana. Planta, 216,
587–596.
DeGray, G., Rajasekaran, K., Smith,
F., Sanford, J., and Daniell, H. (2001)
Expression of an antimicrobial peptide
via the chloroplast genome to control
phytopathogenic bacteria and fungi.
Plant Physiol., 127, 852–862.
Ganapathi, T.R., Ghosh, S.B., Laxmi,
N.H.S., and Bapat, V.A. (2007) Expression of an antimicrobial peptide
(MSI-99) confers enhanced resistance to Aspergillus niger in transgenic
potato. Indian J. Biotechnol., 6, 63–67.
Hong, Y.-b., Liu, S.-p., Zhu, Y.-p.,
Xie, C., Jue, D.-w., Chen, M., Kaleri,
H.A., and Yang, Q. (2013) Expression
of the MSI-99m gene in transgenic
potato plants confers resistance to
phytophthora infestans and ralstonia
solanacearum. Plant Mol. Biol. Rep., 31,
418–424.
Kikkert, J.R., Vidal, J.R., Wallace, P.G.,
Reisch, B.I., Garcia-Zitter, S., Wilcox,
W.F., Gadoury, D.M., Seem, R.C., and
Burr, T.J. (2005) Disease resistance
analyses of transgenic grapevines that
contain endochitinase or antimicrobial peptide genes. Proceedings of the
Seventh International Symposium on
Grapevine Physiology and Biotechnology (ed. L.E. Williams), pp. 493–498.
Thennarasu, S., Huang, R., Lee, D.K., Yang, P., Maloy, L., Chen, Z.,
and Ramamoorthy, A. (2010) Limiting an antimicrobial peptide to the
130.
131.
132.
133.
134.
135.
136.
lipid−water interface enhances its
bacterial membrane selectivity: a case
study of MSI-367. Biochemistry, 49,
10595–10605.
Reisch, B., Kikkert, J., Vidal, J., Ali, G.S.,
Gadoury, D., Seem, R., and Wallace, P.
(2003) Genetic transformation of Vitis
vinifera to improve disease resistance.
Proceedings of the 8th International
Conference on Grape Genetics and
Breeding, vols. 1 and 2 (eds E. Hajdu
and E. Borbas), pp. 303–308.
Kikkert, J.R., Vidal, J.R., Wallace,
P.G., Garcia-Zitter, S., Wilcox, W.F.,
Gadoury, D.M., Seem, R.C., Burr, T.J.,
Rosenfield, C.L., Samuelian, S., and
Reisch, B.I. (2009) Disease resistance
analyses of transgenic grapevines that
contain endochitinase or antimicrobial
peptide genes. IX International Conference on Grape Genetics and Breeding
(eds E. Peterlunger, G. DiGaspero, and
G. Cipriani), pp. 379-383.
Rosenfield, C.-L., Samuelian, S., Vidal,
J.R., and Reisch, B.I. (2010) Transgenic
disease resistance in vitis vinifera:
potential use and screening of antimicrobial peptides. Am. J. Enol. Vitic., 61,
348–357.
Vidal, J.R., Kikkert, J.R., Malnoy, M.A.,
Wallace, P.G., Barnard, J., and Reisch,
B.I. (2006) Evaluation of transgenic
‘Chardonnay’ (Vitis vinifera) containing
magainin genes for resistance to crown
gall and powdery mildew. Transgenic
Res., 15, 69–82.
O’Callaghan, M., Gerard, E.M.,
Waipara, N.W., Young, S.D., Glare, T.R.,
Barrell, P.J., and Conner, A.J. (2004)
Microbial communities of Solanum
tuberosum and magainin-producing
transgenic lines. Plant Soil, 266, 47–56.
Nanna, R. (2010) Expression of antimicrobial peptide gene (magainin) in
soybean. In Vitro Cell. Dev. Biol.-Anim.,
46, S50.
McLane, M.P., McDonald, D.L., Zasloff,
M.A., and Jacob, L.S. (1995) MSI-843,
a magainin for the treatment of Pseudomonas aeruginosa infection in cystic
fibrosis (CF) patients. Abstr. Intersci.
Conf. Antimicrob. Agents Chemother.,
35, 135.
65
66
4 Magainins – A Model for Development of Eukaryotic Antimicrobial Peptides (AMPs)
137. Trullas, C.C.R. and Krutmann, J. (2010)
138.
139.
140.
141.
142.
143.
144.
Use of compositions comprising urea
for treating microbial infections. EP
2241312 A1.
Lichter, J., Lebel, C., Piu, F., and Harris,
J.P. (2011) Compositions and methods for the treatment of sinonasal
disorders. WO Patent 2011049960 A2.
Mechkarska, M., Prajeep, M.,
Leprince, J., Vaudry, H., Meetani,
M.A., Evans, B.J., and Conlon, J.M.
(2013) A comparison of host-defense
peptides in skin secretions of female
Xenopus laevis × Xenopus borealis
and X. borealis × X. laevis F1 hybrids.
Peptides, 45, 1–8.
King, J.D., Mechkarska, M., Meetani,
M.A., and Conlon, J.M. (2013) Peptidomic analysis of skin secretions
provides insight into the taxonomic
status of the African clawed frogs
Xenopus victorianus and Xenopus
laevis sudanensis (Pipidae). Comp.
Biochem. Physiol. D: Genomics Proteomics, 8, 250–254.
Mechkarska, M., Ahmed, E., Coquet,
L., Leprince, J., Jouenne, T., Vaudry,
H., King, J.D., and Conlon, J.M. (2011)
Peptidomic analysis of skin secretions demonstrates that the allopatric
populations of Xenopus muelleri (Pipidae) are not conspecific. Peptides, 32,
1502–1508.
Mechkarska, M., Ahmed, E., Coquet,
L., Leprince, J., Jouenne, T., Vaudry,
H., King, J.D., Takada, K., and Conlon,
J.M. (2011) Genome duplications within
the Xenopodinae do not increase the
multiplicity of antimicrobial peptides
in Silurana paratropicalis and Xenopus
andrei skin secretions. Comp. Biochem.
Physiol. D: Genomics Proteomics, 6,
206–212.
Conlon, J.M., Al-Ghaferi, N., Ahmed,
E., Meetani, M.A., Leprince, J., and
Nielsen, P.F. (2010) Orthologs of magainin, PGLa, procaerulein-derived, and
proxenopsin-derived peptides from
skin secretions of the octoploid frog
Xenopus amieti (Pipidae). Peptides, 31,
989–994.
Rollins-Smith, L.A., Ramsey, J.P., Pask,
J.D., Reinert, L.K., and Woodhams, D.C.
(2011) Amphibian immune defenses
145.
146.
147.
148.
149.
150.
151.
152.
against chytridiomycosis: impacts of
changing environments. Integr. Comp.
Biol., 51, 552–562.
Voyles, J., Rosenblum, E.B., and Berger,
L. (2011) Interactions between Batrachochytrium dendrobatidis and its
amphibian hosts: a review of pathogenesis and immunity. Microbes Infect., 13,
25–32.
Ojo, O.O., Conlon, J.M., Flatt, P.R., and
Abdel-Wahab, Y.H.A. (2013) Frog skin
peptides (tigerinin-1R, magainin-AM1,AM2, CPF-AM1, and PGla-AM1)
stimulate secretion of glucagon-like
peptide 1 (GLP-1) by GLUTag cells.
Biochem. Biophys. Res. Commun., 431,
14–18.
Ojo, O.O., Flatt, P.R., Abdel-Wahab,
Y.H.A., and Conlon, J.M. (2013) in
Handbook of Biologically Active Peptides (ed. A.J. Kastin), Elsevier, San
Diego, CA, pp. 364–370.
Srinivasan, D., Mechkarska, M.,
Abdel-Wahab, Y.H.A., Flatt, P.R., and
Conlon, J.M. (2013) Caerulein precursor fragment (CPF) peptides from
the skin secretions of Xenopus laevis
and Silurana epitropicalis are potent
insulin-releasing agents. Biochimie, 95,
429–435.
Rackelmann-Silber, K. (2012) Development of therapeutic agents handbook.
Edited by Shayne Cox Gad. ChemMedChem, 7, 1504–1505.
Wieprecht, T., Dathe, M., Beyermann,
M., Krause, E., Maloy, W.L.,
MacDonald, D.L., and Bienert, M.
(1997) Peptide hydrophobicity controls
the activity and selectivity of magainin
2 amide in interaction with membranes. Biochemistry, 36, 6124–6132.
Mechkarska, M., Ahmed, E., Coquet,
L., Leprince, J., Jouenne, T., Vaudry,
H., King, J.D., and Conlon, J.M. (2010)
Antimicrobial peptides with therapeutic
potential from skin secretions of the
Marsabit clawed frog Xenopus borealis
(Pipidae). Comp. Biochem. Physiol. C:
Toxicol. Pharmacol., 152, 467–472.
Conlon, J.M., Mechkarska, M., Ahmed,
E., Leprince, J., Vaudry, H., King, J.D.,
and Takada, K. (2011) Purification and
properties of antimicrobial peptides
from skin secretions of the Eritrea
References
153.
154.
155.
156.
157.
158.
159.
160.
161.
162.
clawed frog Xenopus clivii (Pipidae).
Comp. Biochem. Physiol. C: Toxicol.
Pharmacol., 153, 350–354.
Albiol Matanic, V.C. and Castilla, V.
(2004) Antiviral activity of antimicrobial cationic peptides against Junin
virus and herpes simplex virus. Int. J.
Antimicrob. Agents, 23, 382–389.
Aboudy, Y., Mendelson, E., Shalit, I.,
Bessalle, R., and Fridkin, M. (1994)
Activity of 2 synthetic amphiphilic
peptides and magainin-2 against
herpes-simplex virus type-1 and type-2.
Int. J. Pept. Protein Res., 43, 573–582.
Yasin, B., Pang, M., Turner, J.S., Cho,
Y., Dinh, N.N., Waring, A.J., Lehrer,
R.I., and Wagar, E.A. (2000) Evaluation
of the inactivation of infectious herpes
simplex virus by host-defense peptides.
Eur. J. Clin. Microbiol. Infect. Dis., 19,
187–194.
Jenssen, H. (2009) Therapeutic
approaches using host defence peptides to tackle herpes virus infections.
Viruses-Basel, 1, 939–964.
Carriel-Gomes, M.C., Kratz, J.M.,
Barracco, M.A., Bachere, E., Barardi,
C.R.M., and Simoes, C.M.O. (2007) In
vitro antiviral activity of antimicrobial
peptides against herpes simplex virus 1,
adenovirus, and rotavirus. Mem. Inst.
Oswaldo Cruz, 102, 469–472.
Azwa, A. and Barton, S.E. (2009)
Aspects of herpes simplex virus: a
clinical review. J. Fam. Plann. Reprod.
Health Care, 35, 237–242.
Clara, A., Manjramkar, D.D., and
Reddy, V.K. (2004) Preclinical evaluation of magainin-A as a contraceptive
antimicrobial agent. Fertil. Steril., 81,
1357–1365.
Miller, K.E. (2006) Diagnosis and
treatment of Neisseria gonorrhoeae
infections. Am. Fam. Physician, 73,
1779–1784.
Whitlow, C.B. (2004) Bacterial sexually
transmitted diseases. Clin. Colon Rectal
Surg., 17, 209–214.
Wasik, M. and Kachlic, M.D. (2009) A
review of common sexually transmitted
diseases. Formulary, 44, 78–85.
163. Reddy, K.V.R., Yedery, R.D., and
164.
165.
166.
167.
168.
169.
170.
171.
172.
173.
Aranha, C. (2004) Antimicrobial peptides: premises and promises. Int. J.
Antimicrob. Agents, 24, 536–547.
Yedery, R.D. and Reddy, K.V.R. (2005)
Antimicrobial peptides as microbicidal contraceptives: prophecies for
prophylactics – a mini review. Eur. J.
Contracept. Reprod. Health Care, 10,
32–42.
Bedi, G., Jacob, J., Williams, T., and
Zasloff, M. (1996) Method for inhibiting sexually transmitted diseases using
magaining antimicrobials or squalamine
compounds. WO 1996008270 A3.
Shah, H.C., Tatke, P., and Singh, K.K.
(2008) Spermicidal agents. Drug Discovery Ther., 2, 200–210.
de Waal, A., Gomes, A.V., Mensink,
A., Grootegoed, J.A., and Westerhoff,
H.V. (1991) Magainins affect respiratory
control, membrane potential and motility of hamster spermatozoa. FEBS Lett.,
293, 219–223.
Edelstein, M.C., Gretz, J.E., Bauer, T.J.,
Fulgham, D.L., Alexander, N.J., and
Archer, D.F. (1991) Studies on the in
vitro spermicidal activity of synthetic
magainins. Fertil. Steril., 55, 647–649.
Reddy, K.V.R., Shahani, S.K., and
Meherji, P.K. (1996) Spermicidal activity of Magainins: in vitro and in vivo
studies. Contraception, 53, 205–210.
Wojcik, C., Sawicki, W., Marianowski,
P., Benchaib, M., Czyba, J.C., and
Guerin, J.F. (2000) Cyclodextrin
enhances spermicidal effects of
magainin-2-amide. Contraception,
62, 99–103.
Brender, J.R., McHenry, A.J., and
Ramamoorthy, A. (2012) Does cholesterol play a role in the bacterial
selectivity of antimicrobial peptides?
Front. Immunol., 3, 195.
Matsuzaki, K., Sugishita, K., Fujii, N.,
and Miyajima, K. (1995) Molecularbasis for membrane selectivity of an
antimicrobial peptide, magainin-2.
Biochemistry, 34, 3423–3429.
Reddy, V.R.K. and Manjramkar, D.D.
(2000) Evaluation of the antifertility
effect of magainin-A in rabbits: in vitro
and in vivo studies. Fertil. Steril., 73,
353–358.
67
68
4 Magainins – A Model for Development of Eukaryotic Antimicrobial Peptides (AMPs)
174. Dhawan, L., Ghosh, D., Lalitkumar,
175.
176.
177.
178.
179.
180.
181.
182.
P.G.L., Sharma, D.N., Lasley, B.L.,
Overstreet, J.W., and Sengupta, J.
(2000) Anti-nidatory effect of vaginally
administered (Ala8,13,18)-magainin II
amide in the rhesus monkey. Contraception, 62, 39–43.
Mystkowska, E.T., Niemierko, A.,
Komar, A., and Sawicki, W. (2001)
Embryotoxicity of magainin-2-amide
and its enhancement by cyclodextrin,
albumin, hydrogen peroxide and acidification. Hum. Reprod., 16, 1457–1463.
Sawicki, W. and Mystkowska, E.T.
(1999) Contraceptive potential of peptide antibiotics. Lancet, 353, 464–465.
Ghosh, D., Dhawan, L., Lalitkumar,
P.G.L., Wong, V., Rosario, J.F.,
Hendrickx, A.G., Lasley, B.L.,
Overstreet, J.W., and Sengupta, J.
(2001) Effect of vaginally administered
(Ala8,13,18)-magainin II amide on
the morphology of implantation stage
endometrium in the rhesus monkey
(Macaca mulatta). Contraception, 63,
335–342.
Sengupta, J., Khan, M.A., Huppertz, B.,
and Ghosh, D. (2011) In-vitro effects of
the antimicrobial peptide Ala8,13,18magainin II amide on isolated human
first trimester villous trophoblast cells.
Reprod. Biol. Endocrinol., 9, 49.
Maloy, W. L., Macdonald, D., and
Brasseur, M. (1991) Design of broadspectrum antibiotic and host defense
peptides based on magainin and related
peptides.
Eichenlaub, R. and Gartemann, K.H.
(2011) The Clavibacter michiganensis
Subspecies: Molecular Investigation
of Gram-Positive Bacterial Plant
Pathogens. Annu. Rev. Phytopathol.,
49, 445–464.
Morris, C.E., Monteil, C.L., and
Berge, O. (2013) The Life History of
Pseudomonas syringae: Linking Agriculture to Earth System Processes.
Annu. Rev. Phytopathol., 51, 85–104.
Marcet-Houben, M., Ballester, A.-R., de
la Fuente, B., Harries, E., Marcos, J.F.,
Gonzalez-Candelas, L., and Gabaldon,
T. (2012) Genome sequence of the
183.
184.
185.
186.
187.
188.
189.
190.
191.
192.
necrotrophic fungus Penicillium digitatum, the main postharvest pathogen of
citrus. BMC Genomics, 13, 646.
Keymanesh, K., Soltani, S., and Sardari,
S. (2009) Application of antimicrobial
peptides in agriculture and food industry. World J. Microbiol. Biotechnol., 25,
933–944.
Sang, X., Jue, D., Yang, L., Bai, X.,
Chen, M., and Yang, Q. (2013) Genetic
transformation of brassica napus with
msi-99m gene increases resistance in
transgenic plants to sclerotinia sclerotiorum in. Mol. Plant Breed., 4 (30),
247–253.
Kapoor, R., Wadman, M.W., Dohm,
M.T., Czyzewski, A.M., Spormann,
A.M., and Barron, A.E. (2011) Antimicrobial peptoids are effective against
pseudomonas aeruginosa biofilms.
Antimicrob. Agents Chemother., 55,
3054–3057.
Siedenbiedel, F. and Tiller, J.C. (2012)
Antimicrobial polymers in solution and
on surfaces: overview and functional
principles. Polymers, 4, 46–71.
Srey, S., Jahid, I.K., and Ha, S.-D.
(2013) Biofilm formation in food industries: a food safety concern. Food
Control, 31, 572–585.
Lynch, A.S. and Robertson, G.T. (2008)
Bacterial and fungal biofilm infections.
Annu. Rev. Med., 59, 415–428.
Bridier, A., Briandet, R., Thomas,
V., and Dubois-Brissonnet, F. (2011)
Resistance of bacterial biofilms to disinfectants: a review. Biofouling, 27,
1017–1032.
Kostakioti, M., Hadjifrangiskou, M., and
Hultgren, S.J. (2013) Bacterial biofilms:
development, dispersal, and therapeutic
strategies in the dawn of the postantibiotic era. Cold Spring Harbor Perspect.
Med., 3, a010306.
Haynie, S.L., Crum, G.A., and Doele,
B.A. (1995) Antimicrobial activities of
amphiphilic peptides covalently bonded
to a water-insoluble resin. Antimicrob.
Agents Chemother., 39, 301–307.
Bagheri, M., Beyermann, M., and
Dathe, M. (2009) Immobilization
reduces the activity of surface-bound
cationic antimicrobial peptides with no
influence upon the activity spectrum.
References
193.
194.
195.
196.
197.
198.
199.
200.
Antimicrob. Agents Chemother., 53,
1132–1141.
Peyre, J., Humblot, V., Methivier,
C., Berjeaud, J.M., and Pradier,
C.M. (2012) Co-grafting of amino
poly(ethylene glycol) and Magainin I on
a TiO2 surface: tests of antifouling and
antibacterial activities. J. Phys. Chem. B,
116, 13839–13847.
Hequet, A., Humblot, V., Berjeaud,
J.M., and Pradier, C.M. (2011) Optimized grafting of antimicrobial
peptides on stainless steel surface
and biofilm resistance tests. Colloids
Surf., B, 84, 301–309.
Humblot, V., Yala, J.F., Thebault, P.,
Boukerma, K., Hequet, A., Berjeaud,
J.M., and Pradier, C.M. (2009) The
antibacterial activity of Magainin I
immobilized onto mixed thiols SelfAssembled Monolayers. Biomaterials,
30, 3503–3512.
Glinel, K., Jonas, A.M., Jouenne, T.,
Leprince, J., Galas, L., and Huck, W.T.S.
(2009) Antibacterial and antifouling
polymer brushes incorporating antimicrobial peptide. Bioconjugate Chem., 20,
71–77.
Azzaroni, O. (2012) Polymer brushes
here, there, and everywhere: recent
advances in their practical applications
and emerging opportunities in multiple
research fields. J. Polym. Sci., Part A:
Polym. Chem., 50, 3225–3258.
Stenfors Arnesen, L.P., Fagerlund, A.,
and Granum, P.E. (2008) From soil to
gut: bacillus cereus and its food poisoning toxins. FEMS Microbiol. Rev.,
32, 579–606.
Roberts, A.J. and Wiedmann, M. (2003)
Pathogen, host and environmental factors contributing to the pathogenesis
of listeriosis. Cell. Mol. Life Sci., 60,
904–918.
Kulagina, N.V., Lassman, M.E., Ligler,
F.S., and Taitt, C.R. (2005) Antimicrobial peptides for detection of bacteria
201.
202.
203.
204.
205.
206.
207.
208.
209.
in biosensor assays. Anal. Chem., 77,
6504–6508.
Kulagina, N.V., Shaffer, K.M., Anderson,
G.P., Ligler, F.S., and Taitt, C.R. (2006)
Antimicrobial peptide-based array
for Escherichia coli and Salmonella
screening. Anal. Chim. Acta, 575,
9–15.
Mannoor, M.S., Zhang, S.Y., Link, A.J.,
and McAlpine, M.C. (2010) Electrical
detection of pathogenic bacteria via
immobilized antimicrobial peptides.
Proc. Natl. Acad. Sci. U.S.A., 107,
19207–19212.
Yonekita, T., Ohtsuki, R., Hojo, E.,
Morishita, N., Matsumoto, T., Aizawa,
T., and Morimatsu, F. (2013) Development of a novel multiplex lateral flow
assay using an antimicrobial peptide for
the detection of Shiga toxin-producing
Escherichia coli. J. Microbiol. Methods,
93, 251–256.
Pexara, A., Angelidis, D., and Govaris,
A. (2012) Shiga toxin-producing
Escherichia coli (STEC) food-borne
outbreaks. J. Hell. Vet. Med. Soc., 63,
45–53.
Hur, J., Jawale, C., and Lee, J.H. (2012)
Antimicrobial resistance of Salmonella
isolated from food animals: a review.
Food Res. Int., 45, 819–830.
Gaspar, D., Veiga, A.S., and Castanho,
M.A.R.B. (2013) From antimicrobial to
anticancer peptides. A review. Front.
Microbiol., 4, 294.
Harris, F., Dennison, S.R., Singh, J., and
Phoenix, D.A. (2013) On the selectivity
and efficacy of defense peptides with
respect to cancer cells. Med. Res. Rev.,
33, 190–234.
Ledford, H. (2012) FDA under pressure
to relax drug rules. Nature, 492, 19.
Fox, J.L. (2013) Antimicrobial peptides
stage a comeback. Nat. Biotechnol., 31,
379–382.
69
71
5
Antimicrobial Peptides from Prokaryotes
Maryam Hassan, Morten Kjos, Ingolf F. Nes, Dzung B. Diep, and Farzaneh Lotfipour
5.1
Introduction
A significant part of the disease history of mankind concerns the never-ending
fight against microbial pathogens. Invasion by microorganisms has killed millions of human beings, but with the discovery of antibiotics and vaccines the hope
of eradicating bacteria-derived diseases developed. However, due to the recent
alarming rise of antibiotic resistance among pathogens, there is an emerging need
to develop new antimicrobial agents [1, 2].
From a clinical point of view, it is a desirable approach to develop narrowspectrum antibacterial agents that are active against the defined pathogens
associated with microbial infections and the antimicrobial compounds should not
affect the positive, commensal, and natural microflora of the patient [3]. Promising candidates for such treatment are found among the bacteriocins produced by
bacteria and they may provide the key to solve some of the shortcomings of the
classical antibiotics. Antimicrobial peptides (AMPs) from prokaryotic bacteria
have long been seen as potential compounds for use as antimicrobials both in clinical treatments and in food applications, but with a few notable exceptions, their
actual application has so far been very limited. Now, with the alarming increase of
antibiotic resistance in pathogenic bacteria worldwide, AMPs from prokaryotes
should no longer be ignored as an option to combat this problem [4, 5].
Bacteriocins are gene-encoded, ribosomally synthesized peptides with antimicrobial activity directed mostly against closely related bacteria [6]. Although bacteriocins are ribosomally produced, they may be subjected to post-translational
modifications (PTMs) of various kinds and such modifications are often important for stabilization, translocation, and the antimicrobial activity of the peptides.
Bacteriocins have several characteristics that make them excellent candidates for
becoming a new generation of antimicrobials: (i) they are highly potent and inhibit
target cells at nanomolar concentrations, (ii) some AMPs have a remarkably narrow spectrum of inhibition, which makes it possible to develop pathogen-specific
drugs, (iii) yet other AMPs have broader activity spectra (targeting several different bacterial genera) and can be used in more general approaches, (iv) the AMPs
Novel Antimicrobial Agents and Strategies, First Edition.
Edited by David A. Phoenix, Frederick Harris and Sarah R. Dennison.
© 2015 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2015 by Wiley-VCH Verlag GmbH & Co. KGaA.
72
5 Antimicrobial Peptides from Prokaryotes
are often highly stable and are resistant to many proteases, high temperature, and
pH variation, and (iv) the AMPs are amenable to bioengineering, which makes
it possible to create and develop novel variants of the antimicrobials by peptide
synthesis or DNA recombinant technology.
Bacteriocins display great diversity in amino acid sequence, structures, and
targets. Owing to this enormous diversity, classification of bacteriocins is not
straightforward and is still an unsettled issue as demonstrated by different classification schemes suggested over the years [6–8]. In general, bacteriocins from
Gram-positive bacteria have been divided into two classes: Class I with peptides
containing PTMs (lanthibiotics) and Class II containing non-modified bacteriocins. Gram-negative peptide bacteriocins are often referred to as microcins
and these have been further subclassified according to their PTM. Recently, a
comprehensive nomenclature and classification of peptides containing PTMs
(RiPPs; ribosomally synthesized, post-translationally modified peptides) was
proposed, where peptides from both Gram-positive and Gram-negative bacteria
are combined in the same classification scheme. In this scheme, the modified
peptides (Class I) are subdivided into 11 groups [8].
Biosynthesis of bacteriocins generally requires four types of genes, and these are
often found in the same genetic locus: (i) The structural gene(s) encoding the bacteriocin itself – in most cases, the gene encodes a pre-form of the peptide, which
consists of an N-terminal leader sequence and the mature peptide. The leader
sequence, which renders the intracellular peptide inactive and is important for
proper modification and transport across the cell membrane, is cleaved off when
the peptide is exported. It should be mentioned that some Class II bacteriocins
from Gram-positive bacteria are encoded without leader sequences [9, 10]. In the
case of the so-called two-peptide bacteriocins, AMPs whose activity requires the
combined action of both peptides, two structural genes are found in the same
operon and both encode a bacteriocin with an N-terminal leader sequence. (ii)
An immunity system – the AMP producer needs to encode an immunity mechanism to avoid being killed by its own AMP. The immunity protein(s) are often
encoded by the same operon as the AMP structural gene to ensure co-regulation.
The mechanism of immunity is highly different between AMPs and can include
both multiple proteins (including ABC transporters) to obtain optimal resistance
such as in nisin and other lantibiotics or by single, small proteins as often seen
among many Class II bacteriocins. (iii) A transport system – in order to be active
outside the cell, the AMPs need to be exported across the cell envelope. Some
bacteriocins may utilize general secretory pathways for externalization, but often
dedicated transport systems, known as ABC transporters, are encoded in the AMP
locus. In many cases, cleavage of the leader peptide to produce the mature peptide
occurs concomitantly with transport mediated by the ABC transporter. (iv) Genes
responsible for PTMs – many bacteriocins, such as the lantibiotics and most of the
microcins, contain PTMs. Such modifications are inferred by dedicated enzymes
that act on the ribosomally synthesized peptide to produce different PTMs.
The mode of action of bacteriocins varies greatly between peptides, and is often
different from classical antibiotics. Generally, the bacteriocins can kill sensitive
5.2
Bacteriocins
bacteria by two different methods: (i) they may target the cell envelope by forming
pores or inhibiting cell wall biosynthesis or (ii) they may act intracellularly to
inhibit essential enzymatic/metabolic processes that involve inhibition of processes such as DNA synthesis, transcription, and translation. Interestingly, the
latter mechanisms are found only among Gram-negative bacteriocins so far. Conversely, both Gram-negative and Gram-positive bacteriocins are targeting the cell
envelope and the majority of bacteriocins do this by permeabilizing the membrane
or by affecting the cell wall. Such activities cause disruption of the integrity of the
cell free flow of small molecules, dissipation of the proton motive force, and eventually cell death. In order to permeabilize the membrane by forming pores, disrupt
the membrane integrity, or irreversibly open transport channels, most probably
the bacteriocins need to interact with a target on the cell envelope. The initial
attraction of bacteriocins to the cell membrane is probably partly governed by
electrostatic interactions between the positively charged peptides and the anionic
lipids in the bacterial membranes [11], but a number of membrane-targeted bacteriocins interact specifically with a receptor of the sensitive cell, and such specific
interactions are necessary for potent activity of the bacteriocins. Different types of
molecules, including the peptidoglycan precursor lipid II and membrane proteins
involved in sugar transport have been shown to function as receptors for different
types of bacteriocins [6].
In the following, we first describe some of the most important classes of
prokaryotic AMPs, namely, the microcins of Gram-negative bacteria and the
lanthibiotics (Class I) and the Class II bacteriocins of Gram-positive bacteria.
Then we discuss the (potential) applications of these peptides in different fields.
Finally, we present some approaches that are being used to discover and develop
new AMPs.
5.2
Bacteriocins
5.2.1
Microcins – Peptide Bacteriocins from Gram-Negative Bacteria
The small, peptide bacteriocins from Gram-negative Enterobacteria, such as
Escherichia, Klebsiella, and Salmonella, are called microcins. These peptides
are <10 kDa (typically under 100 amino acids (aa)) and generally have a narrow
inhibitory spectrum; thus, they are active only against closely related species
within the Enterobacteria. Microcins are generally hydrophobic and are highly
stable to high temperatures, extreme pH, and proteases. One example of a
highly stable peptide is microcin J25 (MccJ25), a 21 aa long peptide that forms
a remarkably stable threaded lasso structure [12, 13], in which the C-terminal
tail is threaded into a beta-lactam ring. Microcins constitute a heterogeneous
class of peptides, and they are traditionally classified into two groups according
to their size and PTMs [14] (Table 5.1). Class I consist of the smallest peptides
73
74
5 Antimicrobial Peptides from Prokaryotes
Table 5.1 Classification of AMPs.
Subgroup
Microcins
Lantibiotic-containing
bacteriocins (Class I)
I
Examples
Small (<5 kDa),
extensive PTM
Microcin C
IIa
Larger, no-PTM
IIb
Larger, C-terminal PTM
I
Modified by LanC- and
LanD-like enzymes
II
III
IV
Non-modified
bacteriocins (Class II)
Description
a
b
Modified by LanM-like
enzymes
Modified by LanKC-like
enzymes
Modified by LanL-like
enzymes
Pediocin-like
one-peptide
bacteriocins
Two-peptide
bacteriocins
c
Circular bacteriocins
d
Other one-peptide
bacteriocins
Microcin B17
Microcin J25
Microcin V
Microcin L
Microcin 24
Microcin E492
Microcin H47
Microcin M
Nisin
Subtilin
Mersacidin
Lacticin 481
Actagardine
Nukacin ISK-1
Lacticin 3147
(two-peptide)
Haloduracin
(two-peptide)
Labyrinthopeptin
—
Pediocin PA-1
Enterocin A
Sakacin P
Mesentericin Y105
Lactococcin G
Abp-118
Lactacin F
Enterocin A-48
Garvicin ML
Lactococcin A
Lactococcin B
Enterocin B
Lacticin Q
5.2
MccJ25
MccC
MccE492
Outer
membrane
FhuA
OmpF
Inner
membrane
SbmA
Yej
ABEF
Inhibition of
RNA polymerase
Inhibition of
Asp-tRNA synthetase
Figure 5.1 Schematic overview of mode of
action of three different microcins; MccJ25,
MccC, and MccE492. The enzyme parasitizes
different transport systems in the outerand inner membrane to gain access to their
Bacteriocins
Fep/
Cir/
Fiu
IID
IIC
IIC
IID
Man-PTS
Pore formation
target site. Transport of microcins via the
iron siderophore receptors (FhuA, Fep, Cir, or
Fiu, shown in dark) also require energy generated by the inner membrane located Tonsystem, which is not depicted in this figure.
(<5 kDa), which are often plasmid encoded and have extensive PTMs. Class
II consists of larger peptides without (Class IIa) or with (Class IIa) PTMs. It
should be noted that the recent classification scheme proposed by Arnison et al.
[8], separates the microcins in a number of different groups, according to their
different modifications.
Microcins can kill sensitive cells by using either intracellular or extracellular targets. In general, the mode of action of a microcin includes two steps
(Figure 5.1). The first step involves translocation of the peptide across the outer
membrane and sometimes the inner membrane to reach the target site. Such
translocation is often achieved by parasitizing different bacterial transport
systems [15]. This means that the peptides contain molecular structures that
mimic structures that are recognized and taken up by bacteria. For example, the
iron-uptake system is utilized by several microcins [16]. Bacteria secrete the socalled siderophore chelators (low-molecular-mass iron-binding compounds) to
scavenge environmental iron; when siderophores have bound iron extracellularly
they are recognized by siderophore receptors (e.g., FhuA, FepA, Cir, Fiu) in the
outer membrane and are transported into the periplasm. Several microcins (e.g.,
MccE492, MccH47) contain PTMs to mimic the siderophores, causing the bacteria to recognize the microcin as a siderophore and facilitate active translocation
across the outer membrane. Some microcins need to be further translocated
from the periplasm and across the inner membrane, and such transport may be
achieved by different transporters or uptake proteins (e.g., the ABC transporter
YejABEF, which transports MccC or the uptake protein SbmA) [15].
75
76
5 Antimicrobial Peptides from Prokaryotes
The second step in the mode of action of microcins is the cytotoxic action. As
mentioned above, there are, in principle, two different target sites: the inner membrane or the cytoplasm. In the inner membrane, microcins can act by forming
lethal pores, such as microcin E492, which binds to a sugar transporter complex known as the mannose phosphotransferase system to initiate pore formation.
Cytoplasmically active microcins act by inhibiting active enzymatic processes.
For example, the lasso peptide MccJ25 binds and blocks one of the channels of
RNA polymerase [17], while MccC is a non-hydrolysable mimic of the nucleoide
aspartyl adenylate that blocks protein synthesis by inhabiting aspartyl-tRNA synthetase [18].
5.2.2
Lanthibiotics – Post-translationally Modified Peptides from Gram-Positive Bacteria
The lantibiotics (also referred to as lantipeptides or Class I bacteriocins) are small
peptides of <40 aas that contain one or several PTMs. The best known member
of Class I is nisin, which has been studied for decades. The major PTMs of the
lantibiotics are thioether-based internal ring structures known as lanthionine or
𝛽-methyllanthionine. These are formed by dehydration of selected serine and/or
threonine residues followed by formation of a thiol bridge between the dehydrated
residue and the proximally located cysteine residue [19]. Lanthibiotics may also
contain other unusual amino acids formed by post-translational processes, including D-alanine [20].
Because of large structural variations, the subclassification of lantibiotics is not
straightforward. In the recent classification of RiPPs by Arnison et al. [8], lantibiotics are classified on the basis of the biosynthetic enzymes responsible for
the PTMs to create lanthionine or β-methyllanthionine motifs (Table 5.1). Here,
Class I are lantibiotics where dehydration and cyclization to form lanthionine or
β-methyllanthionine rings occur by the action of two separate enzymes known as
LanB (the dehydratase) and LanC (the cyclase). Well-known members of this class
are nisin, Pep5, and subtilin. Class II, III, and IV lantibiotics, on the other hand,
are modified by the action of a single enzyme (LanM for Class II, LanKC for Class
III, and LanL for Class IV) that carries out both the dehydration and cyclization
steps, but by different mechanisms [8]. Notable peptides such as nukacin ISK-1,
mersacidin, as well as the two-peptide lantibiotics lacticin 3147 and haloduracin
belong to Class II. In the latter examples, the antimicrobial activity requires the
combined activity of two individual peptides.
Lantibiotics generally have a broader inhibitory spectrum than other
prokaryotic AMPs and can kill a broad range of Gram-positive bacteria (but not
Gram-negative species). All lantibiotics kill sensitive bacteria by targeting their
cell envelope. The mode of action involves either pore formation or inhibition of
cell wall synthesis, or both these actions combined. For efficient targeting, the
lantibiotic recognizes specific molecules in the membrane of sensitive cells.
The most important lantibiotic target receptor is lipid II. Lipid II is a key
intermediate in peptidoglycan biosynthesis. It is synthesized intracellularly and
5.2
Bacteriocins
translocated across the phospholipid bilayer to supply peptidoglycan subunits
to the growing cell wall [21]. Binding to lipid II by antimicrobial compounds
appears to be an effective means to kill bacteria, as this molecule is targeted
by at least three other types of antibiotics in addition to lantibiotics, namely,
glycopeptides antibiotics (vancomycin), mannopeptimycin, and ramoplanin [21].
Nisin and epidermin were the first lantibiotics shown to use lipid II as a docking
molecule [22, 23], and the mechanism of action for nisin has been characterized
in great detail. Nisin binds lipid II via the characteristic lantibiotic ring structures
in the N-terminal part of the peptide, and pores containing both nisin and lipid
II are formed [24]. In addition, nisin also inhibits target cells by blocking cell
wall formation [25]. This mechanism is independent of the pore-forming activity
and involves relocation of lipid II into patches outside their functional location.
A number of different lantibiotics with N-terminal ring structures similar to
nisin kill target cells by lipid II-mediated pore formation [21]. These include
subtilin and members of the epidermin family as well as some two-component
lantibiotics, such as lacticin 3147 [21, 26]. Mersacidin, on the other hand, also
binds lipid II, but the mechanism of killing involves only inhibition of cell wall
synthesis and pores are not formed in this case [23].
5.2.3
Non-modified Peptides from Gram-Positive Bacteria
In addition to the lantibiotics described above, Gram-positive bacteria also
produce a wide variety of peptide bacteriocins that do not contain PTMs;
these are often referred to as Class II bacteriocins [27]. These peptides can be
classified into subgroups on the basis of their sequence and physiochemical
properties (Table 5.1). Most of these peptides are cationic and amphiphilic and
they kill target cells by creating pores in the membrane, disrupting the membrane
integrity, or permeabilizing the membrane of the target cells in other ways. Such
mechanisms cause leakage of low molecular weight compounds (e.g., ions, K+ ,
PO4 2− , H+ ), leading to dissipation of a proton motive force that is deleterious for
the cell [11]. Similarly to lantibiotics, Class II bacteriocins seem to target specific
molecules in the membrane of sensitive cells to lethally damage the membrane;
however, such receptors/targets have only been identified for some of the Class
II bacteriocins.
The group of pediocin-like bacteriocins (Class IIa) are named after the first
characterized member of this group, pediocin PA-1 from Pediococcus pentosaceus
[28]. These peptides are 36–79 aa long and are grouped in this subclass based on
the presence of a highly conserved N-terminal consensus sequence (YGNGVxCxxxxCxVxWxxA, where x is any amino acid). Structural analysis has shown that
this conserved N-terminal consensus sequence is part of a β-sheet structure,
while the less conserved N-terminal part of the peptides form a hairpin structure
consisting of one of two alpha-helices (helix-turn-helix structure). Pediocin-like
bacteriocins are particularly efficient in inhibiting species within the genera of
Listeria and Enterococcus, and are therefore of medical interest. To kill sensitive
77
78
5 Antimicrobial Peptides from Prokaryotes
Nisin
Outside
Inside
Lipid II
Pore formation
(a)
Class lla bacteriocin
Outside
Outside
IID
IIC
IIC
IID
IID
IIC
IIC
IID
Inside
Inside
Man-PTS
Pore formation
(b)
Figure 5.2 Schematic overview of mode
of action of two different types of poreforming bacteriocins from Gram-positive
bacteria. (a) a pore by insertion of the peptide into the lipid bilayer. Nisin also has a
second mode of action not depicted in the
figure, which involves blocking of cell wall
by binding to lipid II. (b) The class IIa bacteriocins (pediocin-like bacteriocins) bind to the
sugar transporter known as the mannosephosphotransferase system to form pores.
cells, pediocin-like bacteriocins target the membrane-located proteins of a
sugar transporter complex known as the mannose phosphotransferase system on
sensitive cells [29–31] (Figure 5.2b). More specifically, this glucose/mannosespecific translocation channel, consisting of proteins IIC and IID together form
the receptor, and a 40 aa extracellular loop of the IIC protein is involved in the
specific recognition [31, 32]. In order to avoid self-killing, these bacteriocins
have an intracellular immunity protein that blocks the bacteriocin-induced pore
formation by binding tightly to its own mannose phosphotransferase system
[33]. Also worth noting is that resistance toward Class IIa bacteriocins seems
to develop quite easily, and the main mechanism is probably downregulation of
receptor expression [34].
Class IIb consists of bacteriocins, whose activity depends on the complementary action of two individual peptides [35]. The two genes encoding the peptides of
length between 25 and 38 aas are located next to each other on the same operon.
Their structures are characterized by a well-defined, central α-helix region with
flexible regions in both ends. Although no structure has been determined for
the peptide heterodimers, it has been shown that there is direct physical interaction between the complementary peptides when they exert antimicrobial activity
5.3
Applications of Prokaryotic AMPs
[35, 36]. The inhibitory spectra of these bacteriocins are narrow; for example,
lactococcin G is active only against strains of Lactococcus, while plantaricins EF
and JK, which are coproduced by the same host, display activity only against a few
strains of Lactobacillus and Pediococcus.
One group of antimicrobial peptides from Gram-positive bacteria needs to
undergo a cyclization process in order to be active; this is Class IIc, the cyclic
or circular bacteriocins. In these peptides, the N- and C-terminal ends are
covalently linked by an amide bond [37, 38], but the molecular details underlying
this process are still unclear. The circular bacteriocins are both remarkably stable
and display very potent antimicrobial activity [39]. Circular bacteriocins kill
target cells by disrupting the membrane integrity [38]; however, it is not clear
whether circular bacteriocins need a target receptor. While enterocin AS-48 is
able to form pores in lipid bilayers [40], it has been shown for garvicin ML that
the maltose ABC transporter is important for the sensitivity, and may function as
a docking molecule [41].
Some unmodified, linear, non-pediocin-like bacteriocins do not fit into any of
the classes mentioned above and are therefore placed in Class IId. One member of
this group is lactococcin A, a 54 aa long peptide produced by Lactococcus lactis.
Interestingly, this peptide has a similar mode of action as the pediocin-like bacteriocins and microcin MccE492; it uses the mannose phosphotransferase system as
receptor on target cells [33]. This similarity in the mode of action is found despite
a lack of the sequence similarity between these groups of peptides, indicating that
the mannose phosphotransferase system is a weak spot for bacteria when it comes
to antimicrobial attack.
The leaderless bacteriocins constitute a significant group of antibacterial peptides that has been given little attention. They have been isolated from different
Gram-positive bacteria such as Staphylococcus aureus, Enterococcus faecium,
L. lactis, as well as other bacteria.
It has been shown that L. lactis subsp. lactis BGMN1-5 produces a leaderless
Class II bacteriocin called LsbB [42]. Recently, it was published that a membranebound peptidase gene was involved in bacteriocin sensitivity in target cells. It was
concluded that this membrane-bound peptidase acts as a specific receptor for the
LsbB bacteriocins [43].
5.3
Applications of Prokaryotic AMPs
5.3.1
Food Biopreservation
Biopreservation is explained as improved safety of foods as well as extended
storage time using natural microflora in the form of protective cultures and/or
their metabolites as antimicrobial agents [44]. The bacteriocins from lactic acid
bacteria (LAB) are well accepted as natural means of food biopreservation [45],
79
80
5 Antimicrobial Peptides from Prokaryotes
and the well-known lantibiotic nisin has been used for this purpose for decades.
Three different strategies have been explored for application of bacteriocins in
food products: addition in purified form, in situ production by bacteriocinogenic
bacteria, and binding of bacteriocins to polymeric packaging [45, 46].
Bacteriocins such as purified or semi-purified nisin and enterocin AS-48 can
be added to food as biopreservatives to prolong shelf-life and reduce the usage
of conventional preservatives with negative effects for human health such as
nitrates. They have been found to inhibit many Gram-positive and also some
Gram-negative bacterial species. This broad inhibitory spectrum has encouraged
the researchers to evaluate their antimicrobial efficacy in foods. For example,
enterocin AS-48 in ready-to-eat fruit and vegetable products always retained
above 65% of the activity [47]; nisin can be incorporated into kimchi and cooked
mashed potatoes to control and preserve the product by inhibiting lactobacilli
(which are responsible for the over-ripening of kimchi [48]) as well as Bacillus
and Clostridium (which have caused several food-borne outbreaks [46]).
Interestingly, another practical approach is to introduce the bacteriocinogenic
strains (particularly LAB) as a starter culture directly into fermented sausage or
dairy product without purification to increase food quality and safety [49, 50].
Enterocin AS-48 produced by bacteriocinogenic strains can interact directly to
target spoilage and poisonous bacteria such as Listeria monocytogenes in model
sausages [51] and S. aureus in milk [50].
Development of packaging films with antimicrobial entity is a suitable alternative to preserve a complex food environment [52, 53]. Bacteriocins are the most
interesting antimicrobial agents in this regard because of their natural and safe
properties [54]. The efficacy of films coated by nisin [55], enterocin A and B [56],
and enterocin 416K1 [57] in the strong inhibition of L. monocytogenes in cooked
meat products has been proved. Such a method provides longer preservation in
comparison with addition of purified enterocins into foods [58].
5.3.2
Bacteriocinogenic Probiotics
Recent studies have shown that the impact of bacteriocin-producing bacteria
as probiotics on gut health (where they will possibly secrete bacteriocins) are
significantly more effective than purified bacteriocin [59–61]. Probiotic strains
may work better because of their prophylactic benefits (interact directly with the
sensitive organisms in the intestine); however, purified bacteriocins possess the
potential to target specific pathogens in established infection. Furthermore, the
peptide-based structure of AMPs that are susceptible to proteolytic digestion
in the stomach provides some limitation regarding their application in the food
and pharmaceutical industry [49, 62, 63]. Most compelling, Corr et al. [64]
showed that L. monocytogenes infection in the mice gut could be inhibited by a
bacteriocin-producing probiotic strain of Lactobacillus salivarius, and they also
demonstrated that this protection depended strictly on the bacteriocin and not
other probiotic effects. Furthermore, in a study by Bhardwaj et al. [65] it was
5.3
Applications of Prokaryotic AMPs
demonstrated that consuming bacteriocin-producing E. faecium KH24 in mice
significantly modulated fecal microbiota in GI tract of mice. The authors emphasized that Lactobacillus population in mice fed with bacteriocin-producing
E. faecium is much greater than those mice fed with non-producing strains.
These studies highlight the potential of such probiotics to serve to control the
indigenous microbiota in a beneficial manner.
5.3.3
Clinical Application
Nisin A is one of the first AMPs that were studied in animal infection model as
early as 1952 [66]. In spite of the rapid clearance of nisin A from blood, it has
the same efficacy as penicillin in the treatment of infections related to Mycobacterium tuberculosis, Streptococcus pyogenes, and S. aureus in mice. Interestingly,
owing to the rapid bactericidal action of nisin in comparison to vancomycin, two
intravenous doses of 0.16 mg kg−1 nisin are even more effective than two intravenous doses of 1.25 mg kg−1 vancomycin in treating mice infected with Streptococcus pneumonia [66–68]. In another study by Okuda et al. [69] the anti-biofilm
effect of nisin A, lacticin Q, and nukacin ISK-1 were evaluated against a clinical
isolate of methicillin-resistant S. aureus (MRSA). Interestingly, among the abovementioned bacteriocins, nisin A demonstrated the highest bactericidal activity
against planktonic as well as biofilm cells. Mechanistically, the authors suggested
that nisin A can form stable pores on biofilm cells, and therefore it can effectively
be used to prevent biofilm-associated infections, especially in patients using central venous catheters [69]. Nisin has also been tested for functions beyond the
direct antimicrobial effect.
For example, prevention of cell proliferation in vitro and in vivo in head and neck
squamous cell carcinoma (HNSCC) was recently assessed [70]. Apoptosis, also
known as programmed cell death, is a natural process that eliminates older cells
in multicellular organisms; however, cancer cells are resistant to apoptosis [71].
Joo et al. demonstrated that nisin decreases HNSCC tumorigenesis by inducing
preferential apoptosis and inhibiting proliferation in HNSCC cells compared to
primary keratinocytes. Mechanistically, nisin acts by increasing the intracellular
calcium, induction of cell cycle arrest, and activation of CHAC1 (a proapoptotic
cation transport regulator) expression. Joo et al. [70] proposed application of nisin
as a novel antitumor agent for HNSCC in the near future.
The discovery of the sactibiotic thuricin CD by Rea et al. [72], which is produced by Bacillus thuringiensis, showed that Clostridium difficile-associated diarrhea (CDAD) can be treated without prescribing broad-spectrum antibiotics in a
model of the human distal colon. Antibiotics such as vancomycin and metronidazole, which are used clinically to treat CDAD, dramatically modulate the resident
gut microbiota, whereas thuricin CD, which has comparable antimicrobial activity, did not change the composition of the commensal microbiota much. Thus,
thuricin CD can fight CDAD without causing the unwanted side effects often
faced with conventional antibiotics.
81
82
5 Antimicrobial Peptides from Prokaryotes
5.3.4
Applications in Dental Care
Bacteriocins can play a significant role in controlling the ecological balance in
human oral microbial community. S. pyogenes (the important etiological agent
in human dental caries) and Streptococcus salivarius (predominant component
of oral commensal throughout life) both colonize on the oral mucosal surface.
However, bacteriocins of S. salivarius can block the growth of a virulent S. pyogenes biofilm by signal transduction interaction. Therefore, oral bacteria and their
bacteriocins allow bacteria to select their “neighbors” and establish a stable oral
microbiota [73–75].
In recent years, antimicrobial peptides such as nisin, mutacin, and ϵ-poly-Llysine have been widely studied and seem to have practical potential to target
the harmful cariogenic bacteria such as Streptococcus mutans and Streptococcus
sobrinus [76, 77]. Nisin, which is highly active and stable at low pH (2–6) [54, 78],
may suffer from drawbacks regarding instability, low solubility, and consequently
lower efficacy in the neutral and weakly acidic condition in human oral cavity [76].
In this respect, according to the published research of Rollema et al. [78] solubility
and stability of nisin with acceptable activity at neutral pH has been improved by
the application of protein engineering approaches.
ϵ-Poly-L-lysine is another relatively broad spectrum, wide pH range stable and
water soluble generally recognized as safe (GRAS) antimicrobial peptide [79, 80].
In addition, it is the secondary metabolite of Streptomyces albulus, which has
been approved as a natural preservative since the late 1980s and in January 2004
in Japan and the United States, respectively [79, 81]. Interestingly in a study,
Najjar et al. [82] addressed the synergistic activity of nisin and ϵ-poly-L-lysine on
S. mutans. An implication of the above-mentioned finding is the high possibility of
anti-caries combination therapy of nisin and poly-L-lysine.
Despite extensive in vitro research regarding the effectiveness as well as safety
application of nisin in chewing gum and oral formulations against cariogenic
pathogens [83–85] and oral diseases such as gingivitis in animal model [86], very
little progress has been made to develop commercial applications.
5.4
Development and Discovery of Novel AMP
The group of antimicrobials referred to as prokaryotic AMPs offer a wide variety of different peptides that have promising applications in different settings and
against many important pathogens. Since most of these compounds have not yet
undergone clinical tests, knowledge about resistance development only comes
from laboratory-based research [6]. This has shown that sensitive cells can develop
resistance to some Class II bacteriocins (i.e., Class IIa bacteriocins and lactococcin A) by downregulating the expression of the target receptor [34]. Furthermore,
resistance against microcins may develop through mutations in genes encoding
5.4
Development and Discovery of Novel AMP
the bacteriocins target; for example, resistance to MccJ25 can occur by mutation
in RNA polymerase [87]. Lantibiotics seem to have less of these problems, since
they do not target proteins; however, transient lantibiotic resistance could emerge
from changes in the cell envelope composition [88, 89]. Nevertheless, it is worth
noting that nisin has been used in different food applications for decades already
without any major concern about resistance development.
Resistance development among different prokaryotic AMPs still needs to be
examined under in vivo settings; however, one great advantage with this group
of antimicrobials is their gene-encoded nature. This creates huge possibilities for
in silico discovery as well as bioengineering approaches to obtain novel peptide
variants that can overcome potential problems with resistance development. As
mentioned before, biosynthesis of bacteriocins relies on three or four different
types of genes: (i) structural gene(s), (ii) immunity gene(s), (iii) transport genes,
and in some cases, and (iv) genes responsible for PTMs. Since these genes often
are co-located on the genome and also have some well-known characteristics,
softwares such as BAGEL (http://bagel.molgenrug.nl) can be made to automatically mine new bacterial genome sequences for candidate bacteriocin genes.
Furthermore, bioengineering approaches have huge potential to generate new
and even synthetic variants of prokaryotic AMPs [90, 91]. For example, by sitedirected mutagenesis in the bacteriocin structural genes, new peptide variants
can be constructed. As further advancement, by using different synthetic biology
approaches and heterologous expression systems, nonexpressed bacteriocin
genes can be produced and novel combinations of modifications can be introduced in the same peptide [90, 92]. Lastly, chemical synthesis of peptides, based
on knowledge of natural peptides and their mode of action can further increase
the repertoire of prokaryotic AMPs with improved potency, stability, or inhibition
spectrum.
In this context, it is also interesting to note that peptide engineering has been
used to change the activity of other, non-peptide, antimicrobial molecules. For
example, pheromonicin is an engineered peptide that consists of the S. aureus
pheromone gene (AgrD1), which is fused to the colicin Ia [93]. While the poreforming colicin Ia (a pore-forming protein) is only active against Escherichia coli,
fusion of the AgrD1 sequence, directs its activity toward S. aureus (which is not its
natural target). Pheromonicin was found active against both methicillin-sensitive
S. aureus (MSSA) and MRSA. Strikingly, growth was inhibited by 60% in resistant
MRSA (resistant cells). Furthermore, the efficacy of pheromonicin was also shown
in mice infected by MRSA [93]. In another study, colicin Ia was fused to another
pheromone gene, namely, cCF10 of E. faecalis to construct pheromonicin PMCEF. Using this strategy, colicin activity was directed against vancomycin-resistant
Enterococcus (VRE) both in vitro and in mice-made bacteremia by VRE [94]. Furthermore, Janiszewska and Urbanczyk-Lipkowska [95] used peptide engineering
to create a new class of low molecular weight dendrimeric antimicrobial peptides
[95], where the antimicrobial activity of linear peptides was transferred to the
so-called dendrimers. Dendrimers are new synthetic branched macromolecules,
which are characterized by tree-like structures with a high number of functional
83
84
5 Antimicrobial Peptides from Prokaryotes
groups on the molecular surface [95, 96]. Their well-defined scaffold makes them
applicable for various biomedical purposes, for example, protein mimetics [97],
diagnostic purposes [98], and as a carrier in drug and gene delivery [99, 100]. They
obtained peptide dendrimers with amphiphilic properties and positive charges
(two key structural elements for antimicrobial activity) [101–103] and expressed
high antimicrobial effect against E. coli, S. aureus, as well as Candida albicans
because of their interaction with the cell membrane [95].
References
1. Overbye, K.M. and Barrett, J.F. (2005)
2.
3.
4.
5.
6.
7.
Antibiotics: where did we go wrong?
Drug Discovery Today, 10, 45–52.
Hassan, M., Kjos, M., Nes, I.F., Diep,
D.B., and Lotfipour, F. (2012) Natural
antimicrobial peptides from bacteria:
characteristics and potential applications to fight against antibiotic
resistance. J. Appl. Microbiol., 113,
723–736.
Finch, R. and Hunter, P.A. (2006)
Antibiotic resistance–action to promote new technologies: report of an
EU Intergovernmental conference held
in Birmingham, UK, 12-13 December
2005. J. Antimicrob. Chemother., 58
(Suppl. 1), i3–i22.
Belguesmia, Y., Madi, A., Sperandio,
D., Merieau, A., Feuilloley, M., Prevost,
H., Drider, D., and Connil, N. (2011)
Growing insights into the safety of
bacteriocins: the case of enterocin S37.
Res. Microbiol., 162, 159–163.
Hassan, M., Javadzadeh, Y., Lotfipour,
F., and Badomchi, R. (2011) Determination of comparative minimum
inhibitory concentration (MIC) of bacteriocins produced by enterococci for
selected isolates of multi-antibiotic
resistant Enterococcus spp. Adv. Pharm.
Bull., 1, 75–79.
Cotter, P.D., Ross, R.P., and Hill, C.
(2013) Bacteriocins - a viable alternative to antibiotics? Nat. Rev. Microbiol.,
11, 95–105.
Nes, I.F., Diep, D.B., and Holo, H.
(2007) Bacteriocin diversity in Streptococcus and Enterococcus. J. Bacteriol.,
189, 1189–1198.
8. 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.,
Muller, 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.,
Sussmuth, 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 posttranslationally modified peptide natural
products: overview and recommendations for a universal nomenclature. Nat.
Prod. Rep., 30, 108–160.
9. Cintas, L.M., Casaus, P., Holo, H.,
Hernandez, P.E., Nes, I.F., and
Havarstein, L.S. (1998) Enterocins
L50A and L50B, two novel bacteriocins
from Enterococcus faecium L50, are
related to staphylococcal hemolysins.
J. Bacteriol., 180, 1988–1994.
10. Iwatani, S., Yoneyama, F., Miyashita,
S., Zendo, T., Nakayama, J., and
Sonomoto, K. (2012) Identification
of the genes involved in the secretion
and self-immunity of lacticin Q, an
unmodified leaderless bacteriocin from
References
11.
12.
13.
14.
15.
16.
17.
18.
Lactococcus lactis QU 5. Microbiology,
158, 2927–2935.
Eijsink, V.G., Axelsson, L., Diep, D.B.,
Havarstein, L.S., Holo, H., and Nes,
I.F. (2002) Production of class II bacteriocins by lactic acid bacteria; an
example of biological warfare and communication. Antonie Van Leeuwenhoek,
81, 639–654.
Rosengren, K.J. and Craik, D.J. (2009)
How bugs make lassos. Chem. Biol., 16,
1211–1212.
Rosengren, K.J., Blond, A., Afonso,
C., Tabet, J.C., Rebuffat, S., and Craik,
D.J. (2004) Structure of thermolysin
cleaved microcin J25: extreme stability
of a two-chain antimicrobial peptide
devoid of covalent links. Biochemistry,
43, 4696–4702.
Rebuffat, S. (2012) Microcins in action:
amazing defence strategies of Enterobacteria. Biochem. Soc. Trans., 40,
1456–1462.
Duquesne, S., Destoumieux-Garzon,
D., Peduzzi, J., and Rebuffat, S. (2007)
Microcins, gene-encoded antibacterial
peptides from enterobacteria. Nat.
Prod. Rep., 24, 708–734.
Destoumieux-Garzon, D., Duquesne,
S., Peduzzi, J., Goulard, C., Desmadril,
M., Letellier, L., Rebuffat, S., and
Boulanger, P. (2005) The ironsiderophore transporter FhuA is the
receptor for the antimicrobial peptide microcin J25: role of the microcin
Val11-Pro16 beta-hairpin region in the
recognition mechanism. Biochem. J.,
389, 869–876.
Adelman, K., Yuzenkova, J., La Porta,
A., Zenkin, N., Lee, J., Lis, J.T.,
Borukhov, S., Wang, M.D., and
Severinov, K. (2004) Molecular mechanism of transcription inhibition by
peptide antibiotic Microcin J25. Mol.
Cell, 14, 753–762.
Metlitskaya, A., Kazakov, T., Kommer,
A., Pavlova, O., Praetorius-Ibba, M.,
Ibba, M., Krasheninnikov, I., Kolb, V.,
Khmel, I., and Severinov, K. (2006)
Aspartyl-tRNA synthetase is the
target of peptide nucleotide antibiotic Microcin C. J. Biol. Chem., 281,
18033–18042.
19. Bierbaum, G. and Sahl, H.G. (2009)
20.
21.
22.
23.
24.
25.
26.
27.
Lantibiotics: mode of action, biosynthesis and bioengineering. Curr. Pharm.
Biotechnol., 10, 2–18.
Skaugen, M., Nissen-Meyer, J., Jung,
G., Stevanovic, S., Sletten, K., Inger, C.,
Abildgaard, M., and Nes, I.F. (1994)
In vivo conversion of L-serine to
D-alanine in a ribosomally synthesized polypeptide. J. Biol. Chem., 269,
27183–27185.
Breukink, E. and de Kruijff, B. (2006)
Lipid II as a target for antibiotics. Nat.
Rev. Drug Discovery, 5, 321–332.
Breukink, E., Wiedemann, I., van
Kraaij, C., Kuipers, O.P., Sahl, H.G.,
and de Kruijff, B. (1999) Use of the
cell wall precursor lipid II by a poreforming peptide antibiotic. Science,
286, 2361–2364.
Brotz, 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.
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.
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.
Wiedemann, I., Bottiger, T., Bonelli,
R.R., Wiese, A., Hagge, S.O.,
Gutsmann, T., Seydel, U., Deegan, L.,
Hill, C., Ross, P., and Sahl, H.G. (2006)
The mode of action of the lantibiotic
lacticin 3147 - A complex mechanism
involving specific interaction of two
peptides and the cell wall precursor
lipid II. Mol. Microbiol., 61, 285–296.
Kjos, M., Borrero, J., Opsata, M., Birri,
D.J., Holo, H., Cintas, L.M., Snipen, L.,
Hernandez, P.E., Nes, I.F., and Diep,
D.B. (2011) Target recognition, resistance, immunity and genome mining
85
86
5 Antimicrobial Peptides from Prokaryotes
28.
29.
30.
31.
32.
33.
34.
35.
of class II bacteriocins from Grampositive bacteria. Microbiology, 157,
3256–3267.
Nieto Lozano, J.C., Meyer, J.N., Sletten,
K., Pelaz, C., and Nes, I.F. (1992) Purification and amino acid sequence of a
bacteriocin produced by Pediococcus
acidilactici. J. Gen. Microbiol., 138,
1985–1990.
Ramnath, M., Beukes, M., Tamura,
K., and Hastings, J.W. (2000) Absence
of a putative mannose-specific phosphotransferase system enzyme IIAB
component in a leucocin A-resistant
strain of Listeria monocytogenes, as
shown by two-dimensional sodium
dodecyl sulfate-polyacrylamide gel electrophoresis. Appl. Environ. Microbiol.,
66, 3098–3101.
Hechard, Y. and Sahl, H.G. (2002)
Mode of action of modified and
unmodified bacteriocins from Grampositive bacteria. Biochimie, 84,
545–557.
Kjos, M., Nes, I.F., and Diep, D.B.
(2009) Class II one-peptide bacteriocins
target a phylogenetically defined subgroup of mannose phosphotransferase
systems on sensitive cells. Microbiology,
155, 2949–2961.
Kjos, M., Snipen, L., Salehian, Z., Nes,
I.F., and Diep, D.B. (2010) The Abi
proteins and their involvement in bacteriocin self-immunity. J. Bacteriol.,
192, 2068–2076.
Diep, D.B., Skaugen, M., Salehian, Z.,
Holo, H., and Nes, I.F. (2007) Common
mechanisms of target cell recognition
and immunity for class II bacteriocins.
Proc. Natl. Acad. Sci. U.S.A., 104,
2384–2389.
Kjos, M., Nes, I.F., and Diep, D.B.
(2011) Mechanisms of resistance to
bacteriocins targeting the mannose
phosphotransferase system. Appl.
Environ. Microbiol., 77, 3335–3342.
Oppegard, C., Rogne, P., Emanuelsen,
L., Kristiansen, P.E., Fimland, G., and
Nissen-Meyer, J. (2007) The twopeptide class II bacteriocins: structure,
production, and mode of action. J. Mol.
Microbiol. Biotechnol., 13, 210–219.
36. Rogne, P., Fimland, G., Nissen-Meyer,
37.
38.
39.
40.
41.
42.
43.
44.
J., and Kristiansen, P.E. (2008) Threedimensional structure of the two
peptides that constitute the two-peptide
bacteriocin lactococcin G. Biochim.
Biophys. Acta, 1784, 543–554.
van Belkum, M.J., Martin-Visscher,
L.A., and Vederas, J.C. (2011) Structure
and genetics of circular bacteriocins.
Trends Microbiol., 19, 411–418.
Montalban-Lopez, M.,
Sanchez-Hidalgo, M., Cebrian, R.,
and Maqueda, M. (2012) Discovering
the bacterial circular proteins: bacteriocins, cyanobactins, and pilins. J. Biol.
Chem., 287, 27007–27013.
Sanchez-Hidalgo, M.,
Montalban-Lopez, M., Cebrian, R.,
Valdivia, E., Martinez-Bueno, M., and
Maqueda, M. (2012) AS-48 bacteriocin:
close to perfection. Cell. Mol. Life Sci.,
68, 2845–2857.
Maqueda, M., Gálvez, A., Bueno, M.M.,
Sanchez-Barrena, M.J., González, C.,
Albert, A., Rico, M., and Valdivia, E.
(2004) Peptide AS-48: prototype of a
new class of cyclic bacteriocins. Curr.
Protein Pept. Sci., 5, 399–416.
Gabrielsen, C., Brede, D.A., Hernandez,
P.E., Nes, I.F., and Diep, D.B. (2012)
The maltose ABC transporter in Lactococcus lactis facilitates high-level
sensitivity to the circular bacteriocin garvicin ML. Antimicrob. Agents
Chemother., 56, 2908–2915.
Gajic, O., Buist, G., Kojic, M.,
Topisirovic, L., Kuipers, O.P., and
Kok, J. (2003) Novel mechanism of
bacteriocin secretion and immunity
carried out by lactococcal multidrug
resistance proteins. J. Biol. Chem., 278,
34291–34298.
Uzelac, G., Kojic, M., Lozo, J.,
Aleksandrzak-Piekarczyk, T.,
Gabrielsen, C., Kristensen, T., Nes, I.F.,
Diep, D.B., and Topisirovic, L. (2013)
A Zn-dependent metallopeptidase is
responsible for sensitivity to LsbB, a
class II leaderless bacteriocin of Lactococcus lactis subsp. Lactis BGMN1-5.
J. Bacteriol., 195, 5614–5621.
Stiles, M.E. (1996) Biopreservation
by lactic acid bacteria. Antonie Van
Leeuwenhoek, 70, 331–345.
References
45. Galvez, A., Lopez, R.L., Abriouel, H.,
46.
47.
48.
49.
50.
51.
52.
53.
54.
Valdivia, E., and Omar, N.B. (2008)
Application of bacteriocins in the
control of foodborne pathogenic and
spoilage bacteria. Crit. Rev. Biotechnol.,
28, 125–152.
Settanni, L. and Corsetti, A. (2008)
Application of bacteriocins in vegetable food biopreservation. Int. J. Food
Microbiol., 121, 123–138.
Grande, M.J., Lucas, R., Valdivia, E.,
Abriouel, H., Maqueda, M., Omar, N.B.,
Martinez-Canamero, M., and Galvezi,
A. (2005) Stability of enterocin AS-48
in fruit and vegetable juices. J. Food
Prot., 68, 2085–2094.
Choi, M.H. and Park, Y.H. (2000) Selective control of lactobacilli in kimchi
with nisin. Lett. Appl. Microbiol., 30,
173–177.
De Vuyst, L. and Leroy, F. (2007)
Bacteriocins from lactic acid bacteria: production, purification, and
food applications. J. Mol. Microbiol.
Biotechnol., 13, 194–199.
Mañoz, A., Ananou, S., Gálvez, A.,
Martínez-Bueno, M., Rodríguez, A.,
Maqueda, M., and Valdivia, E. (2007)
Inhibition of Staphylococcus aureus in
dairy products by enterocin AS-48 produced in situ and ex situ: bactericidal
synergism with heat. Int. Dairy J., 17,
760–769.
Khan, H., Flint, S., and Yu, P.L. (2010)
Enterocins in food preservation. Int. J.
Food Microbiol., 141, 1–10.
Mauriello, G., Ercolini, D., La Storia,
A., Casaburi, A., and Villani, F. (2004)
Development of polythene films for
food packaging activated with an
antilisterial bacteriocin from Lactobacillus curvatus 32Y. J. Appl.
Microbiol., 97, 314–322.
Suppakul, P., Miltz, J., Sonneveld, K.,
and Bigger, S.W. (2003) Active packaging technologies with an emphasis
on antimicrobial packaging and its
applications. J. Food Sci., 68, 408–420.
Cleveland, J., Montville, T.J., Nes, I.F.,
and Chikindas, M.L. (2001) Bacteriocins: safe, natural antimicrobials
for food preservation. Int. J. Food
Microbiol., 71, 1–20.
55. Neetoo, H., Ye, M., Chen, H., Joerger,
56.
57.
58.
59.
60.
61.
62.
63.
R.D., Hicks, D.T., and Hoover, D.G.
(2008) Use of nisin-coated plastic films
to control Listeria monocytogenes on
vacuum-packaged cold-smoked salmon.
Int. J. Food Microbiol., 122, 8–15.
Marcos, B., Aymerich, T., Monfort,
J.M., and Garriga, M. (2007) Use of
antimicrobial biodegradable packaging
to control Listeria monocytogenes during storage of cooked ham. Int. J. Food
Microbiol., 120, 152–158.
Iseppi, R., Pilati, F., Marini, M., Toselli,
M., de Niederhausern, S., Guerrieri,
E., Messi, P., Sabia, C., Manicardi, G.,
Anacarso, I., and Bondi, M. (2008)
Anti-listerial activity of a polymeric
film coated with hybrid coatings
doped with Enterocin 416 K1 for use as
bioactive food packaging. Int. J. Food
Microbiol., 123, 281–287.
Quintavalla, S. and Vicini, L. (2002)
Antimicrobial food packaging in meat
industry. Meat Sci., 62, 373–380.
Cruchet, S., Obregon, M.C., Salazar,
G., Diaz, E., and Gotteland, M. (2003)
Effect of the ingestion of a dietary
product containing Lactobacillus
johnsonii La1 on Helicobacter pylori
colonization in children. Nutrition, 19,
716–721.
Lesbros-Pantoflickova, D.,
Corthesy-Theulaz, I., and Blum, A.L.
(2007) Helicobacter pylori and probiotics. J. Nutr., 137, 812S–818S.
Coconnier, M.H., Lievin, V., Hemery,
E., and Servin, A.L. (1998) Antagonistic
activity against Helicobacter infection
in vitro and in vivo by the human Lactobacillus acidophilus strain LB. Appl.
Environ. Microbiol., 64, 4573–4580.
Gillor, O., Etzion, A., and Riley, M.A.
(2008) The dual role of bacteriocins as
anti- and probiotics. Appl. Microbiol.
Biotechnol., 81, 591–606.
Hassan, M., Diep, D.B., Javadzadeh, Y.,
Dastmalchi, S., Nes, I.F., Sharifi, Y., Yari,
S., Farajnia, S., and Lotfipour, F. (2012)
Prevalence of bacteriocin activities and
bacteriocin-encoding genes in enterococcal clinical isolates in Iran. Can. J.
Microbiol., 58, 359–368.
87
88
5 Antimicrobial Peptides from Prokaryotes
64. Corr, S.C., Li, Y., Riedel, C.U., O’Toole,
65.
66.
67.
68.
69.
70.
71.
72.
73.
P.W., Hill, C., and Gahan, C.G.M.
(2007) Bacteriocin production as a
mechanism for the antiinfective activity
of Lactobacillus salivarius UCC118.
Proc. Natl. Acad. Sci. U.S.A., 104,
7617–7621.
Bhardwaj, A., Gupta, H., Kapila, S.,
Kaur, G., Vij, S., and Malik, R.K. (2010)
Safety assessment and evaluation of
probiotic potential of bacteriocinogenic
Enterococcus faecium KH 24 strain
under in vitro and in vivo conditions.
Int. J. Food Microbiol., 141, 156–164.
Bavin, E.M., Beach, A.S., Falconer,
R., and Friedmann, R. (1952) Nisin in
experimental tuberculosis. Lancet, 1,
127–129.
Goldstein, B.P., Wei, J., Greenberg, K.,
and Novick, R. (1998) Activity of nisin
against Streptococcus pneumoniae, in
vitro, and in a mouse infection model.
J. Antimicrob. Chemother., 42, 277–278.
Mota-Meira, M., Morency, H., and
Lavoie, M.C. (2005) In vivo activity
of mutacin B-Ny266. J. Antimicrob.
Chemother., 56, 869–871.
Okuda, K.I., Zendo, T., Sugimoto,
S., Iwase, T., Tajima, A., Yamada, S.,
Sonomoto, K., and Mizunoe, Y. (2013)
Effects of bacteriocins on methicillinresistant Staphylococcus aureus biofilm.
Antimicrob. Agents Chemother., 57,
5572–5579.
Joo, N.E., Ritchie, K., Kamarajan,
P., Miao, D., and Kapila, Y.L. (2012)
Nisin, an apoptogenic bacteriocin and
food preservative, attenuates HNSCC
tumorigenesis via CHAC1. Cancer
Med., 1, 295–305.
Mattson, M.P. and Chan, S.L. (2003)
Calcium orchestrates apoptosis. Nat.
Cell Biol., 5, 1041–1043.
Rea, M.C., Sit, C.S., Clayton, E.,
O’Connor, P.M., Whittal, R.M., Zheng,
J., Vederas, J.C., Ross, R.P., and Hill,
C. (2010) Thuricin CD, a posttranslationally modified bacteriocin with
a narrow spectrum of activity against
Clostridium difficile. Proc. Natl. Acad.
Sci. U.S.A., 107, 9352–9357.
Wescombe, P.A., Hale, J.D.F., Heng,
N.C.K., and Tagg, J.R. (2012) Developing oral probiotics from Streptococcus
74.
75.
76.
77.
78.
79.
80.
81.
82.
83.
salivarius. Future Microbiol., 7,
1355–1371.
Favier, C.F., Vaughan, E.E., De Vos,
W.M., and Akkermans, A.D. (2002)
Molecular monitoring of succession
of bacterial communities in human
neonates. Appl. Environ. Microbiol., 68,
219–226.
Wescombe, P.A., Heng, N.C.K., Burton,
J.R., and Tagg, J.R. (2010) Something
old and something new: an update on
the amazing repertoire of bacteriocins
produced by Streptococcus salivarius. Probiotics Antimicrob. Proteins, 2,
37–45.
Pepperney, A. and Chikindas, M.L.
(2011) Antibacterial peptides: opportunities for the prevention and treatment
of dental caries. Probiotics Antimicrob.
Proteins, 3, 68–96.
Chikindas, M.L., Novak, J., Driessen,
A.J., Konings, W.N., Schilling, K.M.,
and Caufield, P.W. (1995) Mutacin II,
a bactericidal antibiotic from Streptococcus mutans. Antimicrob. Agents
Chemother., 39, 2656–2660.
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.
Yoshida, T. and Nagasawa, T. (2003)
epsilon-Poly-L-lysine: microbial production, biodegradation and application
potential. Appl. Microbiol. Biotechnol.,
62, 21–26.
Shima, S., Matsuoka, H., Iwamoto,
T., and Sakai, H. (1984) Antimicrobial action of epsilon-poly-L-lysine.
J. Antibiot. (Tokyo)., 37, 1449–1455.
FDA (2004) Agency Response Letter
GRAS. Notice No. GRN 000135.
Najjar, M.B., Kashtanov, D., and
Chikindas, M.L. (2009) Natural antimicrobials ε-Poly-l-lysine and Nisin A for
control of oral microflora. Probiotics
Antimicrob. Proteins, 1, 143–147.
Blackburn, P. and Goldstein, B. P.
(1995) Nisin compositions to prevent
the promotion of tooth decay by suppressing formation of acid from foods
by oral bacteria, International Patent
References
84.
85.
86.
87.
88.
89.
90.
91.
92.
Application WO 97/10801, Applied
a combination of various lantibiotic
Microbiology, Inc.
modification enzymes. ACS Synth. Biol.,
McConville, P. (1996) A chewing gum
2, 397–404.
composition containing a bacteriocin
93. Qiu, X.Q., Wang, H., Lu, X.F., Zhang,
antibacterial agent, International Patent
J., Li, S.F., Cheng, G., Wan, L., Yang,
Application WO 97/06772, AMBI, Inc.,
L., Zuo, J.Y., Zhou, Y.Q., Wang, H.Y.,
SmithKline Beecham Plc.
Cheng, X., Zhang, S.H., Ou, Z.R.,
Tong, Z., Dong, L., Zhou, L., Tao, R.,
Zhong, Z.C., Cheng, J.Q., Li, Y.P., and
and Ni, L. (2010) Nisin inhibits dental
Wu, G.Y. (2003) An engineered mulcaries-associated microorganism in
tidomain bactericidal peptide as a
vitro. Peptides, 31, 2003–2008.
model for targeted antibiotics against
Howell, T.H., Fiorellini, J.P., Blackburn,
specific bacteria. Nat. Biotechnol., 21,
P., Projan, S.J., de la Harpe, J., and
1480–1485.
Williams, R.C. (1993) The effect of
94. Qiu, X.Q., Zhang, J., Wang, H., and
a mouthrinse based on nisin, a bacWu, G.Y. (2005) A novel engineered
teriocin, on developing plaque and
peptide, a narrow-spectrum antibiotic,
gingivitis in beagle dogs. J. Clin. Periis effective against vancomycin-resistant
odontol., 20, 335–339.
Enterococcus faecalis. Antimicrob.
Yuzenkova, J., Delgado, M., Nechaev, S.,
Agents Chemother., 49, 1184–1189.
Savalia, D., Epshtein, V., Artsimovitch,
95. Janiszewska, J. and UrbanczykI., Mooney, R.A., Landick, R., Farias,
Lipkowska, Z. (2007) Amphiphilic
R.N., Salomon, R., and Severinov, K.
dendrimeric peptides as model non(2002) Mutations of bacterial RNA
sequential pharmacophores with
polymerase leading to resistance to
antimicrobial properties. J. Mol. Micromicrocin j25. J. Biol. Chem., 277,
biol. Biotechnol., 13, 220–225.
50867–50875.
96. Cheng, Y., Xu, Z., Ma, M., and Xu,
Kramer, N.E., van Hijum, S.A., Knol, J.,
T. (2008) Dendrimers as drug carriKok, J., and Kuipers, O.P. (2006) Traners: applications in different routes of
scriptome analysis reveals mechanisms
drug administration. J. Pharm. Sci., 97,
by which Lactococcus lactis acquires
123–143.
nisin resistance. Antimicrob. Agents
97. Ramshaw, J.A., Werkmeister, J.A., and
Chemother., 50, 1753–1761.
Glattauer, V. (1996) Collagen-based
Collins, B., Curtis, N., Cotter, P.D.,
biomaterials. Biotechnol. Genet. Eng.
Hill, C., and Ross, R.P. (2010) The ABC
Rev., 13, 335–382.
transporter AnrAB contributes to the
98.
Misselwitz, B., Schmitt-Willich, H.,
innate resistance of Listeria monocytoEbert, W., Frenzel, T., and Weinmann,
genes to nisin, bacitracin, and various
H.J. (2001) Pharmacokinetics of
beta-lactam antibiotics. Antimicrob.
Gadomer-17, a new dendritic magnetic
Agents Chemother., 54, 4416–4423.
resonance contrast agent. MAGMA, 12,
Knerr, P.J. and van der Donk, W.A.
128–134.
(2012) Discovery, biosynthesis, and
engineering of lantipeptides. Annu. Rev. 99. Dufes, C., Uchegbu, I.F., and Schatzlein,
A.G. (2005) Dendrimers in gene
Biochem., 81, 479–505.
delivery. Adv. Drug Delivery Rev., 57,
Montalban-Lopez, M., Zhou, L.,
2177–2202.
Buivydas, A., van Heel, A.J., and
100. Thankappan, U.P., Madhusudana, S.N.,
Kuipers, O.P. (2012) Increasing the
Desai, A., Jayamurugan, G., Rajesh,
success rate of lantibiotic drug discovY.B., and Jayaraman, N. (2011) Denery by Synthetic Biology. Expert Opin.
dritic poly(ether imine) based gene
Drug Discovery, 7, 695–709.
delivery vector. Bioconjugate Chem., 22,
van Heel, A.J., Mu, D., Montalban115–119.
Lopez, M., Hendriks, D., and Kuipers,
101. Lequin, O., Bruston, F., Convert, O.,
O.P. (2013) Designing and producChassaing, G., and Nicolas, P. (2003)
ing modified, new-to-nature peptides
with antimicrobial activity by use of
Helical structure of dermaseptin B2
89
90
5 Antimicrobial Peptides from Prokaryotes
in a membrane-mimetic environment.
Biochemistry, 42, 10311–10323.
102. Papo, N. and Shai, Y. (2003) Exploring
peptide membrane interaction using
surface plasmon resonance: differentiation between pore formation versus
membrane disruption by lytic peptides.
Biochemistry, 42, 458–466.
103. Powers, J.P. and Hancock, R.E. (2003)
The relationship between peptide structure and antibacterial activity. Peptides,
24, 1681–1691.
91
6
Peptidomimetics as Antimicrobial Agents
Peng Teng, Haifan Wu, and Jianfeng Cai
6.1
Introduction
The abuse of traditional antibiotics has led to the significantly increased
incidence of antibiotic resistance in recent years [1]. The World Health Organization recently specifically identified antimicrobial resistance as one of the three
greatest threats facing mankind in the twenty-first century. Therefore, the generation of new antibiotics with novel mechanisms is urgently needed to treat bacterial
infections that are unresponsive to conventional antibiotics, as many bacteria
(e.g., methicillin-resistant Staphylococcus aureus (MRSA)) cause severe, and often
lethal, community-, and hospital-acquired infections [1, 2]. More noticeably, those
drug-resistant Gram-negative pathogens, such as Pseudomonas aeruginosa, pose
even more severe threats because they are more dangerous and much harder to
kill than Gram-positive bacterial strains. As such, over the last decade, significant
interest has been developed in the exploitation of cationic host defense peptides
(HDPs) as a new generation of antibiotic agents [1, 2]. HDPs, frequently referred
as antimicrobial peptides (AMPs), are small cationic amphiphilic peptides found
in virtually all living organisms [2]. They are an ancient and vital part of the innate
immune system, and play a key role in the defense against bacterial infections
[1–5]. Compared with conventional antibiotics, which target specific metabolic
processes [6], HDPs are able to form globally amphipathic structures, in which
cationic and hydrophobic groups are segregated into two regions, facilitating
interaction with the negatively charged bacterial membranes [7]. However, the
exact mechanism of action of HDPs is still controversial and requires further
exploration. For instance, the implication of cationicity of HDPs is currently under
debate. The cationic charge of HDPs was previously believed to be the determining
factor for the selectivity of HDPs against bacteria over mammalian cells. This is
because bacterial membranes are overall negatively charged owing to the presence
of teichoic acids in the cell wall of Gram-positive bacteria and lipopolysaccharides (LPSs) in Gram-negative bacteria [1–5]. In contrast, the outer layer of
mammalian cell membranes is composed of zwitterionic phosphatidylcholines
[7], resulting in much weaker electrostatic interaction with cationic HDPs.
Novel Antimicrobial Agents and Strategies, First Edition.
Edited by David A. Phoenix, Frederick Harris and Sarah R. Dennison.
© 2015 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2015 by Wiley-VCH Verlag GmbH & Co. KGaA.
92
6 Peptidomimetics as Antimicrobial Agents
However, cationicity alone could not explain the bactericidal activity and strain
selectivity of HDPs according to recent findings [8]. In addition, different modes
of interaction can occur after HDPs dock on bacterial membranes, and a few
models have been proposed, including carpet-like, barrel stave, toroidal channels,
and peptide–lipids aggregates [4, 9–15]. Nevertheless, these interactions all ultimately compromise the physical integrity of the lipid membranes, leading to the
loss of membrane function and bacterial cell death. Furthermore, HDPs can also
translocate into the cells directly and interact with internal targets that are critical
to bacterial signaling pathway. Such mechanisms of action may account for the
structural similarity and diversity of HDPs. Although HDPs have a variety of
secondary structures such as α-helices, β-sheets, and extended conformation [1],
they are overall cationic charged, containing 15–50 residues and sharing globally
amphipathic conformations. Examples of HDPs include defensin [16–18], magainin [19–22], and indolicidin [23–27], which have different secondary structures.
Since interactions between HDPs and bacterial membranes are physical interactions and lack of specific cell wall targets, and their precise sequences are less
critical, it is believed that HDPs are less prone to induce resistance in bacteria
observed for current antibiotic treatments [2], because the circumnavigation
of physical membrane disruption is minimally observed. Therefore, HDPs are
promising therapeutics as a new generation of antibiotics with novel mechanisms.
Despite significant enthusiasm, there are intrinsic drawbacks associated with
the development of HDP antibiotics, which include their common susceptibility
to enzymatic degradation, moderate activity, and their inconvenient optimization
[28–30]. Peptidomimetics rise either from modifications of canonical peptides
by introducing non-natural side chains, by including amide bond isosteres via
the introduction of heteroatoms, or by designing non-natural systems that are
similar to peptides. They are generally highly resistant to proteolytic degradation because they are not well recognized by enzymes [31–33]. It is therefore
envisioned that peptidomimetics that mimic the structure and mechanism of
action of HDPs can potentially overcome the intrinsic drawbacks associated with
HDPs. In addition, there are enormous chemically diverse functional groups
that can be introduced into peptidomimetics, and as such it is promising to
identify and develop lead compounds that are more potent than HDPs. In the
last decade, there has been tremendous effort attributed to the development
of new antimicrobial peptidomimetics, including peptoids [7], β-peptides [34],
arylamide oligomers [35], and AApeptides [1, 2, 36–40]. In this chapter, we
review the recent development of synthetic mimics of HDPs (since 2009) based
on a variety of backbones other than canonical peptides. Some cyclic peptides
[41, 42], although they are potent antimicrobial agents, are not discussed here
as most of their residues are canonical amino acids. Some other HDP mimics
are based on non-peptidic scaffolds including carbohydrates [43], aryls [44, 45],
dendrimers [46], and polymers [47], which are also not included because this
chapter focuses on antimicrobial peptidic oligomers.
6.2
Antimicrobial Peptidomimetics
93
6.2
Antimicrobial Peptidomimetics
6.2.1
Peptoids
Peptoids are unnatural oligomers of N-substituted glycines [48]. Although they
have the same backbone as α-peptides, their side chains are attached to the backbone amide nitrogen rather than to the alpha carbon. As a result, they cannot
form intramolecular hydrogen bonds that are critical for the helical conformation
of α-peptides. However, peptoids can still form helical structure by incorporating
chiral side chains [49], which provides the basis for the rational design of amphipathic antimicrobial peptoids mimicking HDPs (Figure 6.1a). Similar to HDPs,
amphipathic peptoids with a proper balance of cationic charge and hydrophobicity can be potential antimicrobial agents [7]. It is found that antimicrobial peptoids
have good selectivity, with limited toxicity to mammalian cells [7]. Peptoids also
display excellent antifouling activity. They can be fabricated onto the solid surface,
and were found to damage Escherichia coli membranes rapidly [50]. Antimicrobial
peptoids can also affect the phenotype of fungi Candida albicans [51]. It is known
that C. albicans can undergo phenotypic switching between an invasive multicellular form and a latent unicellular type. After administration of an antimicrobial
peptoid (Figure 6.1a) [7], the invasive multicellular phenotype was significantly
(a)
(b)
(c)
H-(NLys-Nspe-Nspe)4-NH2
H-Ntridec-(NLys-Nspe-Nspe-NLys)-NH2
c(NapNdp)3
N
N
O
N
O
O
NH2
NLys, N-(4-aminobutyl)glycine
N
Nspe, (S)-N-(1-phenylethyl)glycine
Ntridec, N-(tridecyl)glycine
N
O
O
NH2
Nap, N-(3-aminopropyl)glycine
Ndp, N-(2,2-diphenylethyl)glycine
Figure 6.1 Antimicrobial peptoids [7, 53–56]: (a) a helical peptoid; (b) a short peptoid with
an alkyl tail; and (c) a cyclic peptoid.
94
6 Peptidomimetics as Antimicrobial Agents
suppressed, and the latent unicellular type was favored [51]. To assess the potential
of antimicrobial peptoids as bioavailable antibiotics, the in vivo pharmacokinetics
of peptoids was studied [52]. The 64 Cu-labeled 1,4,7,10-tetraazacyclododecane1,4,7,10-tetraacetic acid (DOTA) was conjugated to peptoids and the resulting
compounds were analyzed by positron emission tomography (PET). It is interesting that the amphipathicity of peptoids significantly affects their biodistributions,
implying the potential for the rational design of peptoids for the specific delivery
of therapeutic agents to a specific tissue or organs [52].
On the basis of the development of helical antimicrobial peptoids and
structure–function relationship studies [7], Barron et al. [53] found that short
peptoids with alkyl tails could still possess potent and broad-spectrum antimicrobial activity, even though no helical structures are formed. For instance,
an alkylated short peptoid (Figure 6.1b) is comparable to the helical peptoid
(Figure 6.1a) and superior to the synthetic HDP Pexiganan (a synthetic analog of
magainin II) [53]. This alkylated cationic peptoid, as well as the helical peptoid,
shows significant reductions in biomass and cell viability of P. aeruginosa biofilms
[54]. More surprisingly, the alkylated short peptoid was even more potent than
the helical peptoid against tuberculosis, and was most selective [56]. In contrast,
the unalkylated peptoids did not exhibit any activity against Mycobacterium
tuberculosis, further demonstrating the importance of alkylation in antimicrobial peptoids [56]. In addition to helical peptoids and alkylated peptoids, the
antimicrobial activity of cyclic peptoids was also investigated. The head-to-tail
cyclization of peptoids (Figure 6.1c) yielded a few cyclic peptoids that display
enhanced antibacterial activity against MRSA strains (twofold more potent than
linear ones) [55]. It is hypothesized that macrocyclization can pre-organize side
chain groups of peptoids and force them to project toward different faces of
the planar macrocycle in a predicted manner. Therefore, a globally amphiphilic
structure can be formed with enhanced stability, which may improve the ability
of cyclic peptoids for the interaction and disruption of bacteria membranes.
The synergistic effect of peptoids co-administered with HDPs was also studied.
It is interesting that simultaneous administration of both antimicrobial peptoids
and HDPs could further boost the antimicrobial efficacy [57], which opens a new
avenue for the development of HDP-based antibiotics.
6.2.2
𝛃-Peptides
β-peptides are oligomers consisting of β-amino acids, and they have been extensively studied for their ability to mimic the structure and function of peptides
[58, 59]. Similar to other classes of peptidomimetics, β-peptides are completely
resistant to proteolytic degradation. Furthermore, β-peptides exhibit strong folding propensity and can fold into a variety of well-defined secondary structures
including “12-helices” and “14-helices” [60]. As such, the early design of antimicrobial β-peptides was based on their helical conformations, which display segregated amphipathic cationic charges and hydrophobic surfaces (Figure 6.2) [59, 61].
6.2
H-(β3-X-β3-HLys-β3-HLeu)n-OH
Antimicrobial Peptidomimetics
95
n=2-6; X = HVal or HLeu
NH2
O
O
OH
H2N
β3-HAla
(a)
O
H
N
O
H
N
O
β3-HLeu
H
N
N
H2 O
⊕
H
N
O
β3-HLys
H
N
N
H2 O
⊕
OH
H2N
OH
H 2N
β3-HVal
H
N
(b)
OH
H2N
O
O
H
N
O
NH2
O
Figure 6.2 Helical antimicrobial β-peptides: (a) a 14-helix [61] and (b) a 12-helix [59].
The resulting β-peptides are expected to mimic globally amphiphilic conformations of HDPs and display potent and broad-spectrum antimicrobial activities.
Gellman et al. [62] soon after discovered that a helical structure is not necessary
for potent antimicrobial activity, as they observed that a sequence containing both
α and β residues does not show good folding propensity; however, it has excellent
antimicrobial activity. The findings have been used widely for the design of new
antimicrobial peptidomimetics later on.
A recent report on antimicrobial β-peptides focused on their activity against
C. albicans. C. albicans is one of the most prevalent fungal strains found in
humans, and this pathogen has been highly unresponsive to HDPs [63]. To
develop β-peptides that can kill C. albicans effectively, a focused library of
14-helical β-peptides was prepared. Their structure–function relationship was
studied, which reveals that the globally amphiphilic conformation is critical for
the antifungal activity of β-peptides. The lead helical β-peptide (Figure 6.3) was
found to have potent antifungal activity under physiological conditions [63]. In
addition to killing C. albicans cells, this class of β-peptides can even prevent the
formation of biofilms by C. albicans cells. C. albicans biofilms have been found
to contaminate the surfaces of medical device and to induce recurrent infections
[64]. They are also very resistant to drug treatment. As such, prevention of
C. albicans is highly significant. To gain insight on how β-peptides kill planktonic
C. albicans cells as well as biofilms, mechanistic studies were carried out by using
fluorescently labeled β-peptides, which indicate that β-peptides can disrupt cell
membranes of C. albicans and accumulate in the cytoplasm of both planktonic
and biofilm-forming cells [64]. This observation is consistent with the general bactericidal mechanism of HDPs and antimicrobial peptidomimetics, even though
C. albicans is a fungus rather than a bacterium. Nonetheless, the disruption of
C. albicans cell membranes and the cell entry of the β-peptides directly led to the
cell death. These β-peptides were then fabricated into multilayered polyelectrolyte
films to arrest the growth of C. albicans and its biofilm formation [65].
3
96
6 Peptidomimetics as Antimicrobial Agents
β3Tyr-(ACHC-β3Val-β3Lys)3
HO
O
O
N
H
N
H
β3Tyr
Trans-2-aminocyclohexanecarboxylic acid (ACHC)
NH2
O
O
N
H
N
H
β3Lys
β3Val
Figure 6.3 The helical antifungal β-peptide [63].
6.2.3
Arylamides
The development of peptidomimetics that mimic the structure and function
of HDPs has largely overcome the issues associated with HDPs including low
stability and moderate activity. However, the synthesis and scale-up of these
agents may still be obstacles for the development of the next generation of
antibiotics. Therefore, design of molecules with smaller molecular weight
(∼1000 Da) but still capturing the bactericidal mechanism and function of HDPs
may be more appealing. In this regard, Degrado et al. [35] have developed small
arylamide foldamers that mimic the amphipathic structures of HDPs (Figure 6.4).
The arylamide scaffold adopts an almost planar conformation stabilized by
intramolecular hydrogen bonding; thus, hydrophobic and hydrophilic groups
can be appended onto either side to make the global amphipathic structures.
The structure–function relationship and lead optimization of arylamides are
summarized in a recently published review [60]. Certain arylamides exhibited
potent inhibitions toward a number of Gram-positive and Gram-negative strains,
R1
H2N
H
N
NH
R2
Y
H
N
H
N
O
O
R3
X
X
R2
Y
R1
H
N
H
N
H
N
O
O
NH2
NH
R3
Figure 6.4 Antimicrobial arylamides. X is either N or H, Y can be S or H, R1 and R2 are different cationic groups, and R3 represents hydrophobic groups [35, 66, 67].
6.2
Antimicrobial Peptidomimetics
and the lead compounds even successfully arrested the growth of S. aureus in a
mouse thigh burden infection model [35].
Indeed, arylamides are the most successful antimicrobial peptidomimetics that
mimic HDPs. One lead compound PMX-30063 is currently in a phase II clinical
trial for the treatment of skin infections by MRSA strains. The initial results based
on 80 patients are very favorable, demonstrating the effectiveness and minimal
side effects of PMX-30063 treatment, with a maximum tolerated single dose at
2.5 mg kg−1 . Unlike conventional antibiotics, PMX-30063 not only has broader
spectrum activity against a variety of Gram-positive and Gram-negative bacteria
it also does not elicit bacterial resistance even up to 20 serial passages. In addition
to PMX-30063, a few of the other arylamide derivatives have also shown promising activities toward oral Candida strains and biofilms of periodontal pathogens,
and their clinical trials are expected to be conducted in the near future [66, 67].
6.2.4
𝛃-Peptoid–Peptide Hybrid Oligomers
The fact that α/β chimeric peptidomimetics [62] possess potent antimicrobial
activity attracted the study of other antimicrobial peptidomimetics with heterogeneous backbones. One recent example is the development of β-peptoids
[68, 69] and the investigation of the antibacterial activity of β-peptoid–peptide
hybrid oligomers [70]. The building blocks of these sequences consist of cationic
amino acid residues Lys or Arg, and a hydrophobic β-peptoid residue (Figure 6.5).
Therefore, these build blocks can be viewed as amphiphilic. Although the
secondary folding conformations of these hybrids are not further discussed,
these sequences are expected to form a globally amphiphilic structure when
interacting with bacterial membranes. Indeed, these molecules also display
broad-spectrum activity against both bacterial strains and C. albicans. The
structure–function relationship studies reveal that Arg residues are more critical
to boost antimicrobial activity than Lys residues. These β-peptoid–peptide hybrid
oligomers are another class of antimicrobial peptidomimetics with heterogeneous
O
H
N
O
N
R1
NH2
R1 =
O
NH2
n
R2
R2 =
NH
N
H
NH2
Figure 6.5 Antimicrobial β-peptoid–peptide hybrid oligomers [70].
97
98
6 Peptidomimetics as Antimicrobial Agents
backbones, which shed light on the development of new antimicrobial agents
based on currently available molecular frameworks.
6.2.5
Oligourea and 𝛄4 -Peptide-Based Oligomers
It is interesting to understand how the molecular frameworks impact the
antimicrobial activity of peptidomimetics. To address this point, Guichard et al.
[71] prepared a few helical peptidomimetics with similar secondary structures,
including oligoureas, γ4 -peptides, and oligomers with heterogeneous backbones
(Figure 6.6). However, they have shown distinct antimicrobial activity, even
though their side chains are identical. The researchers concluded that the
antimicrobial activity of peptidomimetics is not only related to their amphipathic
conformations but also to their backbones. Polar backbones may promote the
antimicrobial activity by increasing the strength of the interaction with negatively
charged bacterial membranes.
6.2.6
AApeptides
AApeptides are a new class of peptidomimetics developed in our group recently
in order to expand the application of peptidomimetics in chemical biology
[72–79]. They are termed AApeptides because they are composed of N-acylatedN-aminoethyl amino acid units that are derived from the chiral PNA backbone.
Depending on the relative positions of the side chains, two subclasses of AApeptides, α-AApeptides, and γ-AApeptides, have been developed and studied for
their biological activity (Figure 6.7).
Compared with natural α-peptides, the repeating unit (building block) of
AApeptides is comparable to a dipeptide residue. As such, AApeptides project
an identical number of side groups as conventional peptides of the same lengths.
Since half of the side chains come from any carboxylic acids, the potential of
generating AApeptides with chemically diverse functional groups is limitless.
We anticipate that AApeptides would be promising candidates for the design
of antimicrobial agents, not only because of their high stability [75, 80] and
unlimited potential for derivatization [75, 81] but also as a result of their balanced
conformational backbone rigidity and flexibility [36, 37]. In addition, AApeptides
have two more dihedral angles compared to that of α-peptides, resulting in slightly
increased flexibility. As such, we hypothesize that antimicrobial AApeptides
can be designed simply by joining amphiphilic AApeptide building blocks
(containing one hydrophobic and one cationic side chain) together, because such
O
H
N
n=8
X
R
n
Figure 6.6 Antimicrobial oligoureas, γ4 -peptides, and the
related heterogeneous oligomers [71].
6.2
R
N
H
O
H
N
α-peptide
O
R
R
O
O
N
N
H
Antimicrobial Peptidomimetics
α-AApeptide
R
R
N
H
O
γ-AApeptide
N
O
R
Figure 6.7 Structures of α-peptides, α-AApeptides, and γ-AApeptides.
Amphipathic AApeptide
building block
(a)
Cationic group
Hydrophobic group
Membrane
interaction
(b)
Random conformation
Globally amphipathic
conformation
Figure 6.8 Illustration of the mechanism of antimicrobial AApeptide: (a) schematic representation of an amphiphilic AApeptide building block and (b) the amphipathic conformation
of AApeptides upon interaction with bacterial membranes.
AApeptides are expected to adopt overall amphipathic structures by adjusting
their conformation after they sit on bacterial membranes (Figure 6.8). Fine tuning
of their activity and selectivity is readily achieved by adjusting the nature of the
hydrophobic and hydrophilic groups. On the basis of this hypothesis, different
subclasses of antimicrobial AApeptides (linear, lipo-linear, cyclic, lipo-cyclic)
have been developed by our group, all of which can mimic the mechanism
of HDPs by disrupting bacterial membranes [36, 37, 72, 76, 82, 83] and show
broad-spectrum activity against a range of Gram-positive and Gram-negative
bacteria (Figure 6.9).
6.2.6.1 𝛂-AApeptides
To test the design of antimicrobial AApeptides, we initially prepared a focused
library of α-AApeptides composed of amphiphilic building blocks [74]. One
α-AApeptide 𝛂1, containing seven building blocks and comparable to a
14-mer peptide in length (Figure 6.9), showed most potent activity to inhibit the
growth of E. coli (Gram-negative), Bacillus subtilis (Gram-positive), as well as the
99
100
6 Peptidomimetics as Antimicrobial Agents
NH2
H2N
NH2
H
N
N
O
O
NH2
H
N
N
O
NH2
H
N
N
O
H
N
H
N
O
NH2
O
H2N
H
N
N
O
NH2
O
H
N
N
NH2
O
O
O
O
NH2
NH2
O
O
O
O
O
H
N
N
O
O
α2
O
O
H
N
N
O
NH2
N
O
NH2
O
H
N
γ1
NH2
N
O
O
NH2
NH2
H
N
H
N
NH2
N
O
O
N
α1
NH2
N
O
H
N
N
H
N
N
H
N
O
NH2
O
H
N
N
O
O
NH2
N
O
H
N
NH2
O
H
N
O
N
O
NH2
N
O
O
N
O
H
N
N
O
O
NH2
O
O
H
N
N
NH2
H
N
N
O
NH2
O
O
H2N
O
H
N
N
NH2
N
O
O
γ2
O
H
N
N
O
NH2
O
N
O
HN
NH2
γ3
NH
O
N
N
O
N
H
O
O
O
N
O
O
N
H
NH2
NH2
Figure 6.9 The structures of lead antimicrobial AApeptides, including α-AApeptide 𝛂1 and
𝛂2 [74, 84], and γ-AApeptide 𝛄1, 𝛄2, and 𝛄3 [72, 76, 83].
6.2
Antimicrobial Peptidomimetics
multidrug-resistant Staphylococcus epidermidis (Gram-positive). Follow-up SEM
and fluorescence microscopy studies suggested that the bactericidal activity of 𝛂1
may originate from the disruption of bacterial membranes. It is noted that 𝛂1 is
highly selective, and has only negligible hemolytic activity even up to 250 μg ml−1 .
The results also suggest that the size of α-AApeptides is critical since shorter
sequences exhibit minimal antimicrobial activity. The initial studies demonstrated
that the AApeptide backbone is very suitable for antimicrobial development.
Lipopeptide antibiotics have been widely studied in clinical trials and
commercialized in market [85, 86]. Although the mechanism of their antimicrobial functions are different from that of HDPs [87, 88], it is widely accepted
that lipidation can enhance the ability of peptides to interact with bacterial
membranes through membrane insertion. Therefore, we expected that lipidation
of linear α-AApeptides would lead to more potent antimicrobial agents. As such, a
few lipidated α-AApeptides were designed and tested for their capability to arrest
the growth of bacterial strains [76]. One lead lipo-α-AApeptide, α2 (Figure 6.9),
displays potent and broad-spectrum activity against both Gram-positive and
Gram-negative bacteria, including multidrug-resistant strains.
6.2.6.2 𝛄-AApeptides
As γ-AApeptides have a scaffold similar to α-AApeptides, it was speculated that
the same rule employed for the design of antimicrobial α-AApeptides could be
used to direct the development of γ-AApeptides that mimic HDPs. Indeed, with
the same size and slightly different hydrophobic groups, γ-AApeptide 𝛄1 has
shown much improved potency and broader spectrum activity than α-AApeptide
𝛂1 [76]. It also effectively kills the fungus C. albicans. In addition, it does not
induce bacterial resistance even after 17 serial passages, suggesting its potential
as a new class of antimicrobial peptidomimetics.
Akin to lipidated-α-AApeptides, we also explored the ability of lipo-γAApeptides as potential antimicrobial agents [76]. Interestingly, some of these
lipo-peptidomimetics are even better antibiotic agents than 𝛄1. The lead lipo-γAApeptide 𝛄2, is very active against a range of Gram-positive and Gram-negative
bacterial strains including clinically relevant pathogens, as well as fungus
C. albicans. Subsequent fluorescence microscopy, drug-resistance study, and
membrane depolarization again suggest that lipo-γ-AApeptides capture the
mechanism of action of HDPs by damaging bacterial membranes. In addition,
γ-AApeptides with unsaturated lipid tails generally show less hemolytic activity
than γ-AApeptides with saturated alkyl tails, whereas they do not compromise
the antimicrobial activity. The mechanism still remains elusive; however, the
observation can be used for the future design of different classes of antimicrobial
peptidomimetics.
The recent development of antimicrobial γ-AApeptides involves the design
and evaluation of a library of lipo-cyclic γ-AApeptides [83]. Lipo-cyclic peptide
antibiotics such as daptomycin [89] and polymyxins [90] have been used as the
last resort to treat drug-resistant pathogens. Daptomycin is a negatively charged
lipo-cyclic peptide, and it binds to Gram-positive bacterial membranes, causing
101
102
6 Peptidomimetics as Antimicrobial Agents
membrane depolarization. However, daptomycin can only weakly associate with
membranes of Gram-negative bacteria, and thus exhibits almost no activity
toward these pathogens. Polymyxins bind to LPSs in the outer membrane
of Gram-negative bacteria and utilize their hydrophobic tails to damage the
membranes [85]. However, the tails cannot penetrate the thick peptidoglycan
layer in Gram-positive bacteria, and thus polymyxins are not active against
Gram-positive bacterial pathogens. On the basis of the cationic amphipathic
structures of γ-AApeptides, we anticipated that lipo-cyclic γ-AApeptides would
mimic the mechanism of HDPs rather than the above-mentioned lipo-cyclic
peptide antibiotics. As expected, the lead lipo-cyclic γ-AApeptide 𝛄3 effectively
arrests a number of drug-resistant Gram-positive and Gram-negative pathogens
[83]. More noticeably, this class of peptidomimetics can even mimic HDPs [91]
by exhibiting dual functional roles in fighting bacterial infections: they can kill
bacteria directly, and they can also suppress pro-inflammatory cytokines such as
TNF-α. The findings open a new avenue for the development of antimicrobial
peptidomimetics.
6.3
Discussion
Significant advances are achieved in the past several years since the development
of antimicrobial peptidomimetics a decade ago. These peptidomimetics have
cationic amphipathic conformations and mechanism of action similar to HDPs.
The most successful example is arylamide foldamers developed by DeGrado et al.,
which are currently in Phase II clinical trials against MRSA infection. Some other
lead compounds from this series of foldamers have also shown promising efficacy
against bacteria, fungi, virus, and even tropical diseases. It is therefore speculated
that antimicrobial peptidomimetics are promising antibiotics with a novel mechanism to combat drug-resistant bacteria. Future direction for the development
of antimicrobial peptidomimetics could at least focus on the following aspects.
First, the functional groups on the peptidomimetics could be further diversified.
For those examples discussed in this chapter, so far only a few hydrophobic
and hydrophilic groups are explored. One of the most attractive features of
peptidomimetics is the enormous chemical diversity. It is therefore envisioned
that the activity and selectivity of antimicrobial peptidomimetics can be further
improved if a wide variety of hydrophobic, bulky, and hydrophilic groups are
introduced. In addition, it is plausible to design antimicrobial peptidomimetics
with new scaffolds so as to identify more potent antimicrobial agents. Each class
of peptidomimetics has its own framework, which can differ in the backbone
polarity, as well as precise three dimensional structures, which significantly
impact the strength of their interactions with bacteria. Moreover, the selectivity
of antimicrobial peptidomimetics is expected to be further enhanced. It is generally realized that potent antimicrobial peptidomimetics are always accompanied
with visible hemolytic activity and cytotoxicity. However, this problem can be
References
mitigated by varying the ratio and nature of hydrophobic and hydrophilic chains
in peptidomimetics. Emphasis should be given to the identification of more
potent antimicrobial peptidomimetics to kill Gram-negative bacteria such as
P. aeruginosa. It is widely accepted that the prevention of infections caused by
Gram-negative bacteria is more difficult than that of Gram-positive bacteria.
Furthermore, the immunomodulatory function of antimicrobial peptidomimetics
should be considered and elaborated in parallel in the future, since HDPs may
defend organisms by suppressing inflammatory responses to avoid septic shock,
in addition to direct bacterial killing. Peptidomimetics bearing the dual functions
may hold greater promise for antimicrobial development. Last, but not the least,
more systematic and comprehensive mechanistic studies should be conducted
for antimicrobial peptidomimetics in order to evaluate their potential for the
mimicry of HDPs and to assess their probability to induce drug resistance
in bacteria. Such work will provide persuasive proof for the development of
antimicrobial peptidomimetics as the next generation of antibiotics with novel
mechanisms.
Acknowledgments
The work is supported by the USF start-up fund.
References
1. Hancock, R.E., and Sahl, H.G. (2006)
Antimicrobial and host-defense
peptides as new anti-infective therapeutic strategies. Nat. Biotechnol., 24,
1551–1557.
2. Marr, A.K., Gooderham, W.J., and
Hancock, R.E. (2006) Antibacterial peptides for therapeutic use: obstacles and
realistic outlook. Curr. Opin. Pharmacol.,
6, 468–472.
3. Zhang, L., Dhillon, P., Yan, H.,
Farmer, S., and Hancock, R.E. (2000)
Interactions of bacterial cationic peptide
antibiotics with outer and cytoplasmic
membranes of Pseudomonas aeruginosa. Antimicrob. Agents Chemother., 44,
3317–3321.
4. Wu, M., Maier, E., Benz, R., and
Hancock, R.E.W. (1999) Mechanism of
interaction of different classes of cationic
antimicrobial peptides with planar bilayers and with the cytoplasmic membrane
of Escherichia coli. Biochemistry, 38,
7235–7242.
5. Friedrich, C.L., Moyles, D., Beveridge,
6.
7.
8.
9.
T.J., and Hancock, R.E. (2000) Antibacterial action of structurally diverse cationic
peptides on Gram-positive bacteria.
Antimicrob. Agents Chemother., 44,
2086–2092.
Alekshun, M.N. and Levy, S.B. (2007)
Molecular mechanisms of antibacterial multidrug resistance. Cell, 128,
1037–1050.
Chongsiriwatana, N.P., Patch, J.A.,
Czyzewski, A.M., Dohm, M.T., Ivankin,
A., Gidalevitz, D., Zuckermann, R.N.,
and Barron, A.E. (2008) Peptoids that
mimic the structure, function, and
mechanism of helical antimicrobial peptides. Proc. Natl. Acad. Sci. U.S.A., 105,
2794–2799.
Lehrer, R.I. and Lu, W.Y. (2012) alphaDefensins in human innate immunity.
Immunol. Rev., 245, 84–112.
Shai, Y. (2002) Mode of action of membrane active antimicrobial peptides.
Biopolymers, 66, 236–248.
103
104
6 Peptidomimetics as Antimicrobial Agents
10. Matsuzaki, K. (1998) Magainins as
11.
12.
13.
14.
15.
16.
17.
18.
19.
paradigm for the mode of action of pore
forming polypeptides. Biochim. Biophys.
Acta, 1376, 391–400.
Brogden, K.A. (2005) Antimicrobial
peptides: pore formers or metabolic
inhibitors in bacteria? Nat. Rev. Microbiol., 3, 238–250.
Matsuzaki, K., Sugishita, K., and
Miyajima, K. (1999) Interactions of
an antimicrobial peptide, magainin 2,
with lipopolysaccharide-containing liposomes as a model for outer membranes
of Gram-negative bacteria. FEBS Lett.,
449, 221–224.
Sal-Man, N., Oren, Z., and Shai, Y.
(2002) Preassembly of membraneactive peptides is an important factor
in their selectivity toward target cells.
Biochemistry, 41, 11921–11930.
Oren, Z., Ramesh, J., Avrahami, D.,
Suryaprakash, N., Shai, Y., and Jelinek,
R. (2002) Structures and mode of membrane interaction of a short alpha helical
lytic peptide and its diastereomer determined by NMR, FTIR, and fluorescence
spectroscopy. Eur. J. Biochem., 269,
3869–3880.
Sitaram, N. and Nagaraj, R. (2002)
Host-defense antimicrobial peptides:
importance of structure for activity.
Curr. Pharm. Des., 8, 727–742.
Lehrer, R.I., Barton, A., Daher, K.A.,
Harwig, S.S., Ganz, T., and Selsted, M.E.
(1989) Interaction of human defensins
with Escherichia coli. Mechanism of
bactericidal activity. J. Clin. Invest., 84,
553–561.
Tang, Y.Q., Yuan, J., Osapay, G., Osapay,
K., Tran, D., Miller, C.J., Ouellette, A.J.,
and Selsted, M.E. (1999) A cyclic antimicrobial peptide produced in primate
leukocytes by the ligation of two truncated alpha-defensins. Science (New York,
N.Y.), 286, 498–502.
Thennarasu, S. and Nagaraj, R. (1999)
Synthetic peptides corresponding to the
beta-hairpin loop of rabbit defensin NP2 show antimicrobial activity. Biochem.
Biophys. Res. Commun., 254, 281–283.
Patch, J.A. and Barron, A.E. (2003) Helical peptoid mimics of magainin-2 amide.
J. Am. Chem. Soc., 125, 12092–12093.
20. Epand, R.F., Umezawa, N., Porter, E.A.,
21.
22.
23.
24.
25.
26.
27.
28.
29.
Gellman, S.H., and Epand, R.M. (2003)
Interactions of the antimicrobial betapeptide beta-17 with phospholipid
vesicles differ from membrane interactions of magainins. Eur. J. Biochem., 270,
1240–1248.
Chen, H.C., Brown, J.H., Morell, J.L., and
Huang, C.M. (1988) Synthetic magainin
analogues with improved antimicrobial
activity. FEBS Lett., 236, 462–466.
Ge, Y., MacDonald, D.L., Holroyd, K.J.,
Thornsberry, C., Wexler, H., and Zasloff,
M. (1999) In vitro antibacterial properties of pexiganan, an analog of magainin.
Antimicrob. Agents Chemother., 43,
782–788.
Falla, T.J. and Hancock, R.E. (1997)
Improved activity of a synthetic indolicidin analog. Antimicrob. Agents
Chemother., 41, 771–775.
Ahmad, I., Perkins, W.R., Lupan, D.M.,
Selsted, M.E., and Janoff, A.S. (1995)
Liposomal entrapment of the neutrophilderived peptide indolicidin endows it
with in vivo antifungal activity. Biochim.
Biophys. Acta, 1237, 109–114.
Ando, S., Mitsuyasu, K., Soeda, Y.,
Hidaka, M., Ito, Y., Matsubara, K.,
Shindo, M., Uchida, Y., and Aoyagi, H.
(2010) Structure-activity relationship
of indolicidin, a Trp-rich antibacterial
peptide. J. Pept. Sci., 16, 171–177.
Nan, Y.H., Bang, J.K., and Shin, S.Y.
(2009) Design of novel indolicidinderived antimicrobial peptides with
enhanced cell specificity and potent
anti-inflammatory activity. Peptides, 30,
832–838.
Subbalakshmi, C. and Sitaram, N. (1998)
Mechanism of antimicrobial action of
indolicidin. FEMS Microbiol. Lett., 160,
91–96.
Giuliani, A. and Rinaldi, A.C. (2011)
Beyond natural antimicrobial peptides:
multimeric peptides and other peptidomimetic approaches. Cell. Mol. Life
Sci., 68, 2255–2266.
Bruhn, O., Grotzinger, J., Cascorbi,
I., and Jung, S. (2011) Antimicrobial peptides and proteins of the
horse – insights into a well-armed
organism. Vet. Res., 42, 98.
References
30. Zaiou, M. (2007) Multifunctional antimi-
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
crobial peptides: therapeutic targets in
several human diseases. J. Mol. Med.
(Berlin, Germany), 85, 317–329.
Tew, G.N., Scott, R.W., Klein, M.L., and
Degrado, W.F. (2009) De novo design
of antimicrobial polymers, foldamers,
and small molecules: from discovery to
practical applications. Acc. Chem. Res.,
43, 30–39.
Scott, R.W., DeGrado, W.F., and Tew,
G.N. (2008) De novo designed synthetic
mimics of antimicrobial peptides. Curr.
Opin. Biotechnol., 19, 620–627.
Tew, G.N., Liu, D., Chen, B., Doerksen,
R.J., Kaplan, J., Carroll, P.J., Klein, M.L.,
and DeGrado, W.F. (2002) De novo
design of biomimetic antimicrobial polymers. Proc. Natl. Acad. Sci. U.S.A., 99,
5110–5114.
Epand, R.F., Raguse, T.L., Gellman, S.H.,
and Epand, R.M. (2004) Antimicrobial
14-helical beta-peptides: potent bilayer
disrupting agents. Biochemistry, 43,
9527–9535.
Choi, S., Isaacs, A., Clements, D., Liu,
D., Kim, H., Scott, R.W., Winkler, J.D.,
and DeGrado, W.F. (2009) De novo
design and in vivo activity of conformationally restrained antimicrobial
arylamide foldamers. Proc. Natl. Acad.
Sci. U.S.A., 106, 6968–6973.
Niu, Y., Wu, H., Li, Y., Hu, Y., Padhee,
S., Li, Q., Cao, C., and Cai, J. (2013)
AApeptides as a new class of antimicrobial agents. Org. Biomol. Chem., 11,
4283–4290.
Niu, Y., Wang, R.E., Wu, H., and Cai,
J. (2012) Recent development of small
antimicrobial peptidomimetics. Future
Med. Chem., 4, 1853–1862.
Goodman, C.M., Choi, S., Shandler, S.,
and DeGrado, W.F. (2007) Foldamers as
versatile frameworks for the design and
evolution of function. Nat. Chem. Biol.,
3, 252–262.
Gellman, S. (2009) Structure and function in peptidic foldamers. Biopolymers,
92, 293.
Wu, Y.D. and Gellman, S. (2008) Peptidomimetics. Acc. Chem. Res., 41,
1231–1232.
Srinivas, N., Jetter, P., Ueberbacher, B.J.,
Werneburg, M., Zerbe, K., Steinmann, J.,
42.
43.
44.
45.
46.
47.
48.
49.
Van der Meijden, B., Bernardini, F.,
Lederer, A., Dias, R.L., Misson, P.E.,
Henze, H., Zumbrunn, J., Gombert, F.O.,
Obrecht, D., Hunziker, P., Schauer, S.,
Ziegler, U., Kach, A., Eberl, L., Riedel,
K., DeMarco, S.J., and Robinson, J.A.
(2010) Peptidomimetic antibiotics target
outer-membrane biogenesis in Pseudomonas aeruginosa. Science (New York,
N.Y.), 327, 1010–1013.
Obrecht, D., Robinson, J.A., Bernardini,
F., Bisang, C., DeMarco, S.J., Moehle,
K., and Gombert, F.O. (2009) Recent
progress in the discovery of macrocyclic
compounds as potential anti-infective
therapeutics. Curr. Med. Chem., 16,
42–65.
Guell, I., Ferre, R., Sorensen, K.K.,
Badosa, E., Ng-Choi, I., Montesinos, E.,
Bardaji, E., Feliu, L., Jensen, K.J., and
Planas, M. (2012) Multivalent display
of the antimicrobial peptides BP100
and BP143. Beilstein J. Org. Chem., 8,
2106–2117.
Thaker, H.D., Sgolastra, F., Clements, D.,
Scott, R.W., and Tew, G.N. (2011) Synthetic mimics of antimicrobial peptides
from triaryl scaffolds. J. Med. Chem., 54,
2241–2254.
Thaker, H.D., Som, A., Ayaz, F., Lui,
D.H., Pan, W.X., Scott, R.W., Anguita, J.,
and Tew, G.N. (2012) Synthetic mimics of antimicrobial peptides with
immunomodulatory responses. J. Am.
Chem. Soc., 134, 11088–11091.
Young, A., Liu, Z.G., Zhou, C.H., and
Kallenbach, N.R. (2009) Dendrimeric
antimicrobial peptides as antibacterials.
Abstracts of Papers of the American
Chemical Society 238.
Kuroda, K. and Caputo, G.A. (2013)
Antimicrobial polymers as synthetic
mimics of host-defense peptides. Wires
Nanomed. Nanobiotechnol., 5, 49–66.
Simon, R.J., Kania, R.S., Zuckermann,
R.N., Huebner, V.D., Jewell, D.A.,
Banville, S., Ng, S., Wang, L., Rosenberg,
S., Marlowe, C.K. et al. (1992) Peptoids:
a modular approach to drug discovery. Proc. Natl. Acad. Sci. U.S.A., 89,
9367–9371.
Armand, P., Kirshenbaum, K.,
Goldsmith, R.A., Farr-Jones, S., Barron,
A.E., Truong, K.T., Dill, K.A., Mierke,
105
106
6 Peptidomimetics as Antimicrobial Agents
50.
51.
52.
53.
54.
55.
56.
57.
D.F., Cohen, F.E., Zuckermann, R.N., and
Bradley, E.K. (1998) NMR determination of the major solution conformation
of a peptoid pentamer with chiral side
chains. Proc. Natl. Acad. Sci. U.S.A., 95,
4309–4314.
Statz, A.R., Park, J.P., Chongsiriwatana,
N.P., Barron, A.E., and Messersmith, P.B.
(2008) Surface-immobilised antimicrobial
peptoids. Biofouling, 24, 439–448.
Uchida, M., McDermott, G., Wetzler, M.,
Le Gros, M.A., Myllys, M., Knoechel, C.,
Barron, A.E., and Larabell, C.A. (2009)
Soft X-ray tomography of phenotypic
switching and the cellular response to
antifungal peptoids in Candida albicans. Proc. Natl. Acad. Sci. U.S.A., 106,
19375–19380.
Seo, J., Ren, G., Liu, H., Miao, Z., Park,
M., Wang, Y., Miller, T.M., Barron, A.E.,
and Cheng, Z. (2012) In vivo biodistribution and small animal PET of
(64)Cu-labeled antimicrobial peptoids.
Bioconjugate Chem., 23, 1069–1079.
Chongsiriwatana, N.P., Miller, T.M.,
Wetzler, M., Vakulenko, S., Karlsson,
A.J., Palecek, S.P., Mobashery, S., and
Barron, A.E. (2011) Short alkylated
peptoid mimics of antimicrobial lipopeptides. Antimicrob. Agents Chemother., 55,
417–420.
Kapoor, R., Wadman, M.W., Dohm,
M.T., Czyzewski, A.M., Spormann,
A.M., and Barron, A.E. (2011) Antimicrobial peptoids are effective against
Pseudomonas aeruginosa biofilms.
Antimicrob. Agents Chemother., 55,
3054–3057.
Huang, M.L., Shin, S.B., Benson, M.A.,
Torres, V.J., and Kirshenbaum, K. (2012)
A comparison of linear and cyclic peptoid oligomers as potent antimicrobial
agents. ChemMedChem, 7, 114–122.
Kapoor, R., Eimerman, P.R., Hardy, J.W.,
Cirillo, J.D., Contag, C.H., and Barron,
A.E. (2011) Efficacy of antimicrobial
peptoids against Mycobacterium tuberculosis. Antimicrob. Agents Chemother.,
55, 3058–3062.
Chongsiriwatana, N.P., Wetzler, M., and
Barron, A.E. (2011) Functional synergy
between antimicrobial peptoids and
58.
59.
60.
61.
62.
63.
64.
65.
66.
peptides against Gram-negative bacteria. Antimicrob. Agents Chemother., 55,
5399–5402.
Cheng, R.P., Gellman, S.H., and
DeGrado, W.F. (2001) beta-peptides:
from structure to function. Chem. Rev.,
101, 3219–3232.
Porter, E.A., Weisblum, B., and Gellman,
S.H. (2002) Mimicry of host-defense
peptides by unnatural oligomers: antimicrobial beta-peptides. J. Am. Chem. Soc.,
124, 7324–7330.
Tew, G.N., Scott, R.W., Klein, M.L., and
Degrado, W.F. (2010) De novo design
of antimicrobial polymers, foldamers,
and small molecules: from discovery to
practical applications. Acc. Chem. Res.,
43, 30–39.
Liu, D. and DeGrado, W.F. (2001) De
novo design, synthesis, and characterization of antimicrobial beta-peptides.
J. Am. Chem. Soc., 123, 7553–7559.
Schmitt, M.A., Weisblum, B., and
Gellman, S.H. (2004) Unexpected
relationships between structure and
function in alpha,beta-peptides: antimicrobial foldamers with heterogeneous
backbones. J. Am. Chem. Soc., 126,
6848–6849.
Karlsson, A.J., Pomerantz, W.C.,
Weisblum, B., Gellman, S.H., and
Palecek, S.P. (2006) Antifungal activity from 14-helical beta-peptides. J. Am.
Chem. Soc., 128, 12630–12631.
Karlsson, A.J., Pomerantz, W.C., Neilsen,
K.J., Gellman, S.H., and Palecek, S.P.
(2009) Effect of sequence and structural
properties on 14-helical beta-peptide
activity against Candida albicans planktonic cells and biofilms. ACS Chem.
Biol., 4, 567–579.
Karlsson, A.J., Flessner, R.M., Gellman,
S.H., Lynn, D.M., and Palecek, S.P.
(2010) Polyelectrolyte multilayers fabricated from antifungal beta-peptides:
design of surfaces that exhibit antifungal activity against Candida albicans.
Biomacromolecules, 11, 2321–2328.
Hua, J., Scott, R.W., and Diamond, G.
(2010) Activity of antimicrobial peptide
mimetics in the oral cavity: II. Activity
against periopathogenic biofilms and
anti-inflammatory activity. Mol. Oral
Microbiol., 25, 426–432.
References
67. Hua, J., Yamarthy, R., Felsenstein,
68.
69.
70.
71.
72.
73.
74.
S., Scott, R.W., Markowitz, K., and
Diamond, G. (2010) Activity of antimicrobial peptide mimetics in the oral
cavity: I. Activity against biofilms of
Candida albicans. Mol. Oral Microbiol.,
25, 418–425.
Olsen, C.A. (2009) Peptoid-peptide
hybrid backbone architectures. ChemBioChem, 11, 152–160.
Olsen, C.A., Lambert, M., Witt, M.,
Franzyk, H., and Jaroszewski, J.W.
(2008) Solid-phase peptide synthesis
and circular dichroism study of chiral
beta-peptoid homooligomers. Amino
Acids, 34, 465–471.
Olsen, C.A., Ziegler, H.L., Nielsen, H.M.,
Frimodt-Moller, N., Jaroszewski, J.W.,
and Franzyk, H. (2010) Antimicrobial,
hemolytic, and cytotoxic activities of
beta-peptoid-peptide hybrid oligomers:
improved properties compared to
natural AMPs. ChemBioChem, 11,
1356–1360.
Claudon, P., Violette, A., Lamour, K.,
Decossas, M., Fournel, S., Heurtault, B.,
Godet, J., Mely, Y., Jamart-Gregoire,
B., Averlant-Petit, M.C., Briand,
J.P., Duportail, G., Monteil, H., and
Guichard, G. (2010) Consequences
of isostructural main-chain modifications for the design of antimicrobial
foldamers: helical mimics of hostdefense peptides based on a heterogeneous amide/urea backbone. Angew.
Chem. Int. Ed., 49, 333–336.
Niu, Y., Padhee, S., Wu, H., Bai, G.,
Harrington, L., Burda, W.N., Shaw,
L.N., Cao, C., and Cai, J. (2011) Identification of gamma-AApeptides with
potent and broad-spectrum antimicrobial activity. Chem. Commun. (Camb.),
47, 12197–12199.
Niu, Y., Jones, A.J., Wu, H., Varani, G.,
and Cai, J. (2011) gamma-AApeptides
bind to RNA by mimicking RNAbinding proteins. Org. Biomol. Chem., 9,
6604–6609.
Padhee, S., Hu, Y., Niu, Y., Bai, G., Wu,
H., Costanza, F., West, L., Harrington,
L., Shaw, L.N., Cao, C., and Cai, J.
(2011) Non-hemolytic alpha-AApeptides
as antimicrobial peptidomimetics. Chem.
Commun. (Camb.), 47, 9729–9731.
75. Niu, Y., Hu, Y., Li, X., Chen, J., and Cai,
76.
77.
78.
79.
80.
81.
82.
83.
84.
J. (2011) [gamma]-AApeptides: design,
synthesis and evaluation. New J. Chem.,
35, 542–545.
Niu, Y., Padhee, S., Wu, H., Bai, G.,
Qiao, Q., Hu, Y., Harrington, L., Burda,
W.N., Shaw, L.N., Cao, C., and Cai, J.
(2012) Lipo-gamma-AApeptides as a
new class of potent and broad-spectrum
antimicrobial agents. J. Med. Chem., 55,
4003–4009.
Bai, G., Padhee, S., Niu, Y., Wang, R.E.,
Qiao, Q., Buzzeo, R., Cao, C., and Cai,
J. (2012) Cellular uptake of an alphaAApeptide. Org. Biomol. Chem., 10,
1149–1153.
Niu, Y., Bai, G., Wu, H., Wang, R.E.,
Qiao, Q., Padhee, S., Buzzeo, R., Cao, C.,
and Cai, J. (2012) Cellular translocation
of a gamma-AApeptide mimetic of tat
peptide. Mol. Pharm., 9, 1529–1534.
Hu, Y., Li, X., Sebti, S.M., Chen, J., and
Cai, J. (2011) Design and synthesis of
AApeptides: a new class of peptide
mimics. Bioorg. Med. Chem. Lett., 21,
1469–1471.
Yang, Y., Niu, Y., Hong, H., Wu, H.,
Zhang, Y., Engle, J., Barnhart, T., Cai, J.,
and Cai, W. (2012) Radiolabeled gammaAApeptides: a new class of tracers for
positron emission tomography. Chem.
Commun., 48, 7850–7852.
Wu, H., Amin, M.N., Niu, Y., Qiao, Q.,
Harfouch, N., Nimer, A., and Cai, J.
(2012) Solid-phase synthesis of gammaAApeptides using a submonomeric
approach. Org. Lett., 14, 3446–3449.
Wu, H., Niu, Y., Padhee, S., Wang,
R.E., Li, Y., Qiao, Q., Ge, B., Cao,
C., and Cai, J. (2012) Design and
synthesis of unprecedented cyclic
gamma-AApeptides for antimicrobial
development. Chem. Sci., 3, 2570–2575.
Li, Y., Smith, C., Wu, H., Padhee, S.,
Manoj, N., Cardiello, J., Qiao, Q.,
Cao, C., Yin, H., and Cai, J. (2013)
Lipidated cyclic gamma-AApeptides
display both antimicrobial and antiinflammatory activity. ACS Chem. Biol.
doi: 10.1021/cb4006613
Hu, Y., Amin, M.N., Padhee, S., Wang,
R.E., Qiao, Q., Bai, G., Li, Y., Mathew,
A., Cao, C., and Cai, J. (2012) Lipidated
107
108
6 Peptidomimetics as Antimicrobial Agents
85.
86.
87.
88.
peptidomimetics with improved antimi- 89. Nailor, M.D. and Sobel, J.D. (2011)
crobial activity. ACS Med. Chem. Lett., 3,
Antibiotics for Gram-positive bacterial
683–686.
infection: vancomycin, teicoplanin, quinupristin/dalfopristin, oxazolidinones,
Vaara, M. (2010) Polymyxins and their
daptomycin, telavancin, and ceftaroline.
novel derivatives. Curr. Opin. Microbiol.,
Med. Clin. North Am., 95, 723–742.
13, 574–581.
Wu, G., Abraham, T., Rapp, J., Vastey,
90. Kvitko, C.H., Rigatto, M.H., Moro, A.L.,
F., Saad, N., and Balmir, E. (2011) Dapand Zavascki, A.P. (2011) Polymyxin
tomycin: evaluation of a high-dose
B versus other antimicrobials for the
treatment strategy. Int. J. Antimicrob.
treatment of pseudomonas aeruginosa
Agents, 38, 192–196.
bacteraemia. J. Antimicrob. Chemother.,
Muraih, J.K., Pearson, A., Silverman, J.,
66, 175–179.
and Palmer, M. (2011) Oligomerization
91. Mookherjee, N., Brown, K.L., Bowdish,
of daptomycin on membranes. Biochim.
D.M.E., Doria, S., Falsafi, R., Hokamp,
Biophys. Acta, 1808, 1154–1160.
K., Roche, F.M., Mu, R.X., Doho,
G.H., Pistolic, J., Powers, J.P., Bryan,
Mishra, N.N., McKinnell, J., Yeaman,
J., Brinkman, F.S.L., and Hancock, R.E.W.
M.R., Rubio, A., Nast, C.C., Chen, L.,
(2006) Modulation of the TLR-mediated
Kreiswirth, B.N., and Bayer, A.S. (2011)
inflammatory response by the endogeIn vitro cross-resistance to daptomycin
nous human host defense peptide LL-37.
and host defense cationic antimicrobial
J. Immunol., 176, 2455–2464.
peptides in clinical methicillin-resistant
staphylococcus aureus isolates. Antimicrob. Agents Chemother., 55, 4012–4018.
109
7
Synthetic Biology and Therapies for Infectious Diseases
Hiroki Ando, Robert Citorik, Sara Cleto, Sebastien Lemire, Mark Mimee, and Timothy Lu
7.1
Current Challenges in the Treatment of Infectious Diseases
Following the remarkable success of antibiotics in the mid-twentieth century, US
Surgeon General William Stewart is said to have declared, “It is time to close
the book on infectious diseases, and declare the war against pestilence won” [1].
Since this declaration, however, antibiotic-resistant pathogens have progressively
emerged and grown in significance without concomitant improvements in antibiotic development. In addition to drug-resistant superbugs, we are facing newly
emerging and re-emerging infectious diseases. Thus, the book on infectious diseases remains open for the foreseeable future.
Of the 58.8 million annual deaths worldwide, 15 million are estimated to be
directly related to infectious diseases [2, 3] (Figure 7.1). There are many different
types of infectious diseases that afflict mankind. Many established infectious diseases have been prevalent for a long period of time and cause a relatively stable
level of morbidity and mortality. These include viral and bacterial respiratory and
diarrheal diseases, drug-susceptible tuberculosis (TB), tropical diseases such as
helminthic and parasitic diseases including drug-susceptible malaria, and nosocomial infections [3].
There are also newly emerging infectious diseases that are recently adapted to
the human host [3]. For example, in 2011, 34 million people lived with human
immunodeficiency virus (HIV), 2.5 million people were infected for the first time,
and 1.7 million people died from HIV-related illnesses [4]. In addition, some
infectious agents of nonhuman hosts can cross over to humans, but when these
are not passed from person to person, they only achieve dead-end transmission.
However, infections that are transmitted from animals to humans (zoonoses) and
those transmitted between vertebrates by the bite of infected arthropod vectors
(vector-borne diseases) have been identified. For example, rodent species and
viruses such as arenavirus and hantavirus have co-evolved [5]. Environmental
and human behaviors have increased human contact with these animals and
have given these viruses the opportunity to be transmitted from their natural
Novel Antimicrobial Agents and Strategies, First Edition.
Edited by David A. Phoenix, Frederick Harris and Sarah R. Dennison.
© 2015 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2015 by Wiley-VCH Verlag GmbH & Co. KGaA.
110
7 Synthetic Biology and Therapies for Infectious Diseases
Respiratory infections
4.3
2.2
2
Diarrheal diseases
HIV
1.5
Tuberculosis
0.9
Malaria
0.8
Childhood-cluster diseases
0.3
Meningitis
0.2
Tropical-cluster diseases
STDs excluding HIV
Hepatitis B
Hepatitis C
Dengue
0.1
0.1
0.05
0.02
Intestinal nematode
0.01
0.006
Leprosy
0.005
Japanese encephalitis
0.0
1.5
3.0
4.5
No. of deaths (millions)
Figure 7.1 Leading causes of death due
to infectious diseases [2, 3]. Childhoodcluster diseases include pertussis, diphtheria,
poliomyelitis, tetanus, and measles. Tropicalcluster diseases include schistosomiasis,
trypanosomiasis, Chagas disease, and
leishmaniasis. Sexually transmitted diseases
(STDs), excluding HIV, include syphilis,
chlamydia, and gonorrhea. Intestinal nematode infections include ascariasis and
trichuriasis.
hosts to humans. Another example of zoonosis is variant Creutzfeldt–Jakob
disease (vCJD). This disease is thought to be related to the prion associated with
bovine spongiform encephalopathy (BSE), which is commonly called mad cow
disease. Unlike almost all known prions, the new BSE prion easily infects humans
and a number of species, suggesting that the variants could cause unforeseeable transmission from animals to humans and be a serious threat to human
health [6, 7].
Indirect transmission, in which infectious agents are transferred to humans
or between humans, accounts for other zoonotic and non-zoonotic diseases.
Examples include enterohemorrhagic Escherichia coli O157:H7 and Campylobacter jejuni transmitted via food, milk, or direct contact, Legionella pneumophila
transmitted via artificial environments such as hot spring facilities and airconditioning systems, and Vibrio cholerae O139 and Cryptosporidium parvum
transmitted via the fecal–oral route [5]. In addition, infectious agents not only
cause acute health problems but can also result in chronic diseases [5]. For
example, hepatitis viruses type B and C, human herpesvirus 8, and Helicobacter
pylori can lead to liver cirrhosis and cancer, Kaposi sarcoma, and gastric ulcers
and cancer, respectively [8, 9].
Other infectious diseases have long histories of infecting humans but have
recently increased in prevalence in human populations, evolved, or changed their
7.1
Current Challenges in the Treatment of Infectious Diseases
epidemiology or drug susceptibility [3]. These include West Nile virus, influenza,
polio, dengue, drug-resistant TB and malaria, nosocomial infections, and cholera.
Malaria, caused by the protozoan parasite Plasmodium falciparum, is a major
infectious disease that continues to afflict mankind. The number of malaria cases
was reduced due to global malaria eradication efforts in the 1950s and 1960s [10].
However, the introduction of the insecticide dichlorodiphenyltrichloroethane
(DDT) led to the emergence of drug-resistant mosquitoes as well as concerns
about its potentially harmful effects on humans and nature. As DDT use has
decreased, malaria has resurged in prevalence and the emergence of resistance to anti-malarial drugs, such as chloroquine, has worsened the situation
[11].
The emergence and dissemination of drug-resistant bacterial pathogens
constitutes another major growing threat to public health [12]. For example,
methicillin-resistant Staphylococcus aureus (MRSA) is one of the most wellknown drug-resistant pathogens in the world. MRSA emerged several decades
ago and is still a major cause of healthcare-associated infections. For example,
the US Centers for Disease Control and Prevention (CDC) reported that MRSA
killed 11 285 people in the United States in 2011 [13]. The CDC also recently
classified three antibiotic-resistant pathogens at “urgent hazard levels”: Clostridium difficile, carbapenem-resistant Enterobacteriaceae (CRE), and drug-resistant
Neisseria gonorrhoeae [13]. These organisms are resistant to most commonly
used antibiotics. In particular, the treatment of CRE infections is extremely
difficult because these pathogens are resistant to almost all available antibiotics.
In addition, there are many other multidrug-resistant (MDR) and pandrugresistant (PDR) Gram-negative bacteria, including Acinetobacter baumannii,
Pseudomonas aeruginosa, and extended-spectrum β-lactamase-producing Enterobacteriaceae. These bacteria can exhibit broad resistance to antibiotics (e.g.,
β-lactams, aminoglycosides, tetracyclines, and polymyxins) [14].
MDR and extensively drug-resistant (XDR) Mycobacterium tuberculosis also
constitute a significant and growing class of bacterial infections, especially in
developing countries [15]. TB is responsible for causing 8.8 million incident
cases and 1.1 million deaths among HIV-negative patients, as well as 0.35 million
more deaths from TB associated with HIV in 2010 [16]. MDR-TB is defined
as resistance to both isoniazid and rifampicin, with or without resistance to
other first- and second-line drugs [17]. MDR-TB plus resistance to any fluoroquinolone and a second-line injectable drug is defined as XDR-TB [17]. Generally,
MDR-TB requires 2 years of treatment with potentially serious side effects while
the treatment of XDR-TB is even more difficult and often nearly impossible
[18].
As drug-resistant superbugs continue to grow in significance, there have been
few clinical approvals of new antibiotics to meet the challenge. From 1983 to 1992,
the U.S. FDA (Food and Drug Administration) approved 30 new antibiotics. In
the 10 years that followed, that number was reduced to 17 and, in the last decade,
only seven new antimicrobial drugs have been approved for marketing in the
United States [19, 20] (Figure 7.2). Over the last 40 years, only two mechanistically
111
7 Synthetic Biology and Therapies for Infectious Diseases
1983 − 1987
16
1988 − 1992
Years
112
14
1993 − 1997
10
1998 − 2002
7
2003 − 2007
5
2008 − 2012
2
0
6
12
18
No. of drugs
Figure 7.2 Antimicrobial drugs approved by the U.S. FDA for use in humans over the last
30 years. (Modified from Spellberg et al. [20] and Boucher et al. [19].)
and structurally new classes of antibiotics, linezolid and daptomycin, have been
introduced [21]. This lack of productivity has resulted in the continued growth
of life-threatening MDR, XDR, and PDR pathogens [22]. Although government,
academia, and industry are starting to recognize and address this issue, novel
approaches to discovering new antimicrobials are increasingly necessary. In 2012,
the European Union’s Innovative Medicines Initiative launched a public–private
partnership called NewDrugs4BadBugs, which was aimed at bringing together
academia and industry to tackle antibiotic resistance. In addition, the United
States also recently enacted new incentives to encourage drugmakers to develop
new therapies against drug-resistant bacteria [19].
7.2
Introduction to Synthetic Biology
We believe that synthetic biology presents new opportunities to combat infectious diseases. The field of synthetic biology is focused on engineering biological
systems to achieve new functionalities using genetic engineering strategies
that are increasingly more robust, rapid, inexpensive, and multiplexable. Synthetic biology has been accelerated by the development of next-generation
DNA sequencing, synthesis, and assembly techniques, thus enabling rapid and
parallelized construction and modification of biological systems [23]. These
tools enable the design of biomolecular and cellular systems that can address
biomedical applications, including infectious diseases [24–26]. We refer the
interested reader to other reviews for more generalized descriptions of synthetic
biology [23, 27, 28]. Here, we highlight existing approaches to combat infectious
diseases with a primary focus on bacteria. In addition, we discuss advances in the
application of synthetic biology to other infectious diseases and the potential of
synthetic biology to work synergistically with other fields [26] (Figure 7.3).
7.3
Synthetic vaccine strains
with effective antigens,
attenuated virulence,
and delayed lysis
ATGC
TACG
Pyocin
Bacteriocin
Synthetic probiotics
Synthetic bacteriophages
with broad host ranges,
biofilm-dispersing enzymes,
and antibiotic-boosting activity
B
AB
Synthetic biology
In silico mining and synthetic
pathway engineering for
natural products
ATGC
TACG
T7
dspB
113
Synthetic antibiotics
Reverse vaccinology
With desired functions
by in silico analysis
and individual antigen screening
Lysin
P
A
Vaccinology
Next-generation genetics
Gibson assembly, yeast platforms,
CRISPR/Cas-based RGNs
M13
lexA3
Figure 7.3 Schematic overview of synthetic biology approaches to combat infectious
diseases.
7.3
Vaccinology
Vaccination is probably the most successful anti-infection strategy devised by
modern medicine. Interestingly, the majority of vaccines have been developed
empirically with little knowledge regarding how and why they confer protection.
Instead of allowing infections to establish themselves before initiating treatment,
vaccines are given prophylactically to harness the natural capacity of the body
to defend itself. Vaccines have proved effective against many viral and bacterial
diseases such as smallpox, polio, measles, mumps, rubella, yellow fever, rabies,
tetanus, diphtheria, pertussis, and bacterial meningitis. However, vaccination
has also met significant failures, for example, against pathogens such as HIV,
Salmonella, V. cholerae, E. coli, and P. falciparum [29–36]. The scope of the
following chapters is mostly limited to the recent developments of vaccines
against bacterial infections, especially when antibiotic resistance is already or is
becoming a significant issue for treatment.
Vaccinology relies on the establishment of immunological memory of a foreign
agent such that the immune system can elicit a more potent, protective response
against subsequent exposure. The mechanisms underlying vaccinology are well
described in the recent review by Pulendran and Ahmed [37] and we refer the
reader to this work for more in-depth understanding. Briefly, the innate immune
system constitutes the first line of defense, wherein antigen-presenting cells,
114
7 Synthetic Biology and Therapies for Infectious Diseases
such as dendritic cells and macrophages, sense pathogens through various signal
transduction pathways. This information is transmitted to the effectors of the
adaptive immune system (T and B cells), which mount a specific response to
the threat in the form of dedicated killer cells or antibody-producing cells. In
addition, the response can generate memory cells that archive this information in
order to ready the body for a more rapid and robust response in future challenges
with the same pathogen. A successful vaccine can trigger all levels of the immune
response, including memory generation, in order to obtain effective long-term
protection.
Some vaccines are killed or inactivated pathogens that trigger a protective
reaction against a related microbe. In addition, live attenuated vaccines are
derived from actual pathogens that have accumulated mutations that weaken
their disease-causing capabilities while leaving important antigens and immunogenicity intact. Examples of such vaccines include bacille Calmette-Guérin
(BCG) against M. tuberculosis, Ty21a against Salmonella typhi, and MMR against
measles, mumps, and rubella. Inactivated/killed vaccines, such as specific polio
or hepatitis A vaccines, consist of the whole pathogen killed, usually through
heating. It was also discovered that an immune response could be obtained
through exposure to only parts of the pathogen and that this may also decrease
side effects from exposure to the whole pathogen. Those parts may be toxins
or part of toxins; such vaccines are referred to as toxoid vaccines and examples
include vaccines against diphtheria and tetanus. Components of the pathogen
surface, such as the capsule, lipopolysaccharides (LPSs), or outer membrane
proteins, may also be used in subunit vaccines; examples include hepatitis B,
pneumococcal, and meningococcal vaccines [38]. Live attenuated vaccines are
usually longer acting than subunit vaccines, which generally require booster
injections at regular intervals to maintain efficacy [39].
7.3.1
Genetic Engineering and Vaccine Development
The first live, attenuated vaccines, such as BCG or S. typhi Ty21a, were developed
by empirical selection of attenuated strains that confer protective immunity
against otherwise lethal challenges by their respective associated pathogen,
Mycobacterium tuberculosis (Mtb), responsible for TB, and Salmonella typhi
(Sty), responsible for typhoid fever. Both vaccines display incomplete protection,
and therefore, intensive efforts with genetic engineering have been employed to
improve them.
Drug-resistant TB has increased steadily in recent years to the point where
strains refractory to all first-line antibiotics (isoniazid, rifampicin, ethambutol,
and pyrazinamide) now appear in 10–20% of all TB patients [16]. TB is a chronic
infection involving a constant battle between bacteria and the human immune system. After initial infection, acute primary TB develops in 5–10% of patients and an
estimated 20–50% of individuals manifest as asymptomatic carriers [40]. Immunodeficiency, such as in AIDS patients, leads not only to enhanced susceptibility,
7.3
Vaccinology
morbidity, and mortality to the disease but also to increased reactivation in latently
infected individuals, making TB one of the most frequent and lethal of the AIDSassociated co-infections [16].
BCG is an attenuated mutant of Mycobacterium bovis (Mb), the causative
agent of bovine TB, and was obtained by serial passage of the bacilli in laboratory
medium [41]. BCG was first commercialized in the 1920s, but its efficacy has
since been called into question. First, BCG does not provide universal protection
against all pathogenic strains of Mtb and the extent of immunity varies between
individuals. Second, when there is an immune response in the host, the degree of
protection it offers against pulmonary TB, the most dangerous and contagious
form of the disease, is variable and ranges from 77% in some studies to 0% in
others [40]. Finally, it is protective only if administered during early childhood
[42].
As a result of serial passaging, the various strains of BCG have all lost the
RD1 locus, which encodes the majority of Mb’s and Mtb’s virulence factors,
thus resulting in attenuation [42]. Mycobacteria are intracellular pathogens
that reside in heavily modified phagolysosomes, which may restrict immune
access to antigens. Thus, toward the goal of improving the vaccine strain, it was
hypothesized that liberating Mycobacterium into the cytosol could increase the
antigenicity of BCG. A derivative of BCG (VPM1002) was genetically engineered
to carry a replacement of ureC, a gene involved in pH neutralization inside the
Mycobacterium-containing phagolysosome, with the gene encoding listeriolysin
O (LLO), a virulence factor from Listeria monocytogenes that allows it to escape
into the cytosol of infected cells. VPM1002 proved significantly more immunogenic and equivalently safe versus BCG in a randomized Phase 1 trial reported
recently by Grode and colleagues [43] and is now undergoing Phase 2a trials.
Another group led by Marcus Horwitz at the University of California, Los
Angeles, and the company Aeras adopted a different approach by genetically
engineering BCG to express Ag85B (encoded by fbpB), a known antigen of Mtb
absent from Mb. The resultant strain, rBCG30, exhibited increased immunogenicity and also cleared Phase 1 trials, thus demonstrating the power of rational
genome engineering for creating live, attenuated vaccines [44–46]. rBCG30 was
further modified by the deletion of mbtB, a siderophore transport protein, as
well as panCD, which made the strain dependent on high levels of free iron and
vitamin B5, respectively. These engineering steps resulted in an even weaker
pathogen that replicates slowly and has potentially enhanced safety over BCG
in immunocompromised patients [47]. Finally, most of these features were
integrated into the recombinant vaccine strain AERAS-422. AERAS-422 has
the ureC gene replaced by the Clostridium perfringens perforin gene to enable
phagolysosome escape and also overexpresses three key Mtb antigens [48].
The examples provided above illustrate the usefulness of genetic engineering in
live vaccine design, which has led to the development of promising new vaccine
strains. In addition to TB, strain engineering strategies have been applied to other
important diseases. For example, typhoid fever is still endemic in many parts of
the world and causes severe diarrhea and fever that can be lethal in 20% of infected
115
116
7 Synthetic Biology and Therapies for Infectious Diseases
individuals if proper medical treatment is not administered. As with most other
bacterial diseases, multidrug resistance is on the rise and no effective vaccines are
currently available [49]. S. typhi is capable of residing for decades in the gall bladders of infected individuals, resulting in a carrier state that enables shedding and
dissemination of the pathogen from asymptomatic individuals [50]. The first live,
attenuated vaccine available for Salmonella is known as Salmonella typhi Ty21a,
although it does not protect against all typhi strains. Furthermore, longer term
protection requires frequent booster shots and is not indicated in children below
6 years of age, where the toll of the disease is most severe [51].
Salmonella is a facultative intracellular pathogen that has evolved a complex
and tightly orchestrated interaction with infected cells. At the heart of its pathogenicity lies two Type 3 Secretion Systems (T3SSs) encoded by Salmonella
Pathogenicity Islands 1 and 2 (SPI-1 and SPI-2, respectively), which allow the
bacterium to inject a diverse set of proteins and toxins into the cytoplasm of
infected cells [52, 53]. These secreted factors enable the bacteria to remodel
the host cell cytoskeleton for uptake, prevent maturation of the subsequent
phagolysosome, and finally recruit membrane vesicles to form highly specific
Salmonella-containing vacuoles where the bacteria can survive and multiply.
SPI-1 is particularly involved in initial epithelial cell invasion, while SPI-2 is
connected to systemic infection and survival inside of macrophages [54, 55]. The
understanding of this pathogen’s biology along with its genetic tractability has
enabled some of the most refined live, attenuated vaccines designed to date. Various auxotrophic mutations result in attenuation, but the most commonly studied
are 𝛥fur, 𝛥phoP/Q, 𝛥aroC/D, 𝛥rpoS and the double mutant 𝛥cya 𝛥crp [56–59].
Fur (Ferric Uptake Regulator) is a regulator involved in iron homeostasis and
acid tolerance, a requirement for surviving passage through the stomach or the
lysosomes of infected macrophages [60, 61]. The genes phoP/Q encode a sensor
kinase system involved in the sensing of external cues and proper expression of
many virulence genes of Salmonella [62]. The aro genes encode proteins that are
involved in the metabolism of aromatic compounds [56]. Mutations in the aroA,
aroC, and aroD genes also alter membrane integrity, thus increasing sensitivity to
various membrane stresses [63]. RpoS is a sigma factor involved in the response
to stresses and/or regulation of virulence pathways in many bacterial pathogens
[64]. Cya is an adenylate cyclase responsible for the synthesis of cAMP [56]. A
key step in the use of Salmonella as a live vaccine was the development of a
series of plasmids for the expression of heterologous cargo that do not require
antibiotics for selection. In these constructs, stability is ensured through the
complementation of a deleted essential Salmonella gene, such that plasmid loss is
lethal to the bacterium [65, 66]. The genetic cargo carried on these plasmids can
be used either to alter the life cycle of Salmonella, express heterologous antigens,
or a combination of both, as illustrated in the following paragraphs.
As knowledge about Salmonella’s interactions with eukaryotic cells grew, it
became evident that the initial attenuated Salmonella strains described above
were severely limited by pleiotropic effects linked with the mutations that
weakened them and that part of the reduced virulence phenotypes were the
7.3
Vaccinology
result of defective host colonization, which prevents the activation of a robust
immunologic reaction [67]. To further leverage the usefulness of Salmonella
as a vaccine agent, strains were engineered to have essential genes under the
control of inducible promoters. These strains were viable in the laboratory,
but died through lysis after a programmable number of replication cycles due
to exhaustion of an essential bacterial component [65, 66]. The ensuing lysis
enabled the delivery of antigenic cargo directly inside of antigen-presenting cells,
thus enhancing immunogenicity. To further avoid the metabolic load associated
with high-level synthesis of an antigen, the most recent Salmonella vaccine
strains feature delayed overexpression of antigens. The result of this work is a
set of strains collectively known as recombinant attenuated Salmonella vaccines
(RASV s), which incorporate a variable set of genetically engineered solutions to
the known limitations of Ty21a and transform Salmonella into an efficient antigen
delivery system [68]. The various genetic modules, which have been engineered
into RASV strains, are depicted in Figure 7.4. Several aspects regarding the
development of these strains are discussed in the following paragraphs but the
reader is referred to recent publications for a more complete overview [68–70].
Several Salmonella strains that penetrate the host at different depths are available, thus triggering more or less pronounced immune responses to the release of
antigens. Strains completely deleted for phoPQ or rpoS fail to replicate past their
initial port of entry, the intestinal epithelium [71, 72], while a fur deletion mutant
is attenuated because it is incapable of mounting a proper acid tolerance response
to the acidic environment of the stomach [60, 73]. Researchers have replaced the
endogenous promoter for each of these genes with the arabinose-inducible promoter, PBAD . Thus, the expression level of these genes can be maintained in the
laboratory with the addition of arabinose, but will gradually fall over the course
of replication during infection as the human body does not contain arabinose.
This results in delayed attenuation, which allows bacteria to reach deeper into
the body than simple deletion strains. The depth of penetration is at least partially dependent on the number of modified regulatory genes introduced into the
bacterium [74–76]. Delayed antigen overexpression was obtained through highlevel expression of the transcriptional repressors C2 and LacI from the arabinose
promoter, which then downregulate the expression of antigens cloned under the
control of their respective promoters, P22 PR or Plac [77–79]. With the abovementioned PBAD driven genes, C2 or LacI synthesis stops shortly after inoculation
due to absence of arabinose, and the dilution of these repressors by division and
decay gradually leads to antigen expression.
A refinement of this method involved using promoters associated with the
expression of late infection functions to drive the synthesis of antigens, thus
ensuring that antigens are only produced when Salmonella receives specific environmental cues upon reaching its target site inside of macrophages [78]. Another
strategy for delayed antigen liberation is achieved through the replacement of
the murA promoter by the arabinose promoter. MurA is essential for cell wall
synthesis and therefore, its gradual disappearance within arabinose-free host
tissues eventually leads to cell wall collapse and lysis [80].
117
118
7 Synthetic Biology and Therapies for Infectious Diseases
Delayed lysis
Delayed antigen delivery
ΔasdA ΔmurA PBAD-asdA PBAD-murA PR(anti asdA) PR-(anti murA) PBAD-c2
SopE(N-ter)-antigen
OmpC(N-ter)-antigen
ΔsifA
Delayed virulence
attenuation
PBAD-lacI Plac-antigen
PrfaH::PBAD
RASV
Pfur::PBAD
Delayed antigen
overexpression
Pcrp::PBAD
PphoPQ::PBAD
PrpoS::PBAD
Arabinose promoter behaviour
RASV
Growth Arabinose PBAD
condition
In vitro
+
In vivo
-
Figure 7.4 Recombinant attenuated
Salmonella vaccine (RASV) components.
RASV are highly engineered Salmonella
strains that may contain combinations of
modifications to attenuate virulence, maximize immunogenicity, or deliver antigens.
Those components may be chromosomally or plasmid encoded. This figure is not
an exhaustive representation of all possible
mutations that could create RASV strains.
The arabinose promoter used in many of
the strains is regulated by AraC, a chromosomally encoded, arabinose-responsive
transcriptional regulator that activates the
PBAD promoter. As arabinose is not readily
available in eukaryotic cells, PBAD remains
inactive during infection. LacI is a transcriptional repressor of the Plac promoter. If
lacI is controlled by PBAD , any gene downstream of a Plac promoter will be upregulated during intracellular growth of RASV.
Similarly, C2 is a repressor of the phage
P22 promoter PR . rfaH, a transcriptional
PsopE-antigen
Plac-antigen PBAD-lacI
PL-antigen PBAD-c2
anti-terminator essential to the synthesis of
lipopolysaccharide (LPS), is an essential component of Salmonella virulence. The genes
fur, crp, and rpoS encode general regulators
of gene expression, while phoPQ is a twocomponent system that is essential for the
proper expression of many virulence factors. SopE, a Salmonella Pathogenicity Island
1 (SPI-1) secreted effector, possesses an Nterminal domain that is sufficient for targeting proteins to the SPI-1 type III secretion
system. OmpC is an outer membrane protein with an N-terminal export signal. SifA
is a SPI-2 virulence effector involved in the
maturation of Salmonella-containing vacuoles
in which Salmonella resides during intracellular growth. Its absence leads to release of
bacteria into the cytosol. AsdA and MurA are
involved in cell-wall synthesis. As bacteria
replicate, new cell walls need to be synthesized to accommodate growth; thus, a failure
to synthesize cell walls leads to cell lysis.
Engineered Salmonella vaccines have also proved to be useful immunogenic
vectors in various animal models and against a range of non-Salmonella threats
such as enterohemorrhagic E. coli [81], hepatitis B [82], TB [79], cancer [83–85],
Bacillus anthracis [86], H. pylori [87, 88], Yersinia pestis [89, 90], Clostridium
tetani and difficile [91–94], and L. monocytogenes [95]. For example, Salmonella
possesses the capacity to secrete proteins directly from its cytoplasm into the
7.3
Vaccinology
eukaryotic cytoplasm through dedicated channels encoded by the SPI-1 and
SPI-2 T3SSs. Secretion is generally directed by small N-terminal domains,
which can be fused to target proteins and guide them for secretion through the
T3SS of Salmonella [96–98]. This capacity makes it possible for Salmonella
to deliver antigens from within the vacuole it resides in into the cytoplasm of
infected macrophages, a prime target for Salmonella replication during systemic
infection. These macrophages can then display these cytoplasmic antigens on
the macrophage surface for priming T cells and eliciting immune responses
[99]. This strategy was used in the fusion of the 80-amino-acid N-terminal
fragment of SopE, an SPI-1 secreted effector of Salmonella, with the Mtb-specific
antigens ESAT-6 and CFP-10; this design provided protection against infection
by aerosolized Mtb in mice, a model system for TB [100].
Despite all of the promises described above, RASVs are still not used in clinical medicine partially due to concerns that Salmonella is a pathogen. Laboratories
have therefore worked on reproducing RASV features in organisms such as Lactococcus lactis or Bifidobacterium that are generally recognized as safe (GRAS). For
example, Asensi and colleagues [101] successfully cloned staphylococcal enterotoxin B into L. lactis and triggered protective immunity against staphylococcal
challenge in mice. In addition, bifidobacterium was also used to express antigens
of E. coli or Salmonella and trigger at least some level of immune response against
the cognate pathogen [102, 103]. Lactic acid bacteria have been also used as probiotics with reported benefits against inflammatory diseases, obesity, and allergies
[104–107], thus paving the way toward engineered strains functioning at the intersection of vaccines and probiotics.
In any of these live, attenuated vaccine strains, expressing a foreign antigen may
be hampered by low transcriptional efficiency, translational efficiency, mRNA stability, or protein stability, leading to poor immunogenicity. Attempts to address
these problems have leveraged strategies such as codon optimization and mRNA
analysis and design [108, 109]. As synthetic biologists develop systematic strategies to tackle these issues for general organismal engineering [110–112], more
tools will be available for strain engineering and vaccine design in the future. Furthermore, enhanced technologies for genome engineering should enable more
rapid and precise genetic modifications of vaccine strains [113–115].
7.3.2
Rational Antigen Design Through Reverse Vaccinology
Traditional subunit vaccinology starts from a pathogen and dissects it into progressively smaller pieces that are each independently tested for immunogenicity.
This process can identify individual subunits that provide robust and broad protection against the whole pathogen. This long, costly, and unpredictable process
can be challenging to apply against pathogens that evolve more rapidly than our
ability to update vaccines.
119
120
7 Synthetic Biology and Therapies for Infectious Diseases
Reverse vaccinology describes an antigen-design strategy that makes heavy use
of database mining, functional prediction algorithms, and three-dimensional protein structure prediction software to identify antigen candidates that can be tested
experimentally for effectiveness [116]. The tools of synthetic biology can further
advance the field of reverse vaccinology by enabling more rapid construction,
expression, and testing of individual antigens. Reverse vaccinology starts from
identifying open reading frames (ORFs) from the genome sequence of a pathogen.
The predicted proteins are then analyzed for their potential involvement in virulence or their possible display on the bacterial surface. In addition, one attempts to
identify pathways that may lead to LPS and murein synthesis in order to try to predict the resulting structures, as both components may be immunogenic. Further
biological analysis is limited to those candidates identified by these approaches
that usually encompass only a tiny subset of the original ORFs. These ORFs are
then expressed in heterologous hosts and individually tested for immunogenicity.
Although screening is still necessary in reverse vaccinology, the number of possible candidate antigens is drastically reduced by computational pre-screening,
which eliminates genes whose products are unlikely to be immunogenic, because
of reasons such as retention inside the cytoplasm of bacteria, and/or not specific
to the pathogen. Reverse vaccinology also allows for the identification of antigens
Traditional vaccinology
Pathogen
Reverse vaccinology
Fractionation of
cell components
DNA extraction
and sequencing
ATGC…
In silico analysis of ORFs for
candidate antigens (surface
exposed cell components)
Subcloning of antigen
candidate genes
Independent testing of each fraction for immunogenicity
etc…
Expression and testing of
individual antigen candidates
Isolate
Iterative subfractionation and testing
protective
antigens
Isolate protective antigens
Safety and efficacy testing in clinical trials
(a)
(b)
Figure 7.5 Comparison of traditional and
reverse vaccinology methodologies. Panel
(a) presents the general workflow for antigen discovery in a traditional vaccinology
study while the panel (b) illustrates the
workflow in a reverse vaccinology process,
in which the search for antigens starts from
the pathogen’s DNA instead of the pathogen
itself.
7.3
Vaccinology
that may be missed in traditional vaccinology owing to low expression levels from
the original pathogen. A comparative schematic of traditional and reverse vaccinology is presented in Figure 7.5.
One early successful implementation of this technique was for group B
meningococcus (MenB). Meningococci are natural residents of human airways,
but occasionally move into the bloodstream and can cross the blood–brain
barrier to cause meningitis, a highly lethal and contagious disease [117]. Some
vaccines existed, but they did not cover MenB, the most prevalent serogroup
in the Americas and Europe [118], until the introduction of Bexsero® whose
construction is described in the following paragraphs. The most important
virulence factor of Neisseria meningitidis is a polysaccharide capsule that shields
it from the immune response and adverse conditions [117]. The MenB capsule
is made of polysialic acid, which closely resembles human matrix components
and is therefore very poorly immunogenic [119]. Moreover, due to the extreme
variability of the capsule, early subunit vaccines based on capsule components
had a very restricted spectrum (serogroups A, C, Y, and W135). In addition,
variability in the displayed outer membrane proteins from strain to strain within
a given serogroup presents a significant challenge to the development of a MenB
vaccine [120].
Starting from the genome of N. meningitidis strain MC58, about 300 ORFs
were predicted to be either virulence factors or externally displayed proteins.
These were cloned into E. coli, overexpressed, and used to immunize mice [119].
Out of those tested, 91 ORFs were found to elicit an immune response, but
only 28 generated a protective response. Out of those 28, only 3 were found
to have a robust immunogenicity and a reasonably wide coverage spectrum:
Genome-derived Neisseria antigens (GNAs) 2132 (Neisserial Heparin Binding
Antigen, NHBA), 1870 (fHbp), and 1994 (NadA). Later, GNA2091 and GNA1030
were included because, although they did not meet all selection criteria, they still
offered good protection. To ease production of the subunit antigens, GNA1030
and GNA2132 were fused together, while GNA1870 and GNA2091 were also
fused into a single polypeptide. The resulting multicomponent vaccine, named
4CMenB or Bexsero, contained these antigens and outer membrane vesicles
and provided protection against 78% of strains from a large panel of different
N. meningitidis group B strains [119]. Reverse vaccinology was also successfully
applied to other pathogens, including Brucella melitensis [121], an important
animal pathogen, E. coli [122], and various Streptococcus strains [123–125].
Reverse vaccinology also involves the use of structure-based approaches for
antigen design. Here, knowledge and predictions regarding the three-dimensional
conformation of an antigen are used to remodel it for better antigenicity. One
of the antigens used in the 4CMenB vaccine, GNA1870, is known to have more
than 300 variants across Neisseria strains, which fall into only three structural subgroups [126]. Nevertheless, antibodies against one subgroup are not active against
the others. The 3D structures of all three antigen variants were determined by
nuclear magnetic resonance (NMR), leading to the identification of the epitope
structures recognized by antibodies against each individual subtype. Protective
121
122
7 Synthetic Biology and Therapies for Infectious Diseases
monoclonal antibodies against each epitope type targeted different regions of the
protein [119, 127]. Thus, it was conceivable to create a synthetic GNA1870 that
would display all three epitopes, which was expected to have broader immunogenicity. Although challenging, such a protein was eventually developed and was
shown to provide broader spectrum protection [127]. The process involved creating a large bank of GNA1870 hybrids based on variant 1 but harboring various fractions of the GNA1870 C-terminal domain from variant 2 and/or 3. The
screen was limited to the C-terminal domain because it was shown that most of
the protective antibodies against variants 2 and 3 bound within this region. Each
hybrid was assessed for its capacity to elicit production of cross-reacting antibodies against all three classes of GNA1870 after injection into mice and also to protect mice against meningococcal infection [127]. This approach is useful for tackling diseases caused by pathogens with highly variable genomes and is currently
being used to create a vaccine against group B Streptococcus [128]. Structural vaccinology can also be used to create subunit vaccines that have been depleted of
domains that mask strong epitopes or trigger their instability or rapid clearance.
For example, GNA1870 was shown to be more active when its factor H binding
domain was deleted [129], thus exemplifying the power of modern tools in the
fight against pathogenic bacteria.
Next-generation sequencing and high-throughput expression analysis technologies, such as microarrays or RNA-sequencing (RNA-seq), are expected to
play a growing role in reverse vaccinology. These technologies can accelerate the
characterization of the diversity of pathogen populations and gene expression.
Detailed datasets can help refine predictive algorithms and be used to improve
the genome annotations on which reverse vaccinology relies, thus increasing
the precision of antigen predictions. As understanding of the systems biology of
pathogen/host relationships matures, it is also likely to help in the identification
of pathways of importance for effective vaccination.
7.4
Bacteriophages: A Re-emerging Solution?
7.4.1
A Brief History of Bacteriophages
Viruses that infect bacteria, or bacteriophages, were discovered around the time
of World War I for their capacity to lyse cultures of dysentery-causing bacilli. The
stools of patients surviving dysentery were found to contain filterable substances
(now known to be viruses) that formed areas of growth inhibition on lawns of the
disease-causing agent and could lyse the bacteria in culture [130]. When administered to patients, these bacteriophages were able to ameliorate disease symptoms, although the results were unpredictable and were not obtained via rigorous
double-blinded studies [131].
7.4
Bacteriophages: A Re-emerging Solution?
The discovery of phages and their viral nature was received with a high level of
scrutiny that was only resolved when electron microscopes became available and
the viral particles could be visualized [132]. The chaotic history of bacteriophages
delayed their initial general acceptance as antimicrobials and they were further
disregarded in much of the Western world when penicillin became routinely available around the 1950s [28, 133]. However, with the growth of antibiotic resistance
in recent years, interest in phage therapy has rebounded, especially with advances
enabling increasingly more efficient genetic engineering of bacteriophages.
Although not prominent in the West, phage therapy was practiced in the United
States until the 1960s, in France until the early 1990s, and in other countries as well
[134]. Furthermore, phage therapy was used in the Soviet bloc as an alternative to
antibiotics. Two approaches to phage therapy differ on the basis of the diversity
of the pathogenic populations being targeted. The more generalized, symptombased approach is to create complex phage cocktails targeted at many different
pathogens that share commonality in the presentation or location of infection.
For example, Pyophage is a formulation aimed at treating open wounds while
Intestiphage contains a different collection of phage for targeting common causes
of gastrointestinal infections [134]. In both formulations, the actual bacteriophage
composition is updated on a regular basis to cope with the evolution of bacterial
populations and is therefore not statically defined. Alternatively, a more targetspecific approach to phage therapy is to use phages as a last resort when antibiotics
have failed and when phage(s) specific to the disease-causing bacteria can be isolated from a phage library and administered to the patient. In this approach, phage
therapy is a highly personalized medical therapy. We will not review the abundant
literature on phage therapy efficacy in this paper and instead refer the interested
reader to other detailed reviews on the subject [28, 134–141].
Phage therapy presents several potential advantages compared to antibiotics
[28]. First, a given phage often infects only a subset of strains for a given pathogen.
This specificity means that phages are innocuous to the resident commensal
microflora and do not trigger the dysbiosis that can result from harsh antibiotic
treatments [142, 143]. Phages also have very low inherent toxicity as they
cannot infect human cells, and humans are constantly exposed to them in the
environment. Phages can replicate in their bacterial hosts, which means they
can have efficacy at very low doses, although this property can lead to complex
pharmacokinetics/pharmacodynamics (PK/PD) that are quite different from that
of small molecules. A few important principles for effective phage therapy have
been outlined [144]: (i) the phage and the bacterium should be in direct contact,
which makes phage therapy attractive as a topical treatment; (ii) active phage
preparations that target the infectious bacteria should be assembled and used in
a timely manner; and (iii) a regulatory framework to evaluate safety and efficacy
should be established.
Phage therapy must address several areas of concern to become a practical clinical therapy [28]. One of the major obstacles to the Western adoption of phage
therapeutics is the regulatory process. Although the Soviet states accumulated a
wealth of knowledge about phage therapy and there is evidence about its efficacy
123
124
7 Synthetic Biology and Therapies for Infectious Diseases
in specific cases, the vast majority of this data is not freely available and may
be archived in languages and formats not easily interpretable by the rest of the
world. In addition, most existing safety and efficacy studies were not conducted in
well-controlled clinical trials. Moreover, the sudden lysis of infective bacteria due
to phage administration has the potential to cause an excessive release of endotoxins and other toxins that can result in septic and/or toxic shock. Phages can
be immunogenic, which can lead to rapid clearance from the body and establishment of immunological memory, making it potentially ineffective to use the
same phage multiple times in a given patient in case of relapse or re-infection. In
addition, it is challenging to patent non-engineered phages, since they are natural biological particles that can be found in the environment. Furthermore, the
highly specific nature of phages means that one must ensure that the causative
bacteria are susceptible to the phages applied and one must also account for the
evolution of phage-resistant bacteria over time. These difficulties, along with the
PK/PD of phage therapy, may account for a large extent of the lack of interest of
large pharmaceutical companies and investors in phage therapy.
Phages have evolved to prey on bacteria and, in doing so, have adopted countless ways to subvert their hosts and kill them. No single natural phage has all the
desirable features for phage therapy, but synthetic biology is potentially capable of
addressing many of their shortcomings through phage engineering as described
in the following sections. Furthermore, harvesting the killing potential of phages
independent of their viral lifestyle is a way to circumvent many of the regulatory
hurdles linked with phage therapy per se and can lead to either new antibacterial
protein therapeutics and/or new targets for the design of antimicrobials.
7.4.2
Addressing the Problem of the Restricted Host Range of Phages
Most natural phages have restricted host ranges (Figure 7.6a). One way to address
this problem relies on phage cocktails containing several different phages that collectively target all possible variants of a pathogen. However, this approach is challenging for the regulatory framework employed in Western medicine. Although
the U.S. FDA does not disallow the possibility of licensing cocktails, it requires
that each individual phage be tested independently for safety and activity [145].
For infection to proceed, phages first need to recognize and adsorb to specific
receptors on host cell surfaces. These can include proteins, LPS components, capsule components, pili, other molecules, or any combination of these factors [149].
Once adsorption to the receptor is stable, the phage may inject its DNA (and sometimes proteins) into the bacterium where its life cycle will initiate. Much of the
host specificity of a phage is determined by the presence or absence of its specific
receptor(s) on the surface of target cells and thus, mutations in phage receptors
are often sufficient to confer resistance [150]. This mechanism of phage resistance
is the most common under laboratory conditions, with frequencies ranging from
10−5 to 10−8 depending on the particular host/phage considered [151, 152].
7.4
Bacteriophages: A Re-emerging Solution?
Flipping site
Tail A
Tail B
A
B
Tail A
Tail B
Tail A
Tail B
Flipping
C
Tail A
A
B
C
C
Bacteria
A
Bacteria
B
Bacteria
A
Bacteria
B
(a)
Tail B
D
D
D
D
Bacteria
A
Bacteria
B
(b)
(c)
Figure 7.6 Bacteriophage host range and
strategies used to enhance access to host
receptors. (a) Phage A carrying tail A infects
bacteria A, and phage B carrying tail B
infects bacteria B. Phage A cannot infect bacteria B, and phage B cannot infect bacteria
A. Bacterial host range can be dictated by
available bacterial receptors that can be recognized by phages [146]. (b) Phage C carries
two separate tail genes, tail A and B, with
different hydrolytic properties that allow
for drilling through both types of capsules
[147]. (c) Phage D has an invertible segment
that allows switching their host ranges. In
contrast with phage C, phage D does not
express the two versions of the tail proteins
at the same time. The flipping is mediated
by a site-specific recombinase [148].
Receptor recognition by tailed phages usually results from proteins from the
tail itself or associated structures, such as the distal tail fiber, the tail fiber, or
the tail spike [149]. In the most simple adsorption systems, the adhesin at the tip
of a tail fiber will recognize a bacterial outer membrane protein or other bacterial outer membrane components, such as LPS, and then reversibly anchor the
phage to its target [149]. Many phages may require several receptors for efficient
adsorption [146]. Some phages also have the capacity to degrade outer membrane
appendages, such as LPS or capsules, in order to access secondary receptors that
are hidden by these surface components [147, 153, 154]. Phages that are not tailed
still require a specific host receptor on the outer membrane to initiate nucleic acid
injection [155, 156]. For example, for filamentous phages such as Fd, host recognition proceeds through the product of gene pIII, which is attached at only one
end of the virion [155]. Fd’s preferred host is F-pilus-expressing E. coli; however,
phage adsorption to E. coli may still occur in the absence of F although with much
reduced efficiency [157]. Fd can be engineered to infect V. cholerae upon replacement of its pIII gene with its ortholog from vibriophage CTXϕ, orfU [157].
125
126
7 Synthetic Biology and Therapies for Infectious Diseases
While most phages have a single adhesin, some have several recognition elements they can use to widen their host range or to target other hosts when the
supply of a specific host is exhausted (Figure 7.6b and c). The switch can be actuated by a DNA inversion system that inverts a DNA fragment carrying the adhesin
domain and replaces it with another one [158, 159]. In addition, some Bordetella
phages have evolved a complex diversity-generating system to track the fast evolution of the surface of their hosts [160]. One can envision leveraging these natural
strategies and engineering them into synthetic phages to gain more precise control
over phage host ranges.
In addition, K1-5 is a strictly lytic T7-like coliphage that preys on strains harboring either a K1 or a K5 capsule. As these two types of capsule are structurally
different, they cannot be targeted by a single factor. Thus, K1-5 phage carries two
separate tail spikes with different hydrolytic properties to allow drilling through
both types of capsules (Figure 7.6b). The endosialidase-type tail spike (gp47)
enables binding and degradation of the K1 capsule while the lyase-type tail spike
(gp46) enables binding and degradation of the K5 capsule. Inactivation of any one
of these tail spikes restricts K1-5 to a single type of encapsulated E. coli [147].
Both tail spikes are assembled into the phage capsids, side by side [161]. K1-5 is a
very close relative of phage K1E, which lost most of gp46 and therefore targets K1
strains, and phage K5, which only has gp46 and thus only infects K5 strains [162].
Lysogenic coliphages Mu and P1 both have an invertible segment that allows
them to switch their host ranges at a predetermined frequency (Figure 7.6c). In
sharp contrast with K1-5, they do not express the two versions of their tail-fiber
adhesins at the same time, but can adapt to another niche if one specific E. coli
host is depleted or changes its expressed receptor. The switch is operated by a
site-specific recombinase that inverts the C-terminal fragment of the tail fiber
gene while leaving the N-terminus intact, which is required for binding of the
side tail fiber to the tail [158, 163]. A close relative of P1, the defective prophage
p15B, harbors an even more complex multiple inversion system that can create
240 different variants of the tail fiber gene through reassortment of the 5′ end
of the gene. Interestingly, a tail-fiber-deficient P1 phage can be complemented
by the p15B Min locus, which indicates that the Min locus encodes a functional
tail fiber and also that it is capable of binding to the tail-fiber-less capsids of the
above-mentioned P1 mutant [148].
T4, a strictly lytic coliphage of the T-even family, has its host specificity encoded
within the N-terminal segment of gene gp37, which forms the distal portion of its
side tail fibers and binds to receptors on the host surface, the OmpC porin [164].
In most other members of the T-even family, such as T6 or T2, the adhesin is separated from the distal tail fiber gene and is encoded right downstream of it in gene
gp38 [146]. Such adhesins are composed of conserved N- and C-terminal domains
required for binding to the tail fiber. The actual host recognition domain of both
types of adhesins is composed of alternating highly variable regions separated by
highly conserved glycine-rich spacers [165]. Interestingly, the duplication of the
second variable domain of gp37 is sufficient to switch T4 from targeting E. coli to
Yersinia pseudotuberculosis [166]. Moreover, phage T6 and T2 can be brought to
7.4
Bacteriophages: A Re-emerging Solution?
ignore their normal host receptor, Tsx and OmpA, respectively, to recognize other
surface proteins and different bacteria through exchange of their gp37-gp38 region
with that of related phages recognizing those other surface markers [165–168]. In
addition, Salmonella phage vB_SenM-S16 is able to infect E. coli when the latter
expresses the Salmonella receptor of the phage, OmpCSTM , in lieu of the E. coli
OmpC [169].
Another example demonstrating the importance of host adsorption in host
range lies with the Salmonella phage SP6. SP6 shares most of its biology with
coliphages of the T7 family. SP6 normally recognizes the Salmonella LPS through
its six tail spikes encoded by gene gp46 but it has been shown to be infective
of E. coli strains expressing Salmonella LPS after cloning of the Salmonella rfa
(waa) locus into E. coli. SP6 could even be adapted to infect rough strains that
only have the core lipid A and no side chains on their LPS. Inversely, coliphage
T7 was shown by the same authors to be adaptable to rough Salmonella strains,
indicating that host recognition and the ensuing DNA injection are major factors
of host range limitation within the T7 family. Mutations that allowed these
adaptations were located in the tail spikes of SP6, the nozzle gene (gp33), and
other structural genes [170]. The gp33 nozzle gene is responsible for the synthesis
of the spindle-like structure that is found at the very tip of the tail and comes
into direct contact with the bacterium’s outer membrane [171]. These examples
suggest that a major limitation to phage infection is host recognition and that
when brought to recognize alternative receptors, phages can target different
hosts. This capacity is being exploited by a French company, Pherecydes Pharma,
which has devised high-throughput methods to generate T4 mutants that can
infect a variety of Gram-negative hosts [172].
These systems could be, in principle, harvested to create synthetic phages with
potentially wide host ranges. These phages could have their host ranges adjusted
ad libitum by the precise replacement of the recognition elements that confer host
specificity via genetic engineering. This capability could enable the transition from
an empirical search for phages that can target a given pathogen to the rational
design of phages with precisely engineered host ranges through synthetic biology. It may even be possible to build phages that differentiate between pathogenic
and non-pathogenic variations of a bacterium, based on surface targets or DNA
sequences.
Another determinant of phage host range involves anti-phage systems carried
by infected bacteria. The co-evolution of phages and bacteria has led to a continuous arms race wherein bacteria acquire defense mechanisms while phages develop
countermeasures [150, 173]. Many anti-phage systems are carried by prophages
or defective prophage remnants integrated into bacterial genomes and are generally described as superinfection exclusion (SIE) systems. Others appear to be
bacterially evolved and are generally recognized as abortive infection (ABI) systems. Most of these systems prevent the spread of the phage infection by triggering some sort of altruistic suicide wherein the infected bacterium dies before the
incoming phage has sufficient time to reproduce. Others may prevent phage DNA
127
128
7 Synthetic Biology and Therapies for Infectious Diseases
injection by masking the phage receptors or via largely unknown mechanisms
[150, 174, 175].
Phage resistance may also result from possession of a clustered regularly interspaced short palindromic repeats/CRISPR associated system (or CRISPR/Cas).
Detected in nearly 40% of bacterial and 90% of archaeal sequenced genomes
[176], these systems can mediate resistance to bacteriophages and other mobile
genetic elements by catalyzing double-stranded breaks (DSBs) in incoming DNA.
Although CRISPR/Cas modules display a rich, diverse phylogeny across bacterial
species [177], they retain a common mechanism of action. Composed of short,
palindromic, direct repeats separated by spacer sequences, the CRISPR locus
encodes an immature non-coding RNA, called the pre-crRNA, which is processed
into short crRNAs upon transcription. Each processed crRNA corresponds to a
nucleotide spacer sequence which, with the aid of accessory proteins or RNAs,
licenses the Cas endonuclease to cleave DNA sequences complementary to the
spacer (protospacers). In order to distinguish between self and non-self-target
sequences, protospacers must lie adjacent to a short, highly conserved and
system-specific nucleotide motif, called a protospacer adjacent motif (PAM).
Multiple lines of evidence support the role of CRISPR/Cas as an adaptive barrier
of horizontal gene transfer. Bioinformatics analysis has revealed that the majority
of spacer sequences display homology to bacteriophage genomes or mobile
plasmids, and rarely to chromosomal elements [176]. In vitro experiments have
directly demonstrated that the CRISPR/Cas system encodes sequence-specific
resistance to phages and plasmids [178, 179] and can catalyze DSBs in cognate
DNA sequences. Spacer addition and loss occurs in response to beneficial or
deleterious mobile genetic elements circulating in microbial populations [180].
Although adoption of CRISPR loci remains poorly understood, two Cas accessory proteins, Cas1 and Cas2, are implicated in de novo acquisition of spacers in
response to bacteriophage attack [181]. Moreover, protective CRISPR/Cas activity toward target sequences primes the gain of additional spacers derived from the
same piece of DNA [181]. By evolving a rapidly adaptive immunity system, bacteria have developed an effective means to defend against continually changing viral
threats.
Outside of its natural setting, biological engineers have capitalized on the
potential of CRISPR/Cas-based RNA-guided nucleases (RGNs) for a variety of
applications. Most notably, the Type II CRISPR/Cas system from Streptococcus
pyogenes has been engineered as a genome editing tool in bacteria [115], zebrafish
[182], mice [183], and human cells [184–187], among many others. RGNs can
generate targeted gene knockouts in complex eukaryotic genes by creating
DSBs that are erroneously repaired by non-homologous end joining (NHEJ),
leading to insertions and deletions that inactivate gene products. Moreover,
RGNs can elicit specific chromosomal modifications as the DSBs promote
homologous recombination between an editing template and its target. Unlike
zinc-finger nucleases (ZFNs) and transcription-activator-like effector nucleases
(TALENs) that rely on protein engineering, the target specificity of RGNs can
be reprogrammed by simple modifications of the short spacer sequences [188].
7.4
Bacteriophages: A Re-emerging Solution?
Although the specificity of CRISPR/Cas is sufficient to mediate resistance
against mobile genetic elements in prokaryotes, genome editing in human cells
using RGNs also yields high frequency off-target effects, which may need to
be addressed before it can find widespread use [189–192]. By inactivating the
catalytic residues of the Cas nuclease, the CRISPR/Cas system has also been
used to produce RNA-guided transcriptional regulators able to enact tunable
activation or repression of target genes in bacteria, yeast, and mammalian cells
[111, 193, 194]. Owing to the inherent multiplexibility of the CRISPR/Cas system,
these custom transcription factors can be easily engineered to act at multiple
loci and thus provide an effective method to implement synthetic gene networks
and to study natural network dynamics by simultaneously perturbing multiple
nodes.
7.4.3
Phage Genome Engineering for Enhanced Therapeutics
The design of artificial phages to combat infectious diseases is currently limited
by the difficulty of modifying phage genomes. Indeed, viral genomes are not engineerable while encapsidated and are thus often targeted during infection, which
only lasts for a very short time, or in vitro after DNA extraction from the capsid
if the genome is short enough (below about 50 kb) to be amenable to such manipulations. Moreover, there is a lack of information on the biology and function of
many phage genes, especially for phages outside of those well-studied in the laboratory such as λ, T4, or T7. A conventional strategy for phage engineering involves
cloning the region one wishes to modify into a high-copy-number plasmid, introducing the desired mutations in the cloned DNA fragment, infecting cells harboring the modified allele with a phage that may or may not have a counter-selectable
mutation in the locus to modify, and screening the progeny for a virus carrying
the desired modifications resulting from homologous recombination between the
donor plasmid and the recipient phage [146, 168, 195]. This process is hampered
by the fact that many regions of lytic phages are not clonable in their bacterial hosts
because of toxicity and that many phages will degrade resident DNA on entry into
a cell, including the plasmid(s) carrying the DNA to be introduced into the phage,
thus resulting in very low recombination efficiencies. Finally, screening for modified progeny can be extremely time-intensive without selective markers for the
introduced mutations or if the recombinant phage is nonviable or has a growth
defect.
Newly emerging strategies for genome engineering may accelerate the ability to
improve the efficacy of phage therapeutics. As part of their effort to create the first
fully synthetic bacterial genome, Gibson and colleagues [196] developed a whole
toolkit for the cloning and modification of prokaryotic DNAs in the heterologous
host Saccharomyces cerevisiae, where they are expected to be transcriptionally
and translationally silent. This strategy was recently adapted for phage genome
capture and redesign by Sample6 Technologies, a Boston-area startup company
focused on engineering phages as rapid microbial diagnostics [27, 197]. By using
129
130
7 Synthetic Biology and Therapies for Infectious Diseases
yeast as an orthogonal host for phage genomes, the issue of phage toxicity can
be minimized. Furthermore, S. cerevisiae provides a robust platform for targeted
genome engineering using highly efficient homologous recombination. Thus, this
technique has the potential to enable the fast and precise generation of engineered
phages for therapeutic and diagnostic applications.
One example of a function that has been genetically introduced into phages
to boost efficiency is the expression of biofilm-dispersing enzymes that help
the phage to degrade biofilms. Using phage T7 and E. coli biofilms as a model,
Lu and Collins [198] showed that engineering dispersin B expression into the
T7 genome enhanced biofilm destruction by about 2 orders of magnitude
compared with unmodified phages (Figure 7.7a). Moreover, gene expression
dspB
(1)
T7DspB
(2)
(3)
(4)
(a)
Antibiotics
(3)
lexA3
(1)
(2)
SOS response
DNA damage
(4)
M13lexA3
Cell death
(b)
Figure 7.7 (a) Engineered T7 phage
disrupts bacterial biofilms [198]. DspBexpressing T7 phage (T7DspB ) infects E. coli
(1) and produces progeny T7DspB and DspB
(2). Both phage and DspB are released by
cell lysis (3). DspB degrades the biofilm,
enabling phages to further penetrate into
the biofilm and infect more bacterial cells
(4). (b) Engineered M13 phage acts as adjuvant for antibiotic treatment against bacteria [201]. LexA3-expressing M13 phage
(M13lexA3 ) infects E. coli (1) and produces
LexA3 (2). LexA3 suppresses the SOS
response (3), which is thus unable to repair
DNA damage caused by antibiotics, leading
to cell death (4).
7.4
Bacteriophages: A Re-emerging Solution?
studies followed by network analysis can help identify genetic factors that can be
delivered into bacteria via nonlytic phages to interrupt bacterial defense systems.
For example, Dwyer et al. [199] and Kohanski et al. [200] showed that reactive
oxygen species can be generated by bactericidal antibiotics, thus contributing
to cell death. These reactive oxygen species are thought to trigger DNA damage
and subsequent repair via the SOS response. To target this pathway, Lu and
Collins [201] engineered nonlytic M13 phage to express lexA3, a non-inducible
dominant negative version of the SOS regulator lexA, to prevent activation of
the SOS response. This modified phage enhanced the efficacy of three major
classes of co-administered antibiotics, that is, β-lactams, aminoglycosides, and
fluoroquinolones, against E. coli (Figure 7.7b). These engineered phages could
resensitize resistant bacteria to antibiotics, decrease the evolution of antibiotic
resistance, and enhance the survival of infected mice when co-applied with
antibiotics. Another group demonstrated an interesting strategy for using
M13 phage to interfere with horizontal transfer through conjugation, a common dissemination pathway for plasmids carrying antibiotic resistance genes
[202].
Lytic phages kill bacteria before actually lysing them and a primary function of
lysis is to release progeny phages [203]. Lysis can potentially trigger complications
during phage therapy, such as the sudden release of endotoxin and super-antigens.
Such issues could be potentially avoided by engineering phages that do not lyse
bacteria or that lyse bacteria more gradually. Recombinant replication-deficient
phages, which are usually also lysis deficient, and recombinant lysis-deficientonly phages have been tested in several different experimental animal infection
models, where despite the lack of phage multiplication, animals were protected
against lethal bacterial infections [204–207]. These studies show that, at least
in some settings, it is not necessary to completely eradicate the pathogen
population and that simply decreasing its presence below a certain threshold is
sufficient to allow the immune system to complete the process. Another strategy
for decreasing phage virulence without completely abrogating its replication
could be to adopt the approach of Mueller et al. [208] where key genes of
poliovirus were systematically codon deoptimized so that protein synthesis
would be slowed and the viral burst delayed or reduced, thus yielding attenuated
viruses.
The immunogenicity of phages leads to clearance by the reticuloendothelial
system and potential reductions in therapeutic efficacy. Phage immunogenicity
can be diminished in order to increase the therapeutic residence time and to
limit the risk of treatment failure with repeated applications of the same (or a
very related) phage in the same patient. By serially passaging bacteriophage λ
and P22 through mice, Merril and coworkers [209] selected longer circulating
phage mutants that displayed increased antibacterial activity. This approach was
very time-consuming, but it is likely that through genetic engineering strategies
enabled by synthetic biology and antigen prediction via reverse and structural
vaccinology, the process of designing phages with lower immunogenicity could
be simplified and expedited.
131
132
7 Synthetic Biology and Therapies for Infectious Diseases
7.4.4
Phages as Delivery Agents for Antibacterial Cargos
Phages can be used to deliver DNA to recipient bacteria through several distinct
routes. Lysogenic phages propagate their genomes in bacteria either as integrated
DNA elements or as stable plasmids and, in turn, often confer their hosts with
new functions such as phage resistance systems and/or virulence factors that can
be essential for bacterial pathogenicity [150, 210, 211]. Another phage-mediated
transfer system is transduction, whereby bacterial DNA is incidentally encapsulated instead of, or in addition to, viral DNA [212]. This DNA may then be delivered to recipient cells where it can recombine into bacterial genomes. Finally,
phagemids, artificially created plasmids that harbor a phage packaging site, may be
captured during phage infection, specifically incorporated into capsids in lieu of
native phage DNA, and delivered to recipient cells via injection by phage particles
[213, 214].
Applications of all of these methods of delivering DNA via phages can be found
in the literature. For example, Edgar et al. [215] used an engineered version of the
well-studied phage λ to lysogenize antibiotic-resistant recipient cells with mutated
rpsL and gyrA genes by integrating wild-type versions of these genes. Spontaneous
mutations in these genes can confer resistance to streptomycin and nalidixic acid,
respectively, but the mutant gene products are recessive to the wild-type so that
addition of the wild-type genes restores antibiotic sensitivity. In addition, the filamentous phage M13 can be engineered with standard molecular cloning methods
and has a well-characterized packaging signal for targeting phagemids into the
progeny phage [216, 217]. M13-packaged phagemids encoding addiction module
toxins have been used to deliver and express these toxins in recipient cells to exert
antimicrobial effects [218].
Furthermore, phage P1 has been used to transfer genomic markers between
bacterial strains since the discovery of this property by Lennox in 1955 [219].
In the past decade, multiple studies have demonstrated the successful use of P1
phagemid vectors that can be packaged into P1 particles and delivered to a variety of Gram-negative organisms to transduce antibiotic markers or fluorescent
reporters [214, 220]. The advantages of this strategy may be that DNA can be delivered without placing significant selective pressure on the recipient due to phage
propagation and that concerns of phage biocontainment should be mitigated since
phagemids themselves are not self-transmissible after delivery. A recent form of
alternative therapy has been developed by Phico Therapeutics using inactivated
phages to target various bacteria, whereby genes coding for small acid-soluble
proteins (SASPs) are delivered to target bacteria via these phage-derived vehicles (SASPject™). These SASPs are DNA-coating enzymes, a desirable function
for protecting endospore DNA but a lethal one for active vegetative cells [221].
Although commonly used to deliver genetic material into bacterial hosts for
therapeutic purposes, phages can also be engineered to act as diagnostic agents
or to target drugs and other molecules to specific recipients. As mentioned
earlier, Sample6 Technologies is engineering phage cocktails that infect specific
7.5
Isolated Phage Parts as Antimicrobials
microbial pathogens and deliver genes encoding reporter proteins that generate
detectable signals, such as light, as surrogates for the presence of specific
microbes [222]. Phage-based diagnostics can be both specific and sensitive
if care is taken to engineer phages with desired host ranges and high-level
reporter gene expression. Furthermore, a survey of M13 phage delivery in the
contemporary literature reveals its frequent usage in targeting to eukaryotic
cells by fusing the coat protein with peptides that exhibit binding to specific target cells [223]. This technique has been applied toward the treatment
of cells infected by Chlamydia trachomatis. M13 phages, which display an
integrin-binding peptide and a C. trachomatis-derived peptide, are actively
endocytosed by human cells and interfere with propagation of the pathogen,
probably by preventing or lowering the export of bacterial proteins that allow
the pathogen to take control over its host [224]. In addition, bacteriophages
native to S. aureus or filamentous phages engineered to bind to this pathogen
have been used for targeted delivery of a photosensitizing chemical or a chloramphenicol prodrug to specified recipients. These engineered phages have
been used to achieve lethal (and local) effects on activation of the payloads
[225, 226].
Another possibility for the use of phage DNA delivery includes CRISPR-based
systems for genome editing. In a striking example of host mimicry, the bacteriophage ICP1 encodes its own CRISPR/Cas system to prey on V. cholerae strains that
possess a chromosomal phage defense locus [227]. The piracy of CRISPR/Cas as
a predatory strategy suggests that a similar rationale can be used to synthetically
engineer CRISPR/Cas-loaded phages to mediate sequence-specific killing of cells
harboring antibiotic resistance or virulence determinants.
7.5
Isolated Phage Parts as Antimicrobials
In addition to the use of engineered phages as antimicrobial or diagnostic agents,
phage-derived parts may find utility in antimicrobial applications. The synthesis,
expression, and optimization of such parts may be accelerated with synthetic biology tools for DNA synthesis, assembly, and mutagenesis.
7.5.1
Engineered Phage Lysins
Lysins and holins are used by bacteriophages to digest the cell wall of their hosts
and trigger their lysis. Different bacteria have different cell wall structures and
phages have consequently evolved various types of lysins to cleave these structures. In Gram-positive bacteria, the cell wall is directly accessible to the external
milieu and can be attacked by exogenous lysins. In Gram-negative bacteria, the
outer membrane protects the cell wall. Gram-positive phage lysins are usually
organized around two domains: the C-terminal domain is responsible for binding
133
134
7 Synthetic Biology and Therapies for Infectious Diseases
and specificity by anchoring the lysin to specific cell wall moieties before activating
the catalytic N-terminal domain [228]. Lysin activity is tightly controlled to avoid
early lysis and damage to surrounding cells that the released phage may subsequently prey upon. For example, the enzymatic activity of Gram-positive lysins
is only activated once the binding domain has anchored itself to the peptidoglycan. Gram-negative lysins are regulated by their release into the periplasmic space
by holins. Synergy between different lysins and between lysins and antibiotics
may delay the evolution of resistance [229–232]. Several companies are pursuing engineered lysin-based technologies, including Gangagen and Contrafect, as
novel antimicrobial agents.
The modularity of Gram-positive phage lysins allows one to target specific bacteria by engineering the binding domain and the nature of the bond to be cut
through the enzymatic domain. As a proof of concept, Diaz and colleagues [233]
swapped the catalytic domain of a streptococcal lysin, allowing it to keep its recognition specificity, but changing the cell-wall degradation mechanism. Fueled by
the ever-cheaper cost of DNA synthesis and enhanced genetic engineering techniques, this approach is being applied to create new types of lysins that perform
better than natural ones in terms of bactericidal properties and/or pathogen coverage [232, 234–236]. For example, a hybrid lysin made with the catalytic enzyme
of the Twort phage and the binding domain of phiNM3 lysin was easier to produce than either of the two original lysins and also had elevated activity against its
target, MRSA [232].
As mentioned above, a major hurdle for the use of lysins to treat Gramnegative bacteria is the inaccessibility of the cell wall to external agents due to
the presence of the outer membrane. However, Lukacik et al. [237] recently
succeeded in designing synthetic phage lysins that efficiently target Gramnegative bacteria by fusing a lysin to outer membrane transport signals from
bacteriocins. Bacteriocins are small proteinaceous molecules that are synthesized
by a wide variety of bacteria to control the population of other prokaryotes
present in their environment. Bacteriocin import into target cells is dependent
on the recognition of specific receptors, and producing strains are immune
to their own bacteriocins through the production of inactivating antidotes
[238]. For example, the FyuA outer membrane protein is a virulence-related
iron transporter found in some strains of Yersinia and E. coli, which is specifically recognized by the Yersinia bacteriocin pesticin [239, 240], leading to
its uptake into the bacterium’s periplasm. The FyuA protein is a useful target
considering its strong association with pathogenic bacteria [240] and relative
absence from non-pathogenic microbes. This bacteriocin was fused to the
cell-wall-degrading lysozyme of phage T4 based on 3D-structural analyses
to ensure that the fusion would not impair receptor recognition or lysozyme
function (Figure 7.8). The resulting protein efficiently killed FyuA-displaying
Gram-negative bacteria [237]. This strategy could, in principle, be reproduced
with other bacteriocin receptors for specific applications against Gram-negative
pathogens.
7.5
Conserved FyuA binding domain
Figure 7.8 Structural alignment of the wildtype Yersinia toxin pesticin (pink) and the
first anti-Gram-negative engineered lysin
(blue). By keeping the receptor-binding
domain of pesticin and fusing it to the
cell-wall degrading lysozyme of phage T4,
Lucacik and his colleagues created a antimicrobial molecule that specifically recognizes FuyA expressing cells and lyses them
independently of whether those cells are
Isolated Phage Parts as Antimicrobials
Variable toxin domain
sensitive to the original toxin pesticin. This
structural alignment was done using the Vector Alignment Search Tool (VAST) server on
the NCBI website and visualized using Cn3D
software [241] and highlights the conservation of the recognition domain and the
variability of the catalytic domain. (Reference
[417] Reproduced with permission. Copyright
© 2007 Elsevier.)
7.5.2
Pyocins: Deadly Phage Tails
Many pseudomonads are bactericidal to other Pseudomonas strains. Upon stress
conditions, such as starvation or DNA damage, pseudomonads can release phagetail-like particles called pyocins that can kill susceptible pseudomonads by depolarization of their membrane potential. Pyocins belong to two families depending
on their morphology. Those that appear to have contractile sheaths belong to the
R-type family while those that resemble flexible phage tails belong to the S-type
family [242]. In both cases, bactericidal activity is dependent on specific recognition of an outer membrane receptor by the pyocin, thus limiting host range.
Exchange of the host recognition domain located at the tip of the tail fibers has permitted rerouting of R-pyocins from their usual hosts toward various E. coli strains
[243–246]. As myophages having similar tails to R-pyocins are widespread in
nature, it is conceivable to create synthetic pyocins targeting virtually any Gramnegative species. In addition, pyocins have recently been discovered in the Grampositive bacterium C. difficile, which suggests that their scope of action may be
even further broadened via genetic engineering [247].
135
136
7 Synthetic Biology and Therapies for Infectious Diseases
7.5.3
Untapped Reservoirs of Antibacterial Activity
With phages outnumbering bacteria by a factor of 10 or more [248] and most
new phage genomes showing a majority of genes with no known homologs in
databases, it is likely that the vast wealth of phage biology still has much to teach
us about killing bacteria. From the highly limited sample of phages that have been
heavily studied, it is already clear there is a plethora of ways to appropriate host
cell resources. As an example, coliphage T4 encodes proteins that digest bacterial
DNA for selfish reuse (e.g., ndd, various restriction endonucleases) and others
that stop host transcription and redirect transcription toward its own genes (e.g.,
alc, asiA, motA, etc.). Moreover, some of its early promoters are so strong that
they are capable of sequestering RNA polymerase away from bacterial use. All of
those functions are orchestrated to lead to bacterial death within minutes of infection [249]. In addition, T4 [250] as well as other phages such as phiEco32 [251]
and phiEcoM-GJ1 [252] may encode their own sigma factors that can efficiently
compete with endogenous host sigma factors to redirect the host transcription
machinery onto phage genomes. Finally, other phages free themselves from many
host requirements by encoding their own RNA polymerases and other important
functions. Examples include coliphage T7 [253], Xanthomonas phage XP10 [254],
Pseudomonas phage phiKMV [255], and their many relatives. Detailed screens
for phage genes with antimicrobial activity are being performed by academic and
industrial laboratories in the hope of discovering new bioactive compounds or
killing mechanisms that could be exploited in the fight against pathogens with
multidrug resistance. These efforts can be accelerated through cheaper and more
rapid gene synthesis and expression technologies.
7.6
Predatory Bacteria and Probiotic Bacterial Therapy
New genetic engineering technologies and strategies are not only accelerating
phage-based antimicrobials, but are also setting the stage for an entirely novel
class of anti-infectives in the form of cell-based therapeutics [256].
For example, some bacteria have evolved to prey on other bacteria, either exclusively or facultatively, for growth. The most widely studied group of such predatory bacteria is composed Gram-negative prokaryotes that share three features:
a curved rod morphology, small size, and the predatory lifestyle. These are collectively referred to as Bdellovibrio and like organisms (BALO) [257, 258]. BALO
are motile bacteria that swim, invade the periplasm, or attach to the surface of
their prey, and reproduce at the expense of the infected bacteria, producing fewer
than five progeny BALO per infection [259]. Most BALO are promiscuous and
will readily prey on any Gram-negative bacterium. Their wide bactericidal activity
and their ubiquity in the environment [260, 261] and the human gut [262] have
7.6
Predatory Bacteria and Probiotic Bacterial Therapy
led to interest in their potential use as therapeutics [263], and small-animal trials
have yielded promising results [264, 265]. However, very little is known about
how BALO bacteria regulate predation, with concerns over their ability to indiscriminately prey on both pathogenic and commensal bacteria. In time, it may be
possible to engineer more stringent prey specificity, but technologies to manipulate their genomes and study their biology are just emerging [266]. Moreover,
the very high hydrolase content of BALO may also present opportunities for the
development of bactericidal compounds [267].
In addition to BALO, probiotic bacteria are being explored for their antimicrobial and immunomodulatory properties. Probiotic efficacy has been studied in
various applications such as the treatment and prevention of allergies [268], C. difficile colitis [269], inflammatory bowel diseases [270, 271], shigellosis or salmonellosis [272–274], and even complex metabolic diseases such as diabetes, metabolic
syndrome, or obesity [275]. Moreover, consumption of Lactobacillus rhamnosus
GG has been shown to be effective at eliminating vancomycin-resistant Enterococci (VRE) from the gastrointestinal tract of colonized children [276]. However,
the literature on probiotics studies have yielded irregular and conflicting results,
which may be due to factors such as heterogeneity in probiotic strains, preparations, and delivery methods; thus, stronger standards and regulations around the
use and marketing of probiotic bacteria are likely needed to ensure reproducible,
safe, and efficacious use [268, 277, 278].
Although these natural probiotic therapies have demonstrated promise in the
treatment of bacterial infections, biological engineering can potentially further
augment the efficacy of probiotics. An application of our growing understanding of interbacterial communication and its potentiation by synthetic biology can
be found in the work of Duan and colleagues [279], who developed a probiotic
strain that can decrease V. cholerae virulence in a murine model of infection by
targeting its quorum-sensing-based signaling pathways. Quorum sensing refers
to a process through which bacteria modulate gene expression in response to a
critical concentration of self-produced autoinducer (AI) molecules in the environment [280]. Quorum sensing has been described as a social behavior in bacteria
in which populations undergo concerted changes in gene expression to produce
gene products that are only beneficial at high cell densities [280]. For example,
the marine bacterium Vibrio fischeri employs a quorum-sensing mechanism to
only achieve bioluminescence at high cell densities in the light organ of its symbiotic host, but not when free-living in the ocean [281]. Similarly, P. aeruginosa uses
quorum-sensing circuits to control the expression of virulence factors, such as
extracellular proteases and phenazine pigments, which promote infection [281].
This population-level coordination of gene expression ensures that individual cells
will not incur a strong metabolic cost for gene products that are only valuable at
high cell concentrations. In V. cholerae, the expression of virulence genes, such
as cholera toxin and the toxin co-regulated pilus, is repressed in the presence of
two autoinducers, AI-2 and cholera autoinducer 1 (CAI-1). Thus, Duan and colleagues engineered the probiotic E. coli Nissle 1917, already a native producer of
AI-2, to constitutively express and secrete CAI-1 to molecularly confuse virulence
137
138
7 Synthetic Biology and Therapies for Infectious Diseases
(2)
(3)
CqsS
CAI-1
(4)
(1)
Signal transduction
(5)
cqsA
Engineered E. coli
Figure 7.9 Engineered probiotic E. coli
prevents cholera infection in infant mice
[279]. Engineered E. coli carrying cqsA produces cholera autoinducer 1 (CAI-1) (1), and
secretes it into the environment (2). At high
cell densities, the CAI-1 concentration is
Virulence expression
V. cholerae
sufficiently elevated to permit binding and
inhibition of its cognate sensor kinase, CqsS,
in V. cholerae (3). CqsS inactivation triggers a
signal transduction cascade (4) that leads to
the repression of virulence genes (5).
gene regulation in V. cholerae (Figure 7.9). Pre-treatment of mice with the modified probiotic increased survival by 92% and decreased cholera toxin binding to
the mouse intestine by 80%. As this type of therapy targets a component of virulence as opposed to a core component of bacterial physiology, one would expect
that resistance would emerge at a much slower rate than conventional chemical
antibiotics. This work serves as an illustrative example of using synthetic biology
to rationally design probiotic bacteria that can defend against infectious disease.
As discussed above, a myriad of quorum-sensing mechanisms have been
described in a wide range of bacterial species [281], with each quorum-sensing
molecule often acting as a molecular signature of a particular species of microorganism. Using these signatures, both Saeidi et al. [282] and Gupta et al. [283]
developed “sentinel cells” able to sense and destroy P. aeruginosa in vitro. LasR,
a transcription factor that responds to an acyl-homoserine lactone quorumsensing molecule produced specifically by P. aeruginosa, was introduced into
E. coli such that the presence of the pathogen would elicit gene expression of
a toxic bacteriocin and a lysis protein. At a critical intracellular concentration
of the lysis protein, the E. coli chassis would burst open, releasing pyocin and
killing P. aeruginosa. This engineered system could mediate a 100-fold reduction
in P. aeruginosa cell numbers in planktonic cultures and could prevent biofilm
formation by nearly 90%. Further work demonstrated that sentinel cells could
be engineered to actively migrate toward P. aeruginosa biofilms and secrete
antimicrobial peptides and DNases to elicit cell killing and biofilm disassembly
[284]. These E. coli sentinels lay the groundwork for the development of more
sophisticated synthetic “immune cells” that may be able to integrate and compute additional environment cues and eradicate a wider range of pathogens.
Further use of synthetic gene circuits that can implement integrated logic and
memory or perform wide-dynamic-range analog biosensing could enhance the
sense-and-response capabilities of these systems [285–288].
7.7
Natural Products Discovery and Engineering
An exciting extension of probiotic therapy is the prospect of microbiota
engineering with the tools of synthetic biology. The potential of this approach is
inspired by encouraging results from microbiome transplantation, the practice
of colonizing a diseased individual with the microbiome from a healthy donor,
which has shown promise in the treatment of recurrent C. difficile infections
[289]. Although the mysteries of the human microbiome are only starting to be
uncovered [290], we expect that synthetic biology strategies for manipulating
the human microbiota in a precise and controlled manner may open up exciting
new applications in the decades to come. To achieve this promise, new tools for
additive and subtractive manipulation of complex microbial communities at the
level of bacteria and their constituent genes are needed.
7.7
Natural Products Discovery and Engineering
Many clinically relevant molecules are of natural origin, meaning they are produced by living organisms, such as bacteria, fungi, and plants [291]. Antibiotic
discovery has been dominated by microbially secreted molecules since the discovery of penicillin in 1928 [292]. Sir Alexander Fleming observed that microorganisms could produce molecules that kill or inhibit growth of their counterparts. The
realization that these molecules could treat infections led to Fleming, Ernst Chain,
and Sir Howard Florey receiving the Nobel Prize in Physiology [293, 294]. Bacteria have dominated the study of antimicrobial production, especially soil-dwelling
isolates belonging to the Streptomyces genus. Given the success with Streptomyces
spp., many efforts have been made to find novel natural products from microorganisms that can act as antimicrobials [295–297]. Below, we describe extensive
efforts to accelerate the discovery and modification of natural products through
computational and genetic engineering strategies. These efforts can further leverage the tools of synthetic biology to accelerate the construction, expression, and
optimization of new natural product pathways [298, 299], as discussed in greater
detail in the following sections.
Actinomycetales is an order of Gram-positive bacteria that includes the genus
Streptomyces and is among the most prolific sources of natural products. Actinomycetales have yielded about 3000 antibiotics as well as antitumor, antiparasitic,
antiviral, and immunosuppressant drugs [291, 300]. These molecules, which
include non-ribosomal peptides (NRPs), polyketides (PKs), and hybrids of both,
do not result from direct ribosomal translation. Instead, very large enzymatic
complexes, called nonribosomal peptide synthetases (NRPSs) and polyketide
synthases (PKSs), assemble complete natural-product molecules by selecting,
uploading, and modifying particular precursors, namely, aryl and amino acids
(for NRPs) and acetyl/malonyl/methylmalonyl-CoA (for PKs) [301–304]. NRPS
and PKS are modular: each module is usually capable of selecting and loading
unit to the nascent molecule by specific domains [301, 302]. Specialized domains
139
140
7 Synthetic Biology and Therapies for Infectious Diseases
also allow for additional structure modifications and release of the molecule from
the enzymatic machinery [301–304].
In the assembly of NRPs, the monomer to be incorporated into the growing
molecule is selected by the adenylation domain. This domain contains a specific
amino acid signature that determines the aryl or amino acid to be incorporated,
which is subsequently transferred to the downstream thiolation domain. The thiolation domain is characterized by containing a thiol group, to which monomers
or the growing molecule is attached. When additional units are incorporated, an
amide bond is formed between the aminoacyl group of a downstream unit and
the peptidyl group of the upstream one. This step is catalyzed by the condensation domain [301, 303]. For NRPSs, the core domains for chain elongation are the
condensation, adenylation, and thiolation domains, forming the most basic module. The first module – initiation module – is traditionally a two-domain module
consisting of an adenylation and thiolation domain, as no condensation reaction
is required for the first module. Termination of chain elongation is carried out by
a thioesterase domain. This domain catalyzes the release of the molecule from the
enzymatic machinery [301, 303].
The assembly of PKs shares the logic with the assembly of NRPs. Here, the three
core PKS domains for chain elongation have homologous functions with those of
the NRPSs: the acyltransferase is responsible for selecting the unit to be incorporated (either malonyl or methylmalonyl-CoA). Also, a thiolation domain provides
the thiol group for unit and growing chain attachment. C–C bond formation in
PKs is catalyzed by the ketosynthase domain [301, 302, 304]. Chain translocation
across NRPS–PKSs and PKS–NRPSs can occur, thus resulting in hybrid PK–NRP
molecules [301, 302]. We refer the interested reader to extended reviews on NRPS
and PKS biochemistry [301–304].
7.7.1
In Silico and In Vitro Genome Mining for Natural Products
With the rapid development of next-generation DNA sequencing technologies,
the cost for full genome sequencing has plummeted. This has enabled natural
product scientists to collect a wealth of informatics data, mostly freely accessible
from the National Center for Biotechnology Information (NCBI), that can be
screened for gene clusters predicted to biosynthesize natural products [305].
Nonetheless, it is important to highlight that, despite the presence of biosynthetic
genes, one cannot directly assume that these genes are active. In some cases,
biosynthetic clusters are expressed only under very specific conditions or not at
all (cryptic/silent clusters) [295, 296, 306]. Software algorithms, which specifically
search for subregions of genes that will be translated into catalytic domains, can
be used to determine whether a certain microorganism might be able to produce
an NRP or PK. These domains determine enzymatic activity and can be used to
predict the type, and sometimes even the structure, of the molecule that will be
generated (Figure 7.10). Examples of these software tools include the following:
7.7
Natural Products Discovery and Engineering
DNA
Predicted protein domains
Predicted precursors
Predicted molecule
(a)
DNA
Predicted protein domains
Predicted precursors
P
P
Precursor labeling
Labeled molecule
P
(b)
Figure 7.10 In silico genome mining allows
for fast and less laborious screening of large
sets of genomic data for the potential capacity to produce natural products. (a) Software
algorithms can be used to predict the protein coded by a gene and even the structure
of the final natural product. (b) These predictions can be confirmed by feeding a labeled
presumed precursor and tracing its incorporation into molecules, a process that can
culminate with the identification of the natural product.
• ClustScan (cluster scanner) allows for a quick scan of genomes in the search
for DNA sequences encoding NRPSs and PKSs. It offers additional tools
for genome annotation and displays e-values, protein alignments, and both
nucleotide and amino acid coordinates for the regions of interest detected.
Another useful feature of this software is the prediction of possible structures
for the molecules encoded by the identified clusters. It uses a simplified
molecular-input line-entry system (SMILES) that enables the description of
chemical nomenclature as well as the structure of a given molecule. It can
be converted into two-dimensional and three-dimensional structures using
141
142
7 Synthetic Biology and Therapies for Infectious Diseases
appropriate software freely available online. ClustScan is available for download
at http://bioserv.pbf.hr/cms [307].
• NP.Searcher (natural products searcher) is a web-based software that predicts
the structure of natural products from the analysis of input data (full or partial
genome sequences, in FASTA format). It then produces SMILEs that can
be translated into a structure using other available SMILE converter tools.
NP.Searcher can be found at http://dna.sherman.lsi.umich.edu [308].
• antiSMASH (antibiotics and secondary metabolite analysis shell) is a webbased software that allows for full-genome analysis of clusters for a wide range
of molecules: PKs, NRPs, aminoglycosides, beta-lactams, aminocoumarins,
butyrolactones, siderophores, indolocarbazoles, terpenes, bacteriocins, lantibiotics, nucleosides, and melanins, among others. This tool also reports
on the similarity of clusters within the sequence under analysis with fully
characterized and available sequences of clusters for other known molecules.
In addition, one can click on ORFs and automatically submit the corresponding
amino acid sequence for a blast alignment at NCBI. antiSMASH can be found
at http://antismash.secondarymetabolites.org [309].
Despite the utility of the aforementioned tools, careful annotation of catalytic units often requires the use of additional databases. These allow for
more thorough and precise mining, prediction and annotation of protein
domains. Some of the most thorough and reliable are: Web-CD search
tool – http://www.ncbi.nlm.nih.gov/Structure/bwrpsb/bwrpsb.cgi [310–312],
SMART – http://www.ebi.ac.uk/interpro [313, 314], NRPSpredictor – http://ab.
inf.uni-tuebingen.de/software/NRPSpredictor [315], NRPSpredictor2 – http://
nrps.informatik.uni-tuebingen.de/Controller?cmd=SubmitJob [316], PKS/NRPS
analysis website – http://nrps.igs.umaryland.edu/nrps [317], and SBSPKS –
http://www.nii.ac.in/∼pksdb/sbspks/master.html [318].
With these tools, it is possible to glean important information on the structural
and physiochemical characteristics of predicted products, which can help design
targeted isolation procedures [307, 319, 320]. Any such information is useful,
considering the cumbersome task of purifying metabolic products [305, 321].
Predictions made by these bioinformatics tools are usually validated with in vivo
and/or in vitro experimentation. One experimental approach to the identification
of molecules originating from a given cluster is the genomisotopic approach,
which involves feeding stable-isotope-labeled predicted precursors to the producing organism [305, 322]. If incorporated as predicted, the resulting molecule
is immediately labeled, facilitating its identification within the complex mixture
of molecules and thus the subsequent purification of the natural product of
interest. Other authors [323] have devised peptidogenomic methods using massspectrometry (MS)-guided methods that connect the chemotypes of peptidic
natural products to the biosynthetic gene cluster, following the biosynthetic logic
of NRPs [323, 324].
In silico genome screening for natural products is relatively straightforward
using existing tools, but it is limited to microorganisms whose genomes are
7.7
Natural Products Discovery and Engineering
TTCGTAATGCTAGCTATGCCT
ACCCTTAGTCTATTGAAAGG
CCCTAAGCCCTAGATCGATAA
CTCGGATAAATGCCTGCTATC
AAATGCTCCCTGGCTCGCTCT
AGAAATCGCTACTTAGATTTC
(b)
(a)
P
(c)
Figure 7.11 Strategies used for wholegenome mining in the search for naturalproduct-encoding gene clusters: (a) in silico genome mining, using algorithms that
specifically search for natural product-related
enzymes; (b) high-throughput phage display of phosphopantetheinyl transferases;
and (c) in vitro screening of natural-productrelated genes by degenerate primers and
DNA probes.
sequenced. This limits the number and diversity of genomes that can be explored,
especially if one takes into account that the number of genome sequences available is an extremely small fraction of the microorganisms in nature. According
to the Joint Genome Institute (US Department of Energy), more than 4300
genomes belonging to the bacteria, eukarya, and archaea domains have been
fully sequenced so far, and over 12 000 other sequencing projects are in progress
[325]. Yet, this is a very small proportion of the ∼500 000 species of bacteria
predicted to exist [326]. Complementary technologies for genome mining have
been developed, including high-throughput in vitro strategies (Figure 7.11a).
For example, Yin et al. [327] performed high-throughput mining of bacterial
genomes with a phage display strategy. This study leveraged Sfp, a protein with
phosphopantetheinyl transferase activity, which is crucial for the activation of
NRPS and PKS carrier proteins. This enzyme was used to biotinylate carrier
proteins expressed from genome fragments, which were fused to a phagecoat-protein gene. The biotinylated phages were subsequently enriched and
the proteins/genes from NRPS/PKS clusters were then sequenced. Using this
technique, one can perform high-throughput screens for isolates that have been
sequenced as well as for entire communities of microorganisms, including those
considered unculturable (Figure 7.11b) [322, 327].
An alternative technique for the detection of NRP/PK-related biosynthetic
genes is the use of labeled DNA probes and Southern blotting. The labeled probe,
if designed to hybridize with a region of genes coding for natural products, will
hybridize with homologous regions within the sample. Schirmer and colleagues
143
144
7 Synthetic Biology and Therapies for Infectious Diseases
[328–330] designed ketosynthase probes to search at low stringency for PKS
genes within metagenomic libraries containing over 4 Gb of bacterial genomes
obtained from samples of marine sponges; others have performed studies using
a similar approach. Similarly, sets of degenerate PCR primers can be used to
screen for the presence of NRP/PK biosynthetic clusters, expressed or cryptic, in
metagenomic libraries and/or microbial isolates (Figure 7.11c) [331–334].
DNA sequence availability opens a whole range of opportunities for the discovery of novel, bioactive, and natural products. In the absence of genome sequences,
diversity is generated through the screening of organisms for the production of
novel molecules. DNA sequence availability and new DNA synthesis technologies
afford researchers the ability to synthesize parts and pieces that can be altered,
combined, and swapped to generate unnatural diversity. Furthermore, knowing
the genetic makeup of a biosynthetic pathway and understanding how the pathway cooperates to generate a known molecule can help researchers create optimal or novel synthetic pathways through combinatorial and rational design [320,
335–341].
7.7.2
Strain Engineering for Natural Products
There are several reasons why one would choose a heterologous host, rather than
the native producer, for the expression of genes, operons, or clusters, with a major
factor being genetic tractability. For model organisms such as E. coli, Arabidopsis thaliana, and S. cerevisiae, an incredibly wide range of genetic tools has been
developed over the years that enable the facile engineering of their genomes [342,
343]. Yet, the same cannot be said for other organisms, in which routine protocols
used in model organisms are often difficult or even impossible. When engineering pathways and/or overexpressing molecules, the use of well-characterized hosts
allows for a better understanding of the behavior of interest, enhanced abilities to
manipulate strains or pathways and perhaps the capability to achieve higher titers
of the target molecule [344, 345].
Another reason involves growth conditions and speed. For example, although
the diversity of molecules that are produced by Streptomyces spp. remains
unmatched by any other known genus, cell growth is slow and the production of
target molecules is generally dependent on medium composition [321, 346]. The
existence of transcriptional repressors and/or unactivated activators can limit
the titer of a given molecule in its native host. By transferring the pathway of
interest to a heterologous host, one can eliminate unknown regulation and place
the biosynthetic pathway under control of inducible or constitutive promoters,
thus transferring control of expression over to the researcher [347, 348].
Reengineering pathways involved in the biosynthesis of natural products
might, at first, seem a relatively easy undertaking from a technical standpoint.
Yet, the issues faced by synthetic biologists in assembling novel pathways
include challenges associated with manipulating large DNA fragments, dealing
with downstream issues such as ensuring precursor flux, and deciphering the
7.7
Natural Products Discovery and Engineering
grammar underlying the assembly of proteins and domains to achieve proper
molecular synthesis. Although natural products are small, the proteins involved
in their assembly are particularly large, varying between 200 and 2000 kDa [323,
349]. For example, cyclosporin synthase is a 1.7 MDa enzyme catalyzing 40
different reaction steps to synthesize the immunosuppressant cyclosporin from
precursors. This enzyme is encoded by a mega-gene, spanning 45 800 bp of DNA
[350]. Thus, the heterologous expression and rational reprogramming of natural
pathways often result in the challenging cloning and manipulation of synthetic
genes [338]. Fortunately, advances in DNA synthesis and assembly techniques
(e.g., Gibson, Golden Gate, and other assembly strategies) may provide the
ability to tackle these DNA construction challenges [351, 352]. Despite these
potential advantages, heterologous hosts can also be limited by differences in
precursor availability, challenges in achieving optimal levels of gene expression
with synthetic regulation, and variations in codon usage and post-translational
modifications [346, 353, 354].
Given the diversity of genes involved in the biosynthesis of natural products
that display clinically relevant activities, engineered combinations of these clusters might produce functional, novel molecules. For example, the replacement of
a gene from the antitumoral drug echinomycin with one from the poorly characterized SW-163 compound resulted in the novel NRPS, ecolimycin. Although
this hybrid pathway was originally from Streptomyces spp., it was completely and
successfully engineered in E. coli [355]. When constructing synthetic pathways
in heterologous hosts, it is important to keep in mind that precursors may be
lacking or that regulatory mechanisms need to be fine-tuned. These factors can
result in pathway imbalances that result in low product titers; thus, computational
and experimental approaches have been employed to circumvent these challenges
[356]. Below, we discuss several examples where natural product pathways have
been engineered within heterologous hosts as well as their native hosts using the
tools of genetic engineering and synthetic biology.
7.7.2.1 Production of the Antimalarial Artemisinin
Every year, the insect-borne malarial disease caused by the Plasmodium parasite sp. infects over 200 million new people [357]. Malaria is particularly severe
in poor, developing countries and mostly affects children under 5 years of age
who lack adequate immunity [357, 358]. Aggravating this issue is the fact that the
commonly used drugs for malaria, chloroquine and sulfadoxine-pyrimethamine,
are now mostly ineffective, as the most important etiologic agents, P. falciparum
and P. vivax, have developed resistance. Currently, the World Health Organization
(WHO) advises the use of artemisinin and derivatives, in combination with schizontocidal drugs (artemisinin-based combination therapies, ACTs) as efficient
therapies with reduced probabilities for evolving resistance [344, 359]. However,
in 2007–2008, only 15% of infected children under the age of 5 with malarial fever
actually received ACT treatment [357].
Artemisinin is a sesquiterpene lactone peroxide naturally produced by the shrub
Artemisia annua. There is a scarcity of this molecule, with prices fluctuating with
145
146
7 Synthetic Biology and Therapies for Infectious Diseases
variations in supply – for example, its price shifted from $350 kg−1 to as much
as $1000 kg−1 in 2004. Despite the potential to chemically synthesize this natural
molecule, the complexity of the process, aggravated by low yield, has discouraged
production in vitro [357, 360]. An early approach to increase artemisinin availability was to cultivate the shrub and extract the active ingredient. In China and
Vietnam alone, there was a approximately sevenfold increase in the number of
companies extracting artemisinin [360]. Yet, considering the massive number of
new patients every year and the need for low cost production, synthetic biologists have worked to direct the artemisinin production pathway from A. annua to
heterologous hosts, such as bacteria and yeast. The ability to use industrial-scale
fermentation for the growth of an artemisinin-producing organism would allow
for large quantities of an active pharmaceutical ingredient (API) to be produced in
a short period of time [358, 360, 361]. The Keasling group undertook the challenge
of developing a yeast-based platform for the production of artemisinin through a
low-cost fermentation method, in collaboration with Amyris Biotechnologies and
the Institute for OneWorld Health and with funding from the Bill and Melinda
Gates Foundation [344, 357, 358]. The strategy was to engineer the artemisinin
production pathway in S. cerevisiae for the generation of precursors all the way up
to artemisinic acid. This engineering endeavor involved three steps:
1) Engineering the farnesyl pyrophosphate biosynthetic pathway – in this pathway, the production of farnesyl pyrophosphate is upregulated and its conversion to the unwanted sterols downregulated.
2) Cloning of the amorphadiene synthase (ADS) gene from A. annua – this
enabled the heterologous host with the capacity to convert farnesyl
pyrophosphate to amorphadiene, a step required for artemisinin production.
Amorphadiene production was further increased by the overexpression of a
truncated and soluble 3-hydroxy-3-methylglutaryl-coenzyme A reductase.
3) Cloning the gene for a novel cytochrome P450 capable of converting amorphadiene to artemisinic acid through a three-step oxidation reaction. This is
one of the most important enzymes involved in the production of natural
products by plants [361].
The adjustments performed on this heterologous host allowed for a large
increase in amorphadiene production, rising from 370 μg l−1 to 153 mg l−1 . The
artemisinic acid recovery rate reached 96% from the cell pellet via alkaline buffer
washing [344, 362]. Additional advantages of this production method are that it is
relatively environmentally friendly (no additional terpenes produced, contrary to
plant hosts) and that its production depends solely on the fermentation capacities
of an engineered facility rather than being dependent on climatic variations [344].
In 2012, Westfall et al. [357] described S. cerevisiae-based processes for the
production of amorphadiene, a precursor to artemisinic acid. Production levels
above 40 g l−1 were achieved, which was a considerable increase from the 0.5 g l−1
achieved previously in the bacterial host (E. coli) and 153 mg l−1 in yeast. The subsequent step converting this molecule to dihydroartemisinic acid was also engineered [357, 363]. Yet another yeast strain was engineered for the production
7.7
Natural Products Discovery and Engineering
of the artemisinin precursor artemisinic acid by building a construct expressing three plant genes involved in this process: ADS, amorphadiene oxidase, and
cytochrome P450 reductase. Together, these genes have the capability to channel carbon flux from farnesyl diphosphate to the artemisinin pathway, resulting in
titers of 1 g l−1 of artemisinic acid [364–366].
7.7.2.2 Daptomycin (Cubicin)
Daptomycin is a potent U.S. FDA-approved lipopeptide antibiotic naturally produced by Streptomyces roseosporus. Used in the treatment of skin and skin structure infections, this NRP targets both the cell wall and membrane, acting as a
cationic dipeptide in Gram-positive bacteria [367, 368]. Daptomycin’s fermentation process depends highly on the precursor feed into the bioreactor. Lipid
removal from the medium leads to the production of a family of lipoproteins
(A21978C) by S. roseosporus, which vary in their N-terminal fatty acid chain [367,
369]. By adding decanoic acid during the fermentation process, biosynthesis can
be directed toward the production of daptomycin.
Attempts have also been made at improving the yields of daptomycin production by expressing daptomycin biosynthetic genes in heterologous hosts.
Yet, results so far have not reached a high level or even the same levels as the
natural producer strains, thus requiring additional engineering. For example,
Streptomyces lividans has been used for chromosomal and extrachromosomal
expression of these genes under the control of a constitutive strong Streptomyces
promoter, ermE*p [369]. Engineered S. lividans produced 10 mg l−1 daptomycin,
which was lower than the native S. roseosporus host, where it varies from
150 mg l−1 to 1 g l−1 . By deletion of the biosynthetic cluster coding for the
benzoisochromanequinone polyketide antibiotic actinorhodin and adjustment
of growth medium composition, production levels rose to 55 mg l−1 [367, 369].
Additional engineering is still required to achieve yields comparable to those
obtained by native lipopeptide expression. Interestingly, novel compounds
with increased structural diversity and activity spectra were also generated by
genetically engineering Streptomyces fradiae clusters for daptomycin and the
related A54145 molecule, also a lipopeptide antibiotic [370, 371].
7.7.2.3 Echinomycin
Echinomycin is a quinoxaline antibiotic that contains quinoxaline chromophores
attached to the depsipeptide core structure and displays potent antibacterial,
anticancer, and antiviral activities [372]. It has been isolated from an assortment of bacteria, namely, Streptomyces lasaliensis and Streptomyces echinatus
[373–375]. Upon discovery of the gene cluster responsible for the biosynthesis
of this molecule within a linear plasmid in S. lasaliensis, successful attempts
were made to overexpress it in E. coli, at a concentration of 0.6 mg l−1 [372].
Levels as high as 126 mg l−1 of echinomycin have been reported, upon medium
optimization [375].
147
148
7 Synthetic Biology and Therapies for Infectious Diseases
The endeavor to express echinomycin in E. coli involved the construction of a
synthetic cluster derived from the original echinomycin one. The cluster was separated into three plasmids comprising 15 genes from S. lasaliensis and sfp, which
encodes an acyl carrier protein (ACP) activator, from Bacillus subtilis. Each gene
was given a dedicated T7 promoter, ribosome-binding site, and T7 transcriptional
terminator. Moreover, by removing a single gene from this multi-plasmid and
multi-monocistronic synthetic pathway, it was possible to convert echinomycin
to the antitumor agent triostin A [372].
7.7.2.4 Clavulanic Acid
β-lactam antibiotics contain a lactam ring and can be used to treat Gram-positive
and Gram-negative bacteria by targeting bacterial cell wall biosynthetic enzymes,
called penicillin-binding proteins (PBPs). This class, which includes penicillins,
cephalosporins, carbapenems, and others, represents over 65% of the antibiotic
market owing to its high effectiveness, low cost, and minimal side effects [376,
377]. However, some groups of antibiotic-resistant bacteria produce β-lactamases,
which specifically inactivate these molecules by hydrolyzing the amide bond of
their β-lactam ring [376, 377]. Clavulanic acid, discovered in 1976 as being produced by the soil organism Streptomyces clavuligerus, can be used in combination
therapy with β-lactams to inhibit the action of β-lactamases. Similarly to β-lactam
antibiotics, clavulanic acid displays a structure with a β-lactam ring. However,
rather than being hydrolyzed from the enzyme after catalysis and permitting the
turnover of an additional substrate, this compound remains irreversibly bound to
the active site of β-lactamase. Formulations combining this β-lactamase inhibitor
and β-lactams are available and represent several billion U.S. dollars in annual sales
[378].
The production of clavulanic acid still relies on the fermentation of
S. clavuligerus. Thus, there is interest in potentiating this bacterium’s capacity to produce the β-lactamase inhibitor. Seeking to increase the titers of
clavulanic acid, researchers have resorted to intensive random mutagenesis and
screening-based techniques. For example, Li and Townsend were able to double
the yields obtained from the fermentation of S. clavuligerus. To achieve this,
two distinct glyceraldehyde-3-phosphate dehydrogenases (gap1 and gap2) were
removed from the glycolytic pathway. By disrupting these genes, glyceraldehyde3-phosphate converted from glycerol is channeled to the carboxyethyl-L-arginine
synthesis pathway, instead of the Krebs cycle, thus leading to the production of
clavulanic acid [378].
Deletion of other genes also enhanced production levels of clavulanic acid. By
generating null mutants of relA, which codes for a protein involved in the regulation of phosphorylation of guanosine nucleotides, Gomes-Escribano et al. [379]
engineered S. clavuligerus to overproduce both clavulanic acid and the β-lactam
antibiotic cephamycin C. Clavulanic acid levels were three to fourfolds higher than
in the wild type, and cephamycin C was overproduced by up to 2.6-fold.
Efforts have also been undertaken to understanding the difference, at the
whole-genome level, between clavulanic-acid-producing strains and their
7.7
Natural Products Discovery and Engineering
wild-type counterparts. This research can enable one to better understand the
natural mutations, which may result in a higher yield of the drug and can suggest
additional targets for mutagenesis to further enhance production levels [380].
7.7.2.5 Production of the Antiparasitic Avermectin and Its Analogs Doramectin and
Ivermectin
In the late 1970s, Merck and the Kitasato Institute in Japan discovered a family
of potent antihelminthic PKs [381]. Produced by Streptomyces avermetilis, these
molecules were specifically antiparasitic and did not display activity against bacteria or fungi [381, 382]. The biosynthetic cluster, made up of nine genes spanning
over 90 000 bp, was elucidated over 20 years later [383], with AveR as a putative
regulatory protein [381, 383, 384]. Soon after the discovery of avermectin, Pfizer
produced a more efficient analog, doramectin, through targeted mutagenesis of
natural precursor pathways and exogenous feeding of alternative ones [385, 386].
A collection of analogs with different C25 substituents and different levels of activity were obtained by feeding carboxylic acids or their precursors to a producing
mutant strain [386]. Furthermore, Wang and colleagues [387] were able to produce
doramectin without the need to feed these exogenous precursors by modifying
the avermectin biosynthetic pathway and replacing its loading module with the
cyclohexanecarboxylic (CHC) module from the PK phoslactomycin.
Ivermectin, another antiparasitic drug used in veterinary medicine, is also an
analog of avermectin [388, 389]. Ivermectin is now a semisynthetic drug, resulting from the chemical reduction of bonds in the intermediate avermectin B1,
produced by Streptomyces avermitilis [385, 388]. Zhang et al. [388] was able to
fully synthesize ivermectin in S. avermitilis by synthetically engineering the avermectin pathway through the replacement of two domains from one of the modules
from the avermectin PKS. More precisely, the dehydratase and ketoreductase were
replaced by the corresponding domains of the pikromycin PKS from a strain of
Streptomyces venezuelae. Despite the resulting low yield, this represented a major
advancement and allowed for an easier route to lower cost industrial production
of ivermectin [388, 389].
7.7.2.6 Production of Doxorubicin/Daunorubicin
In addition to antibiotics, the production of other highly important molecules
can be optimized through genetic engineering. Initially valued for their antibiotic
properties, daunorubicin (DNR) and doxorubicin (DXR) are two related anthracyclines produced by Streptomyces peucetius strains that became invaluable
chemotherapeutic agents. Currently, DXR is partly synthesized from the more
abundant DNR. This is due to the fact that strains capable of producing high
levels of DNR, but not DXR, have been found. DNR can be converted to DXR
by hydroxylation of its C14 [390–392]. The high demand for DXR, exceeding
225 kg year−1 in 1999, has led researchers to synthetically engineer strains for
increased production.
Attempts have been made to increase the yield of DXR produced by the native
host. Since production can lead to host toxicity and thus put a ceiling on maximal
149
150
7 Synthetic Biology and Therapies for Infectious Diseases
yield, one of these attempts involved overexpressing the DXR resistance genes
drrABC under the control of a stronger constitutive promoter, ermE*p [393]. A
fivefold increase in the production of DXR was observed when drrC was overexpressed and an approximately twofold increase when drrAB and drrABC were
overexpressed. In all cases, the increase was only observed at 72 and 84 h of incubation [393].
Overexpression of dnmT, required for daunosamine moiety biosynthesis, also
led to an increase of almost 10-fold of DNR production. The authors found that
the inactivation of drnH (which encodes a glycosyl transferase) resulted in a DNR
increase of 8.5-fold, with DXR itself being overproduced as well. These mutated
genes are located within the DXR cluster [394].
Other studies have found that disrupting some of the genes in the pathway – dnrU and dnrX – in S. peucetius resulted in a 36% increase in DXR yield,
with the triple mutant of dnrX (unknown function), dnrU (putative ketoreductase), and dnrH producing 86% more compound than the wild-type strain [395].
To this date, and to our best knowledge, the maximum reported increase in DXR
production was 11.4-fold by Lomovkaya et al. in 1999 [395]. Further work is
therefore needed to increase DXR production levels, as this molecule is currently
in shortage [396].
7.7.2.7 Development of Hosts for the Expression of Nonribosomal Peptides and
Polyketides
For many NRP and PK compounds, heterologous hosts have been explored when
the natural hosts are difficult to grow due to nutrient requirements or growth
kinetics and because the natural yields are low. One of the first strains to be engineered for the biosynthesis of heterologous natural products was Streptomyces
coelicolor CH999. Seeking to build a strain for expression of foreign PKs in considerable amounts, McDaniel et al. [397] removed the antibiotic actinorodin cluster
from S. coelicolor A3(2) by homologous recombination. This strain also carried a
mutation that disabled the synthesis of the pink antibiotic prodigiosin, and was
used to express a multitude of antibiotics, including actinorhodin, granaticin, and
tetracenomycin, from extrachromosomal material introduced into the new host
strain. Yet, CH999 only accepts non-methylated DNA and exhibits low transformation efficiencies. Kosan Biosciences then built S. lividans strains (K4-114,
K4-115) that can be transformed much more efficiently and with methylated DNA
and also lack the actinorhodin cluster, similarly to CH999. These three strains,
K4-114, K4-115, and CH999, have been tested for the capacity to produce the erythromycin precursor 6-deoxyerythronolide B (6dEB). They displayed equal efficiency in its production [339].
S. avermitilis, another soil isolate, is known for its capacity to produce the
PK antibiotic avermectin [346, 398]. In order to engineer S. avermitilis into a
versatile and ideal host for the expression of heterologous PK clusters, several
rounds of gene deletions were performed. The initial strategy used was deletion
by homologous recombination. However, this strategy turned out to be inefficient
and resulted in irregular deletions, although two useful mutants were recovered
7.7
Natural Products Discovery and Engineering
in strain SUKA2. The authors then decided to follow a different approach in which
they used the Cre-Lox recombination system to remove a 1.5 Mb section of the
9 Mb chromosome of S. avermitilis to create a new strain named SUKA3. With
the deletion of that large region, both the avermectin and filipin biosynthetic
clusters were removed. Additional deletion strains were built upon SUKA2
and SUKA3 by removing the oligomycin synthesis genes, named SUKA4 and
SUKA5, respectively. Further deletions removed terpene biosynthesis (geosmin,
neopentalenolactone, and carotenoids) as well as 78% of the transposases found
in the chromosome. This mutant, with a chromosome reduced to about 82% of
its original size, is a useful host for the expression of clusters from both culturable
and unculturable bacteria. It is also well suited for combinatorial synthesis
approaches [399].
Gomez-Escribano and Bibb [400] took the genetic engineering of Streptomyces
further. The availability of precursors and the capacity to withstand the production of potentially toxic molecules are major goals for engineering of Streptomyces
spp. as heterologous hosts. These authors engineered S. coelicolor M145 by removing its capacity to produce antibiotics. This involved the elimination of the cluster responsible for the biosynthesis of actinorhodin (blue pigmented antibiotic),
prodigiosin (red pigmented antibiotic), cryptic Type I polyketide (CPK), and the
NRPS-derived calcium-dependent antibiotic (CDA). The goal of this work was to
focus the host’s metabolic machinery primarily on the synthesis of heterologous
pathways introduced into this strain. In order to pleiotropically achieve higher
levels of metabolite production, an additional step was undertaken: point mutations were introduced in the rpoB (strain M1152) and both rpoB and rpsL genes
(strain M1154), as it had been previously shown that these mutations resulted
in an increased level of antibiotic production without affecting bacterial growth
[401, 402]. These engineered strains not only facilitate the overexpression of heterologous natural products but also simplify the identification and study of these
products. This latter effect is due to the simplicity of the HPLC chromatogram,
which results from the removal of the major metabolites that are natively produced by this strain. This empowers the use of chromatography, coupled with MS,
for the identification of novel peaks and molecules [400].
The well-studied host E. coli has also been used for the heterologous expression of natural products. However, there are additional challenges that must be
overcome when trying to express natural products in E. coli. For example, heterologous pathways often need precursors that are not all naturally available in E. coli.
Moreover, the biosynthesis of PKs requires proper folding and post-translational
modifications of the biosynthetic enzymes. Therefore, when Pfeifer et al. [345]
sought to synthesize complex PKs in E. coli, genetic engineering of this host was
required. This involved the introduction of pathways for the production of certain precursors in the proper stoichiometries as well as the expression of genes
involved in the post-translational modification of heterologous enzymes. In order
to engineer E. coli to synthesize 6dEB, the DEBS genes were introduced, as well
as a phosphopantetheinyl transferase from B. subtilis (for activation of the carrier proteins) and overexpression of genes prpE (for conversion of propionate into
151
152
7 Synthetic Biology and Therapies for Infectious Diseases
propionyl-CoA) and birA (for overexpression of propionyl CoA carboxylase). This
resulted in high-level synthesis of 6dEB from propionate in E. coli instead of the
native Saccharoplyspora erythraea.
7.7.3
Generation of Novel Molecules by Rational Reprogramming
Given the highly modular nature of NRPSs and PKSs, several attempts have been
made at generating new molecules from the rational combination of genes or parts
thereof [337]. Exchanging modules that determine the unit to be incorporated
into the product can lead to the selection and loading of a different precursor. For
example, and as briefly mentioned above, in the case of the antibiotic avermectin,
switching the isobutyryl-CoA-uploading module with a CHC-uploading unit was
successful. This single switch resulted in a novel molecule – doramectin – being
produced [337, 385, 387]. Notably, the replacement of the unit-uploading machinery is independent of the unit biosynthesis machinery, meaning that although a
given cluster can upload a new molecule, it does not necessarily mean that it can
make it. Therefore, the genes involved in the biosynthesis of the unit CHC acid
had to be cloned into the doramectin biosynthetic host as well [337, 387].
Attempting to engineer natural pathways can often result in failure; for
example, when downstream processing domains do not recognize the precursor,
no molecule can be successfully assembled [356]. Thus, the capacity to successfully reengineer biosynthetic pathways is directly related to the understanding
of the pathways of interest or the ability to rapidly construct and test many
pathway variants. Such knowledge is not merely at the level of DNA and protein
sequences, but involves understanding how the proteins interact with each
other, how specific the biosynthetic enzymes are to a certain substrate, and the
stoichiometry of the pathway itself [320]. The first successful crystallization of
NRPSs and PKSs, as well as NMR studies, has contributed to understanding how
these enzymes cooperate and are organized in particular pathways.
Reprogramming specialized metabolite production can be carried out by engineering the specific pathway of interest as well as global regulators (Figure 7.12)
[403]. A regulator capable of controlling metabolite production across all species
has yet to be discovered, but several regulators have been found to be involved in
the expression control of a variety of molecules. Such is the case for DasR, PhoP,
and ArpA, pleiotropic regulators that control compound production in several
species and are useful targets for improving the titers of molecules produced
by unknown pathways. For example, regulators such as SARPs (Streptomyces
antibiotic regulatory proteins) act as transcriptional activators for a given
pathway (Figure 7.12a and b), whereas others, such as TetR, are transcriptional
repressors (Figure 7.12c and d). In addition, some families of proteins, such as
MarR and LysR, include both repressors and activators and thus, their function
cannot be predicted on the basis of family alone. A two- to threefold increase in
natural product titers has been achieved in Streptomyces by the overexpression
7.7
Natural Products Discovery and Engineering
(a)
(b)
(c)
(d)
Figure 7.12 (a) In the absence of molecule
production, (b) cloning genes that encode
pathway activator proteins can result in the
transcription of the regulated pathway genes
and thus, a metabolic product. In the same
way, (c) if a repressor is disabling the biosynthesis of a molecule, (d) disruption of the
gene encoding this repressor (inverted triangle) can lead to the resumption of transcription and the production of a molecule.
153
154
7 Synthetic Biology and Therapies for Infectious Diseases
of pathway activators. For example, the transcriptional activator SARP was overexpressed to enhance the production of the drug C-1027. In addition, knocking
out the transcriptional repressor encoded by ptmR1 increased the production of
the antibiotics platensimycin and platencin in Streptomyces platensis MA7327 by
about 100-fold [340] (Figure 7.12).
Reprogramming biosynthetic pathways has been useful in the case of the antibiotic daptomycin. As described earlier, this antibiotic is approved for the treatment
of skin infections, bacteremia, and right-side endocarditis by Gram-positive bacteria, such as S. aureus. Yet, it has failed the “at least as good as” criteria, also known
as the non-inferiority criteria, in a clinical trial to treat community-acquired pneumonia (CAP) [367]. It is thought that this failure was due to the sequestration of
this molecule in pulmonary surfactants, which help gas exchange in the alveolar space. NRPSs involved in the assembly of daptomycin were expressed from
several different chromosomal loci that were determined experimentally to result
in higher compound yield. The reengineering of the biosynthetic pathway also
allowed for modifications of the core peptide, which were not possible to achieve
synthetically [370, 404]. Hybrid lipopeptides were built by exchanging the third
unit of DptD, the NRPS for daptomycin, with that of A54145, a CDA structurally
very similar to daptomycin, together with the deletion of dptI and integration
of heterologous replacements of dptI coding for different subunits. Seventy new
lipopeptides varying in their side chains were generated, many with strong antimicrobial properties even in the presence of surfactants such as those found in the
alveolar space [346, 367, 404].
7.7.4
Engineering NRPS and PKS Domains
Instead of just combining genes, the modular nature of NRPSs and PKSs also
allows for combinatorics at the level of domains. One of the main determinants of
enzyme selectivity is the adenylation domain, which is responsible for the selection of the amino acid to be incorporated [301]. By reengineering domains, it has
also been found that the thiolation domain has no particular selectivity [405]. Similarly, the condensation domain only seems to display some selectivity at the donor
level and for the size of the molecule it accepts. This same study showed that an
elongation module has the potential to be converted into an initiation module for
peptide bond formation by simple module relocation within the gene [405].
For example, several dozen different 6-dEB analogs were generated by swapping
the adenylation, thiolation, and β-carbon processing domains for the enzymes
responsible for synthesizing 6-dEB and rapamycin, both PKs. This process was
carried out in a heterologous host and some of the generated variants had a fourfold increased affinity to the Hsp90 heat shock protein, which was chosen as a
target because it is crucial for the survival of cells under stress such as tumor
cells [346, 406]. In addition, as noted above, the semi-synthetic drug ivermectin
is produced in a strain of S. avermitilis by PKSs. By switching the dehydratase
and ketoreductase domains of one of the modules of the avermectin with the
7.7
Natural Products Discovery and Engineering
dehydratase, enoyl reductase, and ketoreductase domains from another module
involved in the synthesis of rapamycin, ivermectin was generated [388].
Furthermore, in an attempt to understand the mechanisms underlying the unidirectional growth of polyketide chains, Kapur et al. [407] engineered 6dEB synthase to repeat an elongation step by replacing the ACP domain with the ACP of an
upstream enzyme. In other words, a non-iterative PKS was converted into an iterative one, such that intermediates undergo multiple rounds of elongation within
the same module. The unidirectionality of an elongation domain depends on the
helix I of a given ACP being unable to dock onto the ketosynthase-acyltransferase
domains of the same module, such that the downstream didomain is the only
option for elongation.
When performing domain and module swaps, it is important to maintain
proper protein–protein interactions. The small peptide regions between domains
and modules, termed the interdomain and intermodule linkers respectively, have
been found to play an important role in the molecular assembly line [408, 409].
Although these linkers are not conserved at the levels of primary and secondary
structure, they seem to be responsible for the proper folding of their associated
enzymes. Studies indicate that these regions have a tendency to form defined
secondary folds, with the resulting turn being conformationally sensitive [405].
A certain degree of conservation has been found at the level of a single proline
residue, which acts as a switch to coordinate domain activity. Thus, the role of
these linkers seems to be structural, maintaining the domains in their proper
locations [409]. To facilitate the identification of linker regions, software with
specific algorithms that account for their low levels of conservation has been
developed, such as the Udwary–Merski algorithm (UMA). These algorithms can
endow synthetic biologists with the ability to predict candidate locations for the
division of genes into parts to be reassembled. Without such tools, one risks
subdividing genes into pieces that will not be functional because of nonideal
protein–protein interactions [410].
7.7.5
Activation of Cryptic Genes/Clusters
The awakening of cryptic or silent gene clusters encoding natural molecules can
gain much from the tools of synthetic biology (Figure 7.13a and b). These clusters, although seemingly capable of leading to the production of a given polyketide or NRP of a predicted structure, are not expressed at significant levels under
tested conditions [308]. One successful strategy for activating these cryptic pathways has involved the overexpression of global secondary metabolism activators,
such as laeA, in Aspergillus nidulans, which led to the expression of antimicrobial
biosynthetic clusters [305, 411–413]. However, in certain cases, the key regulators are specific to a given pathway rather than being global in nature. In such
situations, authors have successfully expressed putative pathway-specific activators (e.g., apdR) extrachromosomally under the control of inducible promoters
(Figure 7.13) [295, 305, 413].
155
156
7 Synthetic Biology and Therapies for Infectious Diseases
(a)
(b)
(c)
(d)
Figure 7.13 (a) Silent cryptic biosynthetic
pathways can be activated by (b) cloning a
promoter upstream of the operon of interest that will drive its expression. HPLC chromatograms can then reveal the molecule
that a given cluster is responsible for synthesizing. In the same way, (c) when a certain
product is observed but whose biosynthetic
origin is unknown, (d) interfering with the
transcription of genes (e.g., by transposon
mutagenesis, shown as an inverted triangle)
permits connection of specific biosynthetic
genes with specific product molecules with
the aid of chromatograms.
7.8
Summary
7.7.6
Mutasynthesis as a Source of Novel Analogs
Structural diversity in NRP and PK molecules can also be introduced by disrupting the endogenous production of precursors followed by exogenous feeding
of alternative precursors to generate molecular analogs (mutasynthons), if the
biosynthetic enzymes are promiscuous enough to accept different precursors
[346]. This approach, termed mutasynthesis, has been used to generate new
molecules, such as aminocoumarin antibiotics. For example, the selection of a
less selective amide synthase (ClocL instead of CouL and SimL), responsible for
linking the aminocoimarin moiety to the acyl group, allowed for more efficient
generation of mutasynthons by expanding the diversity of precursors that could
be utilized. This resulted in the generation of 12 new aminocoumarin-derived
molecules, displaying both antibacterial and gyrase-inhibitory activities that were
comparable or superior to that of the aminocoumarin novobiocin [414].
An improved version of the Hsp90 inhibitor geldanamycin was obtained by
using this same approach. The precursor’s biosynthetic pathway was removed and
a heterologous precursor was fed into the culture. The novel analog was a chlorosubstituted nonbenzoquinone and exhibited improved antibiotic properties [346,
415]. Analogs of the immunosuppressant rapamycin [346, 416] and the antibiotic balhimycin are also among many of the other molecules generated through
mutasynthesis [346, 417].
7.8
Summary
Here, we have provided an overview of the fields of vaccinology, alternatives
to antibiotics for the treatment of bacterial diseases, and natural products
discovery. There has been a tremendous amount of research in these fields,
but significant limitations still exist. We envision that these challenges can be
tackled with advances across interdisciplinary areas, including synthetic biology.
In particular, we have highlighted examples where the genetic engineering of
pathways, proteins, phages, and organisms has led to new and/or optimized
capabilities to treat infectious diseases. Synthetic biology expands upon the
power of genetic engineering by providing tools for more rapid, precise, and
high-throughput biological design. These tools include cheaper and faster DNA
synthesis techniques, more scalable DNA assembly and genome engineering
strategies, and complex gene regulatory systems with tunable functionalities. We
envision that these technologies will continue to advance the development of
therapeutics for infectious diseases.
Acknowledgments
The authors thank members of the Lu lab for comments. All authors are listed in
alphabetical order, except for the last author (T.K.L.). We acknowledge support
157
158
7 Synthetic Biology and Therapies for Infectious Diseases
from the National Science Foundation (1124247), National Institutes of Health
(1DP2OD008435 and 1P50GM098792), the Ellison Foundation, the Office of
Naval Research (N00014-11-1-0725, N00014-11-1-0687, and N00014-13-10424), the Army Research Office (W911NF-11-1-0281), the Defense Advanced
Research Projects Agency, and the United States Presidential Early Career
Award for Scientists and Engineers. H.A. acknowledges support from a JSPS
Postdoctoral Fellowship for Research Abroad and the Naito Foundation.
References
1. Spellberg, B. (2008) Dr. William H.
2.
3.
4.
5.
6.
7.
8.
9.
10.
Stewart: mistaken or maligned? Clin.
Infect. Dis., 47, 294.
World Health Organization (2008) The
Global Burden of Disease: 2004 Update.
Fauci, A.S. and Morens, D.M. (2012)
The perpetual challenge of infectious diseases. N. Engl. J. Med., 366,
454–461.
World Health Organization (2011)
Global Summary of the AIDS Epidemic.
Morens, D.M., Folkers, G.K., and
Fauci, A.S. (2004) The challenge of
emerging and re-emerging infectious
diseases. Nature, 430, 242–249.
Beisel, C.E. and Morens, D.M. (2004)
Variant Creutzfeldt-Jakob disease and
the acquired and transmissible spongiform encephalopathies. Clin. Infect.
Dis., 38, 697–704.
Casalone, C., Zanusso, G., Acutis, P.,
Ferrari, S., Capucci, L., Tagliavini, F.
et al. (2004) Identification of a second bovine amyloidotic spongiform
encephalopathy: molecular similarities
with sporadic Creutzfeldt-Jakob disease. Proc. Natl. Acad. Sci. U.S.A., 101,
3065–3070.
Chang, Y., Cesarman, E., Pessin, M.S.,
Lee, F., Culpepper, J., Knowles, D.M.
et al. (1994) Identification of
herpesvirus-like DNA sequences in
AIDS-associated Kaposi’s sarcoma.
Science, 266, 1865–1869.
Sanders, M.K. and Peura, D.A. (2002)
Helicobacter pylori-associated diseases.
Curr. Gastroenterol. Rep., 4, 448–454.
Breeveld, F.J., Vreden, S.G., and
Grobusch, M.P. (2012) History of
11.
12.
13.
14.
15.
16.
17.
18.
19.
malaria research and its contribution to the malaria control success in
Suriname: a review. Malar. J., 11, 95.
Wellems, T.E. and Miller, L.H. (2003)
Two worlds of malaria. N. Engl. J. Med.,
349, 1496–1498.
Fischbach, M.A. and Walsh, C.T. (2009)
Antibiotics for emerging pathogens.
Science, 325, 1089–1093.
Centers for Disease Control and Prevention (2013) Antibiotic Resistance
Threats in the United States, 2013.
Falagas, M.E., Bliziotis, I.A.,
Kasiakou, S.K., Samonis, G.,
Athanassopoulou, P., and
Michalopoulos, A. (2005) Outcome
of infections due to pandrug-resistant
(PDR) Gram-negative bacteria. BMC
Infect. Dis., 5, 24.
Dorman, S.E. and Chaisson, R.E. (2007)
From magic bullets back to the magic
mountain: the rise of extensively drugresistant tuberculosis. Nat. Med., 13,
295–298.
World Health Organization (2011)
Global Tuberculosis Report 2012.
World Health Organization (2006)
WHO Global Task Force Outlines Measures to Combat XDR-TB Worldwide.
Kim, D.H., Kim, H.J., Park, S.K.,
Kong, S.J., Kim, Y.S., Kim, T.H. et al.
(2008) Treatment outcomes and
long-term survival in patients with
extensively drug-resistant tuberculosis.
Am. J. Respir. Crit. Care Med., 178,
1075–1082.
Boucher, H.W., Talbot, G.H.,
Benjamin, D.K. Jr., Bradley, J.,
Guidos, R.J., Jones, R.N. et al. (2013)
10 x ′ 20 Progress–development of new
drugs active against Gram-negative
bacilli: an update from the Infectious
References
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
Diseases Society of America. Clin.
Infect. Dis., 56, 1685–1694.
Spellberg, B., Powers, J.H., Brass, E.P.,
Miller, L.G., and Edwards, J.E. Jr.,
(2004) Trends in antimicrobial
drug development: implications
for the future. Clin. Infect. Dis., 38,
1279–1286.
Roemer, T. and Boone, C. (2013)
Systems-level antimicrobial drug and
drug synergy discovery. Nat. Chem.
Biol., 9, 222–231.
Boucher, H.W., Talbot, G.H.,
Bradley, J.S., Edwards, J.E., Gilbert, D.,
Rice, L.B. et al. (2009) Bad bugs, no
drugs: no ESKAPE! An update from the
Infectious Diseases Society of America.
Clin. Infect. Dis., 48, 1–12.
Cheng, A.A. and Lu, T.K. (2012) Synthetic biology: an emerging engineering
discipline. Annu. Rev. Biomed. Eng., 14,
155–178.
Khalil, A.S. and Collins, J.J. (2010) Synthetic biology: applications come of age.
Nat. Rev. Genet., 11, 367–379.
Lu, T.K. (2010) Engineering scalable
biological systems. Bioeng. Bugs, 1,
378–384.
Ruder, W.C., Lu, T., and Collins, J.J.
(2011) Synthetic biology moving into
the clinic. Science, 333, 1248–1252.
Lu, T.K., Bowers, J., and Koeris, M.S.
(2013) Advancing bacteriophagebased microbial diagnostics with
synthetic biology. Trends Biotechnol.,
31, 325–327.
Lu, T.K. and Koeris, M.S. (2011) The
next generation of bacteriophage
therapy. Curr. Opin. Microbiol., 14,
524–531.
Clemens, J. (2011) Evaluation of vaccines against enteric infections: a
clinical and public health research
agenda for developing countries. Philos.
Trans. R. Soc. Lond. B Biol. Sci., 366,
2799–2805.
Walker, R.I., Steele, D., Aguado, T.,
and Ad Hoc, E.T.E.C. (2007) Analysis
of strategies to successfully vaccinate
infants in developing countries against
enterotoxigenic E. coli (ETEC) disease.
Vaccine, 25, 2545–2566.
31. Brumbaugh, A.R. and Mobley, H.L.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
(2012) Preventing urinary tract infection: progress toward an effective
Escherichia coli vaccine. Expert Rev.
Vaccines, 11, 663–676.
Girard, M.P., Reed, Z.H., Friede, M.,
and Kieny, M.P. (2007) A review of
human vaccine research and development: malaria. Vaccine, 25, 1567–1580.
Girard, M.P., Osmanov, S.K., and
Kieny, M.P. (2006) A review of vaccine
research and development: the human
immunodeficiency virus (HIV). Vaccine,
24, 4062–4081.
Girard, M.P., Steele, D., Chaignat, C.L.,
and Kieny, M.P. (2006) A review of
vaccine research and development:
human enteric infections. Vaccine, 24,
2732–2750.
Heppner, D.G. (2013) The malaria vaccine – status quo 2013. Travel Med.
Infect. Dis., 11, 2–7.
Schiffner, T., Sattentau, Q.J., and
Dorrell, L. (2013) Development of
prophylactic vaccines against HIV-1.
Retrovirology, 10, 72.
Pulendran, B. and Ahmed, R. (2011)
Immunological mechanisms of vaccination. Nat. Immunol., 12, 509–517.
Bonanni, P. and Santos, J.I. (2011) Vaccine evolution. Perspect. Vaccinol., 1,
1–24.
Lauring, A.S., Jones, J.O., and
Andino, R. (2010) Rationalizing the
development of live attenuated virus
vaccines. Nat. Biotechnol., 28, 573–579.
Svenson, S., Kallenius, G.,
Pawlowski, A., and Hamasur, B. (2010)
Towards new tuberculosis vaccines.
Hum. Vaccin., 6, 309–317.
Calmette, A. (1931) Preventive vaccination against tuberculosis with BCG.
Proc. R. Soc. Med., 24, 1481–1490.
Kaufmann, S.H. and Gengenbacher, M.
(2012) Recombinant live vaccine candidates against tuberculosis. Curr. Opin.
Biotechnol., 23, 900–907.
Grode, L., Ganoza, C.A., Brohm, C.,
Weiner, J. III, Eisele, B., and
Kaufmann, S.H. (2013) Safety and
immunogenicity of the recombinant
BCG vaccine VPM1002 in a phase 1
open-label randomized clinical trial.
Vaccine, 31, 1340–1348.
159
160
7 Synthetic Biology and Therapies for Infectious Diseases
44. Hoft, D.F., Blazevic, A., Abate, G.,
45.
46.
47.
48.
49.
50.
51.
52.
53.
Hanekom, W.A., Kaplan, G., Soler, J.H.
et al. (2008) A new recombinant bacille
Calmette-Guerin vaccine safely induces
significantly enhanced tuberculosisspecific immunity in human volunteers.
J. Infect. Dis., 198, 1491–1501.
Horwitz, M.A., Harth, G., Dillon, B.J.,
and Maslesa-Galic, S. (2006) A novel
live recombinant mycobacterial vaccine
against bovine tuberculosis more potent
than BCG. Vaccine, 24, 1593–1600.
Horwitz, M.A. (2005) Recombinant
BCG expressing Mycobacterium tuberculosis major extracellular proteins.
Microbes Infect., 7, 947–954.
Tullius, M.V., Harth, G.,
Maslesa-Galic, S., Dillon, B.J., and
Horwitz, M.A. (2008) A ReplicationLimited Recombinant Mycobacterium
bovis BCG vaccine against tuberculosis
designed for human immunodeficiency
virus-positive persons is safer and more
efficacious than BCG. Infect. Immun.,
76, 5200–5214.
Sun, R., Skeiky, Y.A., Izzo, A.,
Dheenadhayalan, V., Imam, Z.,
Penn, E. et al. (2009) Novel recombinant BCG expressing perfringolysin
O and the over-expression of key
immunodominant antigens; pre-clinical
characterization, safety and protection
against challenge with Mycobacterium
tuberculosis. Vaccine, 27, 4412–4423.
Zaki, S.A. and Karande, S. (2011)
Multidrug-resistant typhoid fever:
a review. J. Infect. Dev. Ctries., 5,
324–337.
Gonzalez-Escobedo, G., Marshall, J.M.,
and Gunn, J.S. (2011) Chronic and
acute infection of the gall bladder by
Salmonella Typhi: understanding the
carrier state. Nat. Rev. Microbiol., 9,
9–14.
Martin, L.B. (2012) Vaccines for
typhoid fever and other salmonelloses.
Curr. Opin. Infect. Dis., 25, 489–499.
Ehrbar, K. and Hardt, W.D. (2005)
Bacteriophage-encoded type III effectors in Salmonella enterica subspecies
1 serovar Typhimurium. Infect. Genet.
Evol., 5, 1–9.
Srikanth, C.V., Mercado-Lubo, R.,
Hallstrom, K., and McCormick, B.A.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
(2011) Salmonella effector proteins and
host-cell responses. Cell. Mol. Life Sci.,
68, 3687–3697.
Valdez, Y., Ferreira, R.B., and
Finlay, B.B. (2009) Molecular mechanisms of Salmonella virulence and
host resistance. Curr. Top. Microbiol.
Immunol., 337, 93–127.
Bakowski, M.A., Braun, V., and
Brumell, J.H. (2008) Salmonellacontaining vacuoles: directing traffic
and nesting to grow. Traffic, 9,
2022–2031.
Tacket, C.O., Hone, D.M., Curtiss, R.
III, Kelly, S.M., Losonsky, G., Guers, L.
et al. (1992) Comparison of the safety
and immunogenicity of delta aroC delta
aroD and delta cya delta crp Salmonella
typhi strains in adult volunteers. Infect.
Immun., 60, 536–541.
Schmitt, C.K., Darnell, S.C., and
O’Brien, A.D. (1996) The attenuated phenotype of a Salmonella
typhimurium flgM mutant is related to
expression of FliC flagellin. J. Bacteriol.,
178, 2911–2915.
Levine, M.M., Galen, J., Barry, E.,
Noriega, F., Chatfield, S., Sztein, M.
et al. (1996) Attenuated Salmonella as
live oral vaccines against typhoid fever
and as live vectors. J. Biotechnol., 44,
193–196.
Cardenas, L. and Clements, J.D. (1992)
Oral immunization using live attenuated Salmonella spp. as carriers of
foreign antigens. Clin. Microbiol. Rev.,
5, 328–342.
Hall, H.K. and Foster, J.W. (1996)
The role of fur in the acid tolerance
response of Salmonella typhimurium
is physiologically and genetically separable from its role in iron acquisition.
J. Bacteriol., 178, 5683–5691.
Foster, J.W. (1995) Low pH adaptation and the acid tolerance response
of Salmonella typhimurium. Crit. Rev.
Microbiol., 21, 215–237.
Kato, A., Groisman, E.A., and I.
Howard Hughes Medical (2008) The
PhoQ/PhoP regulatory network of
Salmonella enterica. Adv. Exp. Med.
Biol., 631, 7–21.
Sebkova, A., Karasova, D.,
Crhanova, M., Budinska, E., and
References
64.
65.
66.
67.
68.
69.
70.
71.
72.
73.
Rychlik, I. (2008) aro mutations in
Salmonella enterica cause defects in cell
wall and outer membrane integrity. J.
Bacteriol., 190, 3155–3160.
Dong, T. and Schellhorn, H.E. (2010)
Role of RpoS in virulence of pathogens.
Infect. Immun., 78, 887–897.
Curtiss, R. III, Galan, J.E.,
Nakayama, K., and Kelly, S.M. (1990)
Stabilization of recombinant avirulent
vaccine strains in vivo. Res. Microbiol.,
141, 797–805.
Galan, J.E., Nakayama, K., and
Curtiss, R. III, (1990) Cloning and
characterization of the asd gene of
Salmonella typhimurium: use in stable
maintenance of recombinant plasmids
in Salmonella vaccine strains. Gene, 94,
29–35.
Galen, J.E. and Curtiss, R. III, (2014)
The delicate balance in genetically
engineering live vaccines. Vaccine, in
press.
Curtiss, R. III, Xin, W., Li, Y., Kong, W.,
Wanda, S.Y., Gunn, B. et al. (2010)
New technologies in using recombinant
attenuated Salmonella vaccine vectors.
Crit. Rev. Immunol., 30, 255–270.
Kong, W., Brovold, M.,
Koeneman, B.A., Clark-Curtiss, J.,
and Curtiss, R. III, (2012) Turning
self-destructing Salmonella into a universal DNA vaccine delivery platform.
Proc. Natl. Acad. Sci. U.S.A., 109,
19414–19419.
Wang, S., Kong, Q., and Curtiss, R. III,
(2013) New technologies in developing
recombinant attenuated Salmonella
vaccine vectors. Microb. Pathog., 58,
17–28.
Miller, S.I. (1991) Phop
Phoq – macrophage-specific modulators of Salmonella virulence. Mol.
Microbiol., 5, 2073–2078.
Nickerson, C.A. and Curtiss, R.
III, (1997) Role of sigma factor
RpoS in initial stages of Salmonella
typhimurium infection. Infect. Immun.,
65, 1814–1823.
Lee, I.S., Lin, J., Hall, H.K., Bearson, B.,
and Foster, J.W. (1995) The stationaryphase sigma factor sigma S (RpoS) is
required for a sustained acid tolerance response in virulent Salmonella
74.
75.
76.
77.
78.
79.
80.
typhimurium. Mol. Microbiol., 17,
155–167.
Gunn, B.M., Wanda, S.Y., Burshell, D.,
Wang, C., and Curtiss, R. III, (2010)
Construction of recombinant attenuated Salmonella enterica serovar
typhimurium vaccine vector strains
for safety in newborn and infant mice.
Clin. Vaccine Immunol., 17, 354–362.
Curtiss, R. III, Wanda, S.Y.,
Gunn, B.M., Zhang, X., Tinge, S.A.,
Ananthnarayan, V. et al. (2009)
Salmonella enterica serovar
typhimurium strains with regulated
delayed attenuation in vivo. Infect.
Immun., 77, 1071–1082.
Li, Y., Wang, S., Scarpellini, G.,
Gunn, B., Xin, W., Wanda, S.Y. et al.
(2009) Evaluation of new generation Salmonella enterica serovar
Typhimurium vaccines with regulated delayed attenuation to induce
immune responses against PspA. Proc.
Natl. Acad. Sci. U.S.A., 106, 593–598.
Wang, S., Li, Y., Scarpellini, G.,
Kong, W., Shi, H., Baek, C.H. et al.
(2010) Salmonella vaccine vectors displaying delayed antigen synthesis in
vivo to enhance immunogenicity. Infect.
Immun., 78, 3969–3980.
Wang, S., Li, Y., Shi, H., Sun, W.,
Roland, K.L., and Curtiss, R. III,
(2011) Comparison of a regulated
delayed antigen synthesis system with
in vivo-inducible promoters for antigen
delivery by live attenuated Salmonella
vaccines. Infect. Immun., 79, 937–949.
Juarez-Rodriguez, M.D., Yang, J.,
Kader, R., Alamuri, P., Curtiss, R.
III, and Clark-Curtiss, J.E. (2012) Live
attenuated Salmonella vaccines displaying regulated delayed lysis and delayed
antigen synthesis to confer protection
against Mycobacterium tuberculosis.
Infect. Immun., 80, 815–831.
Kong, W., Wanda, S.Y., Zhang, X.,
Bollen, W., Tinge, S.A., Roland, K.L.
et al. (2008) Regulated programmed
lysis of recombinant Salmonella in host
tissues to release protective antigens
and confer biological containment.
Proc. Natl. Acad. Sci. U.S.A., 105,
9361–9366.
161
162
7 Synthetic Biology and Therapies for Infectious Diseases
81. Oliveira, A.F., Cardoso, S.A.,
82.
83.
84.
85.
86.
87.
88.
Almeida, F.B., de Oliveira, L.L.,
Pitondo-Silva, A., Soares, S.G. et al.
(2012) Oral immunization with attenuated Salmonella vaccine expressing
Escherichia coli O157:H7 intimin
gamma triggers both systemic and
mucosal humoral immunity in mice.
Microbiol. Immunol., 56, 513–522.
Londono, L.P., Chatfield, S.,
Tindle, R.W., Herd, K., Gao, X.M.,
Frazer, I. et al. (1996) Immunisation
of mice using Salmonella typhimurium
expressing human papillomavirus type
16 E7 epitopes inserted into hepatitis B virus core antigen. Vaccine, 14,
545–552.
Chorobik, P. and Marcinkiewicz, J.
(2011) Therapeutic vaccines based on
genetically modified Salmonella: a novel
strategy in cancer immunotherapy. Pol.
Arch. Med. Wewn., 121, 461–466.
Xiong, G., Husseiny, M.I., Song, L.,
Erdreich-Epstein, A., Shackleford, G.M.,
Seeger, R.C. et al. (2010) Novel cancer
vaccine based on genes of Salmonella
pathogenicity island 2. Int. J. Cancer,
126, 2622–2634.
Nishikawa, H., Sato, E., Briones, G.,
Chen, L.M., Matsuo, M., Nagata, Y.
et al. (2006) In vivo antigen delivery
by a Salmonella typhimurium type
III secretion system for therapeutic
cancer vaccines. J. Clin. Invest., 116,
1946–1954.
Osorio, M., Wu, Y., Singh, S.,
Merkel, T.J., Bhattacharyya, S.,
Blake, M.S. et al. (2009) Anthrax protective antigen delivered by Salmonella
enterica serovar Typhi Ty21a protects
mice from a lethal anthrax spore challenge. Infect. Immun., 77, 1475–1482.
Liu, D.S., Hu, S.J., Zhou, N.J., Xie, Y.,
and Cao, J. (2011) Construction and
characterization of recombinant
attenuated Salmonella typhimurium
expressing the babA2/ureI fusion
gene of Helicobacter pylori. Clin. Res.
Hepatol. Gastroenterol., 35, 655–660.
Liu, K.Y., Shi, Y., Luo, P., Yu, S.,
Chen, L., Zhao, Z. et al. (2011)
Therapeutic efficacy of oral immunization with attenuated Salmonella
typhimurium expressing Helicobacter
89.
90.
91.
92.
93.
94.
95.
pylori CagA, VacA and UreB fusion
proteins in mice model. Vaccine, 29,
6679–6685.
Branger, C.G., Fetherston, J.D.,
Perry, R.D., and Curtiss, R. III, (2007)
Oral vaccination with different antigens
from Yersinia pestis KIM delivered by
live attenuated Salmonella typhimurium
elicits a protective immune response
against plague. Adv. Exp. Med. Biol.,
603, 387–399.
Branger, C.G., Sun, W.,
Torres-Escobar, A., Perry, R.,
Roland, K.L., Fetherston, J. et al. (2010)
Evaluation of Psn, HmuR and a modified LcrV protein delivered to mice
by live attenuated Salmonella as a vaccine against bubonic and pneumonic
Yersinia pestis challenge. Vaccine, 29,
274–282.
Kulkarni, R.R., Parreira, V.R., Jiang, Y.F.,
and Prescott, J.F. (2010) A live oral
recombinant Salmonella enterica
serovar typhimurium vaccine expressing Clostridium perfringens antigens
confers protection against necrotic
enteritis in broiler chickens. Clin.
Vaccine Immunol., 17, 205–214.
Kulkarni, R.R., Parreira, V.R., Sharif, S.,
and Prescott, J.F. (2008) Oral immunization of broiler chickens against
necrotic enteritis with an attenuated
Salmonella vaccine vector expressing Clostridium perfringens antigens.
Vaccine, 26, 4194–4203.
Zekarias, B., Mo, H., and Curtiss, R.
III, (2008) Recombinant attenuated Salmonella enterica serovar
typhimurium expressing the carboxyterminal domain of alpha toxin from
Clostridium perfringens induces protective responses against necrotic enteritis
in chickens. Clin. Vaccine Immunol.,
15, 805–816.
Ward, S.J., Douce, G., Figueiredo, D.,
Dougan, G., and Wren, B.W. (1999)
Immunogenicity of a Salmonella
typhimurium aroA aroD vaccine
expressing a nontoxic domain of
Clostridium difficile toxin A. Infect.
Immun., 67, 2145–2152.
Verma, N.K., Ziegler, H.K., Wilson, M.,
Khan, M., Safley, S., Stocker, B.A. et al.
(1995) Delivery of class I and class
References
96.
97.
98.
99.
100.
101.
102.
103.
II MHC-restricted T-cell epitopes of
listeriolysin of Listeria monocytogenes
by attenuated Salmonella. Vaccine, 13,
142–150.
Dong, X., Zhang, Y.J., and Zhang, Z.
(2013) Using weakly conserved motifs
hidden in secretion signals to identify type-III effectors from bacterial
pathogen genomes. PLoS One, 8,
e56632.
Niemann, G.S., Brown, R.N.,
Mushamiri, I.T., Nguyen, N.T.,
Taiwo, R., Stufkens, A. et al.
(2013) RNA type III secretion signals that require Hfq.
J. Bacteriol., 195, 2119–2125.
Buchko, G.W., Niemann, G.,
Baker, E.S., Belov, M.E., Smith, R.D.,
Heffron, F. et al. (2010) A multipronged search for a common
structural motif in the secretion signal of Salmonella enterica serovar
Typhimurium type III effector proteins.
Mol. Biosyst., 6, 2448–2458.
Hegazy, W.A., Xu, X., Metelitsa, L.,
and Hensel, M. (2012) Evaluation of
Salmonella enterica type III secretion
system effector proteins as carriers for
heterologous vaccine antigens. Infect.
Immun., 80, 1193–1202.
Russmann, H., Shams, H., Poblete, F.,
Fu, Y., Galan, J.E., and Donis, R.O.
(1998) Delivery of epitopes by the
Salmonella type III secretion system
for vaccine development. Science, 281,
565–568.
Asensi, G.F., de Sales, N.F., Dutra, F.F.,
Feijo, D.F., Bozza, M.T., Ulrich, R.G.
et al. (2013) Oral immunization with
Lactococcus lactis secreting attenuated
recombinant staphylococcal enterotoxin B induces a protective immune
response in a murine model. Microb.
Cell Fact., 12, 32.
Ma, Y., Luo, Y., Huang, X., Song, F.,
and Liu, G. (2012) Construction of
Bifidobacterium infantis as a live oral
vaccine that expresses antigens of the
major fimbrial subunit (CfaB) and the B
subunit of heat-labile enterotoxin (LTB)
from enterotoxigenic Escherichia coli.
Microbiology, 158, 498–504.
Takata, T., Shirakawa, T., Kawasaki, Y.,
Kinoshita, S., Gotoh, A., Kano, Y.
104.
105.
106.
107.
108.
109.
110.
111.
112.
113.
et al. (2006) Genetically engineered
Bifidobacterium animalis expressing
the Salmonella flagellin gene for the
mucosal immunization in a mouse
model. J. Gene Med., 8, 1341–1346.
Mortaz, E., Adcock, I.M., Folkerts, G.,
Barnes, P.J., Paul Vos, A., and
Garssen, J. (2013) Probiotics in the
management of lung diseases. Mediators Inflamm., 2013, 751068.
Sarowska, J., Choroszy-Krol, I.,
Regulska-Ilow, B., Frej-Madrzak, M.,
and Jama-Kmiecik, A. (2013) The therapeutic effect of probiotic bacteria on
gastrointestinal diseases. Adv. Clin. Exp.
Med., 22, 759–766.
Isolauri, E., Rautava, S., and
Salminen, S. (2012) Probiotics in the
development and treatment of allergic
disease. Gastroenterol. Clin. North Am.,
41, 747–762.
Million, M., Lagier, J.C., Yahav, D., and
Paul, M. (2013) Gut bacterial microbiota and obesity. Clin. Microbiol.
Infect., 19, 305–313.
Pathangey, L., Kohler, J.J., Isoda, R.,
and Brown, T.A. (2009) Effect of
expression level on immune responses
to recombinant oral Salmonella enterica
serovar Typhimurium vaccines. Vaccine,
27, 2707–2711.
Kindsmuller, K. and Wagner, R. (2011)
Synthetic biology: impact on the design
of innovative vaccines. Hum. Vaccin., 7,
658–662.
Lou, C., Stanton, B., Chen, Y.J.,
Munsky, B., and Voigt, C.A. (2012)
Ribozyme-based insulator parts buffer
synthetic circuits from genetic context.
Nat. Biotechnol., 30, 1137–1142.
Qi, L.S., Larson, M.H., Gilbert, L.A.,
Doudna, J.A., Weissman, J.S.,
Arkin, A.P. et al. (2013) Repurposing
CRISPR as an RNA-guided platform
for sequence-specific control of gene
expression. Cell, 152, 1173–1183.
Salis, H.M., Mirsky, E.A., and
Voigt, C.A. (2009) Automated design
of synthetic ribosome binding sites
to control protein expression. Nat.
Biotechnol., 27, 946–950.
Isaacs, F.J., Carr, P.A., Wang, H.H.,
Lajoie, M.J., Sterling, B., Kraal, L.
et al. (2011) Precise manipulation of
163
164
7 Synthetic Biology and Therapies for Infectious Diseases
114.
115.
116.
117.
118.
119.
120.
121.
122.
chromosomes in vivo enables genomewide codon replacement. Science, 333,
348–353.
Wang, H.H., Isaacs, F.J., Carr, P.A.,
Sun, Z.Z., Xu, G., Forest, C.R. et al.
(2009) Programming cells by multiplex
genome engineering and accelerated
evolution. Nature, 460, 894–898.
Jiang, W., Bikard, D., Cox, D., Zhang, F.,
and Marraffini, L.A. (2013) RNAguided editing of bacterial genomes
using CRISPR-Cas systems. Nat.
Biotechnol., 31, 233–239.
Rappuoli, R. (2000) Reverse vaccinology. Curr. Opin. Microbiol., 3, 445–450.
Rouphael, N.G. and Stephens, D.S.
(2012) Neisseria meningitidis: biology, microbiology, and epidemiology.
Methods Mol. Biol., 799, 1–20.
Papaevangelou, V. and Spyridis, N.
(2012) MenACWY-TT vaccine for
active immunization against invasive
meningococcal disease. Expert Rev.
Vaccines, 11, 523–537.
Serruto, D., Bottomley, M.J., Ram, S.,
Giuliani, M.M., and Rappuoli, R. (2012)
The new multicomponent vaccine
against meningococcal serogroup B,
4CMenB: immunological, functional
and structural characterization of
the antigens. Vaccine, 30 (Suppl. 2),
B87–B97.
Tondella, M.L., Popovic, T.,
Rosenstein, N.E., Lake, D.B.,
Carlone, G.M., Mayer, L.W. et al.,
and The Active Bacterial Core Surveillance Team (2000) Distribution of
Neisseria meningitidis serogroup B
serosubtypes and serotypes circulating
in the United States. J. Clin. Microbiol.,
38, 3323–3328.
Gomez, G., Pei, J., Mwangi, W.,
Adams, L.G., Rice-Ficht, A., and
Ficht, T.A. (2013) Immunogenic and
invasive properties of Brucella melitensis 16M outer membrane protein
vaccine candidates identified via a
reverse vaccinology approach. PLoS
One, 8, e59751.
Nesta, B., Spraggon, G., Alteri, C.,
Moriel, D.G., Rosini, R., Veggi, D. et al.
(2012) FdeC, a novel broadly conserved
123.
124.
125.
126.
127.
128.
129.
130.
Escherichia coli adhesin eliciting protection against urinary tract infections.
MBio, 3, e00010.
Liu, L., Cheng, G., Wang, C., Pan, X.,
Cong, Y., Pan, Q. et al. (2009) Identification and experimental verification of
protective antigens against Streptococcus suis serotype 2 based on genome
sequence analysis. Curr. Microbiol., 58,
11–17.
Tettelin, H., Medini, D., Donati, C., and
Masignani, V. (2006) Towards a universal group B Streptococcus vaccine using
multistrain genome analysis. Expert
Rev. Vaccines, 5, 687–694.
Maione, D., Margarit, I., Rinaudo, C.D.,
Masignani, V., Mora, M., Scarselli, M.
et al. (2005) Identification of a universal Group B streptococcus vaccine by
multiple genome screen. Science, 309,
148–150.
Masignani, V., Comanducci, M.,
Giuliani, M.M., Bambini, S.,
Adu-Bobie, J., Arico, B. et al. (2003)
Vaccination against Neisseria meningitidis using three variants of the
lipoprotein GNA1870. J. Exp. Med.,
197, 789–799.
Scarselli, M., Arico, B., Brunelli, B.,
Savino, S., Di Marcello, F., Palumbo, E.
et al. (2011) Rational design of a
meningococcal antigen inducing broad
protective immunity. Sci. Transl. Med.,
3, 91ra62.
Nuccitelli, A., Cozzi, R., Gourlay, L.J.,
Donnarumma, D., Necchi, F., Norais, N.
et al. (2011) Structure-based approach
to rationally design a chimeric protein
for an effective vaccine against Group
B Streptococcus infections. Proc. Natl.
Acad. Sci. U.S.A., 108, 10278–10283.
Johnson, S., Tan, L., van der Veen, S.,
Caesar, J., Goicoechea De Jorge, E.,
Harding, R.J. et al. (2012) Design and
evaluation of meningococcal vaccines
through structure-based modification
of host and pathogen molecules. PLoS
Pathog., 8, e1002981.
D’Herelle, F. (2007) On an invisible
microbe antagonistic toward dysenteric
bacilli: brief note by Mr. F. D’Herelle,
presented by Mr. Roux. 1917. Res.
Microbiol., 158, 553–554.
References
131. D’Herelle, F. (1931) Bacteriophage as a
132.
133.
134.
135.
136.
137.
138.
139.
140.
141.
142.
143.
treatment in acute medical and surgical
infections. Bull. N. Y. Acad. Med., 7,
329–348.
Duckworth, D.H. (1976) Who discovered bacteriophage? Bacteriol. Rev., 40,
793–802.
Summers, W.C. (2012) The strange history of phage therapy. Bacteriophage, 2,
130–133.
Abedon, S.T., Kuhl, S.J., Blasdel, B.G.,
and Kutter, E.M. (2011) Phage
treatment of human infections. Bacteriophage, 1, 66–85.
Burrowes, B., Harper, D.R.,
Anderson, J., McConville, M., and
Enright, M.C. (2011) Bacteriophage
therapy: potential uses in the control of
antibiotic-resistant pathogens. Expert
Rev. Anti Infect. Ther., 9, 775–785.
Denou, E., Bruttin, A., Barretto, C.,
Ngom-Bru, C., Brussow, H., and
Zuber, S. (2009) T4 phages against
Escherichia coli diarrhea: potential and
problems. Virology, 388, 21–30.
Filippov, A.A., Sergueev, K.V., and
Nikolich, M.P. (2012) Can phage effectively treat multidrug-resistant plague?
Bacteriophage, 2, 186–189.
Gilmore, B.F. (2012) Bacteriophages as
anti-infective agents: recent developments and regulatory challenges. Expert
Rev. Anti Infect. Ther., 10, 533–535.
Hunter, P. (2012) The return of the
phage. EMBO Rep., 13, 20–23.
Loc-Carrillo, C. and Abedon, S.T.
(2011) Pros and cons of phage therapy.
Bacteriophage, 1, 111–114.
Pirnay, J.P., De Vos, D., Verbeken, G.,
Merabishvili, M., Chanishvili, N.,
Vaneechoutte, M. et al. (2011) The
phage therapy paradigm: pret-aporter or sur-mesure? Pharm. Res.,
28, 934–937.
Balakrishnan, M. and Floch, M.H.
(2012) Prebiotics, probiotics and digestive health. Curr. Opin. Clin. Nutr.
Metab. Care, 15, 580–585.
Stecher, B., Maier, L., and Hardt, W.D.
(2013) ‘Blooming’ in the gut: how dysbiosis might contribute to pathogen
evolution. Nat. Rev. Microbiol., 11,
277–284.
144. Krylov, V., Shaburova, O., Krylov, S.,
145.
146.
147.
148.
149.
150.
151.
152.
153.
154.
and Pleteneva, E. (2013) A genetic
approach to the development of
new therapeutic phages to fight
pseudomonas aeruginosa in wound
infections. Viruses, 5, 15–53.
Brussow, H. (2012) What is needed
for phage therapy to become a reality
in Western medicine? Virology, 434,
138–142.
Trojet, S.N., Caumont-Sarcos, A.,
Perrody, E., Comeau, A.M., and
Krisch, H.M. (2011) The gp38 adhesins
of the T4 superfamily: a complex
modular determinant of the phage’s
host specificity. Genome Biol. Evol., 3,
674–686.
Scholl, D., Rogers, S., Adhya, S., and
Merril, C.R. (2001) Bacteriophage K1-5
encodes two different tail fiber proteins,
allowing it to infect and replicate on
both K1 and K5 strains of Escherichia
coli. J. Virol., 75, 2509–2515.
Sandmeier, H., Iida, S., and Arber, W.
(1992) DNA inversion regions Min of
plasmid p15B and Cin of bacteriophage
P1: evolution of bacteriophage tail fiber
genes. J. Bacteriol., 174, 3936–3944.
Rakhuba, D.V., Kolomiets, E.I.,
Dey, E.S., and Novik, G.I. (2010) Bacteriophage receptors, mechanisms
of phage adsorption and penetration
into host cell. Pol. J. Microbiol., 59,
145–155.
Labrie, S.J., Samson, J.E., and
Moineau, S. (2010) Bacteriophage resistance mechanisms. Nat. Rev. Microbiol.,
8, 317–327.
Projan, S. (2004) Phage-inspired antibiotics? Nat. Biotechnol., 22, 167–168.
Schoolnik, G.K., Summers, W.C., and
Watson, J.D. (2004) Phage offer a
real alternative. Nat. Biotechnol., 22,
505–506; author reply 506-507.
Bessler, W., Freund-Molbert, E.,
Knufermann, H., Rduolph, C.,
Thurow, H., and Stirm, S. (1973) A
bacteriophage-induced depolymerase
active on Klebsiella K11 capsular
polysaccharide. Virology, 56, 134–151.
Dobbins, A.T., George, M. Jr.,
Basham, D.A., Ford, M.E., Houtz, J.M.,
Pedulla, M.L. et al. (2004) Complete
genomic sequence of the virulent
165
166
7 Synthetic Biology and Therapies for Infectious Diseases
155.
156.
157.
158.
159.
160.
161.
162.
163.
Salmonella bacteriophage SP6. J. Bacteriol., 186, 1933–1944.
Rakonjac, J., Bennett, N.J.,
Spagnuolo, J., Gagic, D., and Russel, M.
(2011) Filamentous bacteriophage: biology, phage display and nanotechnology
applications. Curr. Issues Mol. Biol., 13,
51–76.
Lindberg, A.A. (1973) Bacteriophage
receptors. Annu. Rev. Microbiol., 27,
205–241.
Heilpern, A.J. and Waldor, M.K. (2003)
pIIICTX, a predicted CTXphi minor
coat protein, can expand the host
range of coliphage fd to include Vibrio
cholerae. J. Bacteriol., 185, 1037–1044.
Iida, S. (1984) Bacteriophage P1 carries
two related sets of genes determining its host range in the invertible C
segment of its genome. Virology, 134,
421–434.
Plasterk, R.H., Kanaar, R., and van de
Putte, P. (1984) A genetic switch in
vitro: DNA inversion by Gin protein of
phage Mu. Proc. Natl. Acad. Sci. U.S.A.,
81, 2689–2692.
Doulatov, S., Hodes, A., Dai, L.,
Mandhana, N., Liu, M., Deora, R.
et al. (2004) Tropism switching in Bordetella bacteriophage defines a family
of diversity-generating retroelements.
Nature, 431, 476–481.
Leiman, P.G., Battisti, A.J.,
Bowman, V.D., Stummeyer, K.,
Muhlenhoff, M., Gerardy-Schahn, R.
et al. (2007) The structures of bacteriophages K1E and K1-5 explain
processive degradation of polysaccharide capsules and evolution of new
host specificities. J. Mol. Biol., 371,
836–849.
Scholl, D., Adhya, S., and Merril, C.R.
(2002) Bacteriophage SP6 is closely
related to phages K1-5, K5, and K1E
but encodes a tail protein very similar
to that of the distantly related P22.
J. Bacteriol., 184, 2833–2836.
Iida, S., Huber, H., Hiestand-Nauer, R.,
Meyer, J., Bickle, T.A., and Arber, W.
(1984) The bacteriophage P1 sitespecific recombinase cin: recombination
events and DNA recognition sequences.
Cold Spring Harb. Symp. Quant. Biol.,
49, 769–777.
164. Miller, E.S., Kutter, E., Mosig, G.,
165.
166.
167.
168.
169.
170.
171.
172.
Arisaka, F., Kunisawa, T., and Ruger, W.
(2003) Bacteriophage T4 genome.
Microbiol. Mol. Biol. Rev., 67, 86–156,
table of contents.
Tetart, F., Desplats, C., and
Krisch, H.M. (1998) Genome plasticity in the distal tail fiber locus of
the T-even bacteriophage: recombination between conserved motifs swaps
adhesin specificity. J. Mol. Biol., 282,
543–556.
Tetart, F., Repoila, F., Monod, C., and
Krisch, H.M. (1996) Bacteriophage T4
host range is expanded by duplications of a small domain of the tail fiber
adhesin. J. Mol. Biol., 258, 726–731.
Mahichi, F., Synnott, A.J.,
Yamamichi, K., Osada, T., and Tanji, Y.
(2009) Site-specific recombination of
T2 phage using IP008 long tail fiber
genes provides a targeted method for
expanding host range while retaining
lytic activity. FEMS Microbiol. Lett.,
295, 211–217.
Yoichi, M., Abe, M., Miyanaga, K.,
Unno, H., and Tanji, Y. (2005) Alteration of tail fiber protein gp38 enables
T2 phage to infect Escherichia coli
O157:H7. J. Biotechnol., 115, 101–107.
Marti, R., Zurfluh, K., Hagens, S.,
Pianezzi, J., Klumpp, J., and
Loessner, M.J. (2013) Long tail fibres
of the novel broad-host-range T-even
bacteriophage S16 specifically recognize
Salmonella OmpC. Mol. Microbiol., 87,
818–834.
Nguyen, A.H., Molineux, I.J.,
Springman, R., and Bull, J.J. (2012)
Multiple genetic pathways to similar
fitness limits during viral adaptation to
a new host. Evolution, 66, 363–374.
Cuervo, A., Pulido-Cid, M.,
Chagoyen, M., Arranz, R.,
Gonzalez-Garcia, V.A., Garcia-Doval, C.
et al. (2013) Structural characterization
of the bacteriophage T7 tail machinery.
J. Biol. Chem., 288, 26290–26299.
Pouillot, F., Blois, H., and Iris, F. (2010)
Genetically engineered virulent phage
banks in the detection and control of
emergent pathogenic bacteria. Biosecur.
Bioterror., 8, 155–169.
References
173. Stern, A. and Sorek, R. (2011) The
174.
175.
176.
177.
178.
179.
180.
181.
182.
phage-host arms race: shaping the
evolution of microbes. Bioessays, 33,
43–51.
Otsuka, Y. and Yonesaki, T. (2012) Dmd
of bacteriophage T4 functions as an
antitoxin against Escherichia coli LsoA
and RnlA toxins. Mol. Microbiol., 83,
669–681.
Friedman, D.I., Mozola, C.C., Beeri, K.,
Ko, C.C., and Reynolds, J.L. (2011)
Activation of a prophage-encoded tyrosine kinase by a heterologous infecting
phage results in a self-inflicted abortive
infection. Mol. Microbiol., 82, 567–577.
Grissa, I., Vergnaud, G., and Pourcel, C.
(2007) The CRISPRdb database and
tools to display CRISPRs and to generate dictionaries of spacers and repeats.
BMC Bioinformatics, 8, 172.
Makarova, K.S., Haft, D.H.,
Barrangou, R., Brouns, S.J.,
Charpentier, E., Horvath, P. et al.
(2011) Evolution and classification of
the CRISPR-Cas systems. Nat. Rev.
Microbiol., 9, 467–477.
Garneau, J.E., Dupuis, M.E., Villion, M.,
Romero, D.A., Barrangou, R.,
Boyaval, P. et al. (2010) The
CRISPR/Cas bacterial immune system cleaves bacteriophage and plasmid
DNA. Nature, 468, 67–71.
Barrangou, R., Fremaux, C., Deveau, H.,
Richards, M., Boyaval, P., Moineau, S.
et al. (2007) CRISPR provides acquired
resistance against viruses in prokaryotes. Science, 315, 1709–1712.
Jiang, W., Maniv, I., Arain, F., Wang, Y.,
Levin, B.R., and Marraffini, L.A. (2013)
Dealing with the evolutionary downside
of CRISPR immunity: bacteria and
beneficial plasmids. PLoS Genet., 9,
e1003844.
Fineran, P.C. and Charpentier, E.
(2012) Memory of viral infections by
CRISPR-Cas adaptive immune systems: acquisition of new information.
Virology, 434, 202–209.
Hwang, W.Y., Fu, Y., Reyon, D.,
Maeder, M.L., Tsai, S.Q., Sander, J.D.
et al. (2013) Efficient genome editing in
zebrafish using a CRISPR-Cas system.
Nat. Biotechnol., 31, 227–229.
183. Wang, H., Yang, H., Shivalila, C.S.,
184.
185.
186.
187.
188.
189.
190.
191.
192.
Dawlaty, M.M., Cheng, A.W., Zhang, F.
et al. (2013) One-step generation of
mice carrying mutations in multiple genes by CRISPR/Cas-mediated
genome engineering. Cell, 153,
910–918.
Mali, P., Yang, L., Esvelt, K.M., Aach, J.,
Guell, M., DiCarlo, J.E. et al. (2013)
RNA-guided human genome engineering via Cas9. Science, 339, 823–826.
Cong, L., Ran, F.A., Cox, D., Lin, S.,
Barretto, R., Habib, N. et al. (2013)
Multiplex genome engineering using
CRISPR/Cas systems. Science, 339,
819–823.
Cho, S.W., Kim, S., Kim, J.M., and
Kim, J.S. (2013) Targeted genome
engineering in human cells with the
Cas9 RNA-guided endonuclease. Nat.
Biotechnol., 31, 230–232.
Jinek, M., East, A., Cheng, A., Lin, S.,
Ma, E., and Doudna, J. (2013) RNAprogrammed genome editing in human
cells. Elife, 2, e00471.
Gaj, T., Gersbach, C.A., and
Barbas, C.F. III, (2013) ZFN, TALEN,
and CRISPR/Cas-based methods for
genome engineering. Trends Biotechnol., 31 (7), 397–405.
Pattanayak, V., Lin, S., Guilinger, J.P.,
Ma, E., Doudna, J.A., and Liu, D.R.
(2013) High-throughput profiling of
off-target DNA cleavage reveals RNAprogrammed Cas9 nuclease specificity.
Nat. Biotechnol., 31, 839–843.
Mali, P., Aach, J., Stranges, P.B.,
Esvelt, K.M., Moosburner, M.,
Kosuri, S. et al. (2013) CAS9 transcriptional activators for target specificity
screening and paired nickases for
cooperative genome engineering. Nat.
Biotechnol., 31, 833–838.
Hsu, P.D., Scott, D.A., Weinstein, J.A.,
Ran, F.A., Konermann, S., Agarwala, V.
et al. (2013) DNA targeting specificity
of RNA-guided Cas9 nucleases. Nat.
Biotechnol., 31, 827–832.
Fu, Y., Foden, J.A., Khayter, C.,
Maeder, M.L., Reyon, D., Joung, J.K.
et al. (2013) High-frequency off-target
mutagenesis induced by CRISPRCas nucleases in human cells. Nat.
Biotechnol., 31, 822–826.
167
168
7 Synthetic Biology and Therapies for Infectious Diseases
193. Bikard, D., Jiang, W., Samai, P.,
194.
195.
196.
197.
198.
199.
200.
201.
202.
203.
204.
Hochschild, A., Zhang, F., and
Marraffini, L.A. (2013) Programmable
repression and activation of bacterial
gene expression using an engineered
CRISPR-Cas system. Nucleic Acids Res.,
41, 7429–7437.
Farzadfard, F., Perli, S.D., and Lu, T.K.
(2013) Tunable and multifunctional
eukaryotic transcription factors based
on CRISPR/Cas. ACS Synth. Biol., 2,
604–613.
Switala-Jelen, K., Dabrowska, K.,
Gorski, A., and Sliwa, L. (2002) Mutations in bacteriophage T4 genome. Acta
Virol., 46, 57–62.
Gibson, D.G. (2011) Gene and genome
construction in yeast. Curr. Protoc. Mol.
Biol., 94, 3.22.1–3.22.17.
Lu, T.K., Koeris, M., Chevalier, B.,
Holder, J., McKensie, G., and
Brownell, D. (2012) Recombinant
phage and methods. US Patent US
13/627,060.
Lu, T.K. and Collins, J.J. (2007) Dispersing biofilms with engineered enzymatic
bacteriophage. Proc. Natl. Acad. Sci.
U.S.A., 104, 11197–11202.
Dwyer, D.J., Kohanski, M.A., Hayete, B.,
and Collins, J.J. (2007) Gyrase inhibitors
induce an oxidative damage cellular
death pathway in Escherichia coli. Mol.
Syst. Biol., 3, 91.
Kohanski, M.A., Dwyer, D.J., Hayete, B.,
Lawrence, C.A., and Collins, J.J. (2007)
A common mechanism of cellular
death induced by bactericidal antibiotics. Cell, 130, 797–810.
Lu, T.K. and Collins, J.J. (2009) Engineered bacteriophage targeting gene
networks as adjuvants for antibiotic
therapy. Proc. Natl. Acad. Sci. U.S.A.,
106, 4629–4634.
Lin, A., Jimenez, J., Derr, J., Vera, P.,
Manapat, M.L., Esvelt, K.M. et al.
(2011) Inhibition of bacterial conjugation by phage M13 and its protein g3p:
quantitative analysis and model. PLoS
One, 6, e19991.
Young, R. (1992) Bacteriophage lysis:
mechanism and regulation. Microbiol.
Rev., 56, 430–481.
Paul, V.D., Sundarrajan, S.,
Rajagopalan, S.S., Hariharan, S.,
205.
206.
207.
208.
209.
210.
211.
212.
213.
Kempashanaiah, N., Padmanabhan, S.
et al. (2011) Lysis-deficient phages as
novel therapeutic agents for controlling
bacterial infection. BMC Microbiol., 11,
195.
Hagens, S., Habel, A., von Ahsen, U.,
von Gabain, A., and Blasi, U. (2004)
Therapy of experimental pseudomonas
infections with a nonreplicating genetically modified phage. Antimicrob.
Agents Chemother., 48, 3817–3822.
Hagens, S. and Blasi, U. (2003) Genetically modified filamentous phage as
bactericidal agents: a pilot study. Lett.
Appl. Microbiol., 37, 318–323.
Matsuda, T., Freeman, T.A.,
Hilbert, D.W., Duff, M., Fuortes, M.,
Stapleton, P.P. et al. (2005) Lysisdeficient bacteriophage therapy
decreases endotoxin and inflammatory mediator release and improves
survival in a murine peritonitis model.
Surgery, 137, 639–646.
Mueller, S., Papamichail, D.,
Coleman, J.R., Skiena, S., and
Wimmer, E. (2006) Reduction of the
rate of poliovirus protein synthesis
through large-scale codon deoptimization causes attenuation of viral
virulence by lowering specific infectivity. J. Virol., 80, 9687–9696.
Merril, C.R., Biswas, B., Carlton, R.,
Jensen, N.C., Creed, G.J., Zullo, S. et al.
(1996) Long-circulating bacteriophage
as antibacterial agents. Proc. Natl.
Acad. Sci. U.S.A., 93, 3188–3192.
Brussow, H., Canchaya, C., and
Hardt, W.D. (2004) Phages and the
evolution of bacterial pathogens: from
genomic rearrangements to lysogenic
conversion. Microbiol. Mol. Biol. Rev.,
68, 560–602, table of contents.
Moons, P., Faster, D., and Aertsen, A.
(2013) Lysogenic conversion and phage
resistance development in phage
exposed Escherichia coli biofilms.
Viruses, 5, 150–161.
Sternberg, N.L. and Maurer, R. (1991)
Bacteriophage-mediated generalized
transduction in Escherichia coli and
Salmonella typhimurium. Methods
Enzymol., 204, 18–43.
Melnikov, A.A., Tchernov, A.P.,
Fodor, I., and Bayev, A.A. (1984)
References
214.
215.
216.
217.
218.
219.
220.
221.
222.
223.
Lambda phagemids and their transducing properties. Gene, 28, 29–35.
Westwater, C., Schofield, D.A.,
Schmidt, M.G., Norris, J.S., and
Dolan, J.W. (2002) Development of
a P1 phagemid system for the delivery
of DNA into Gram-negative bacteria.
Microbiology, 148, 943–950.
Edgar, R., Friedman, N.,
Molshanski-Mor, S., and Qimron, U.
(2012) Reversing bacterial resistance
to antibiotics by phage-mediated delivery of dominant sensitive genes. Appl.
Environ. Microbiol., 78, 744–751.
Qi, H., Lu, H.Q., Qiu, H.J., Petrenko, V.,
and Liu, A.H. (2012) Phagemid vectors
for phage display: properties, characteristics and construction. J. Mol. Biol.,
417, 129–143.
Yanischperron, C., Vieira, J., and
Messing, J. (1985) Improved M13
phage cloning vectors and host strains nucleotide-sequences of the M13mp18
and Puc19 vectors. Gene, 33, 103–119.
Westwater, C., Kasman, L.M.,
Schofield, D.A., Werner, P.A.,
Dolan, J.W., Schmidt, M.G. et al. (2003)
Use of genetically engineered phage to
deliver antimicrobial agents to bacteria:
an alternative therapy for treatment of
bacterial infections. Antimicrob. Agents
Chemother., 47, 1301–1307.
Lennox, E.S. (1955) Transduction of
linked genetic characters of the host by
bacteriophage P1. Virology, 1, 190–206.
Kittleson, J.T., Deloache, W.,
Cheng, H.Y., and Anderson, J.C. (2012)
Scalable plasmid transfer using engineered P1-based phagemids. ACS
Synth. Biol., 1, 583–589.
Fairhead, H. (2009) Sasp gene delivery:
a novel antibacterial approach. Drug
News Perspect., 22, 197–203.
Lu, T.K., Bowers, J., and Koeris, M.S.
(2013) Advancing bacteriophagebased microbial diagnostics with
synthetic biology. Trends Biotechnol.,
31, 325–327.
Haq, I.U., Chaudhry, W.N.,
Akhtar, M.N., Andleeb, S., and Qadri, I.
(2012) Bacteriophages and their implications on future biotechnology: a
review. Virol. J., 9, 9.
224. Bhattarai, S.R., Yoo, S.Y., Lee, S.W., and
225.
226.
227.
228.
229.
230.
231.
232.
Dean, D. (2012) Engineered phagebased therapeutic materials inhibit
Chlamydia trachomatis intracellular
infection. Biomaterials, 33, 5166–5174.
Embleton, M.L., Nair, S.P.,
Heywood, W., Menon, D.C.,
Cookson, B.D., and Wilson, M. (2005)
Development of a novel targeting system for lethal photosensitization of
antibiotic-resistant strains of Staphylococcus aureus. Antimicrob. Agents
Chemother., 49, 3690–3696.
Yacoby, I., Shamis, M., Bar, H.,
Shabat, D., and Benhar, I. (2006) Targeting antibacterial agents by using
drug-carrying filamentous bacteriophages. Antimicrob. Agents Chemother.,
50, 2087–2097.
Seed, K.D., Lazinski, D.W.,
Calderwood, S.B., and Camilli, A.
(2013) A bacteriophage encodes its own
CRISPR/Cas adaptive response to evade
host innate immunity. Nature, 494,
489–491.
Fischetti, V.A. (2011) Exploiting what
phage have evolved to control Grampositive pathogens. Bacteriophage, 1,
188–194.
Djurkovic, S., Loeffler, J.M., and
Fischetti, V.A. (2005) Synergistic killing
of Streptococcus pneumoniae with
the bacteriophage lytic enzyme Cpl-1
and penicillin or gentamicin depends
on the level of penicillin resistance.
Antimicrob. Agents Chemother., 49,
1225–1228.
Loeffler, J.M. and Fischetti, V.A.
(2003) Synergistic lethal effect of a
combination of phage lytic enzymes
with different activities on penicillinsensitive and -resistant Streptococcus
pneumoniae strains. Antimicrob. Agents
Chemother., 47, 375–377.
Tafesh, A., Najami, N., Jadoun, J.,
Halahlih, F., Riepl, H., and Azaizeh, H.
(2011) Synergistic antibacterial effects
of polyphenolic compounds from
olive mill wastewater. Evid. Based
Complement. Alternat. Med., 2011,
431021.
Daniel, A., Euler, C., Collin, M.,
Chahales, P., Gorelick, K.J., and
Fischetti, V.A. (2010) Synergism
169
170
7 Synthetic Biology and Therapies for Infectious Diseases
233.
234.
235.
236.
237.
238.
239.
240.
between a novel chimeric lysin and
oxacillin protects against infection by
methicillin-resistant Staphylococcus
aureus. Antimicrob. Agents Chemother.,
54, 1603–1612.
Diaz, E., Lopez, R., and Garcia, J.L.
(1990) Chimeric phage-bacterial
enzymes: a clue to the modular evolution of genes. Proc. Natl. Acad. Sci.
U.S.A., 87, 8125–8129.
Donovan, D.M., Dong, S., Garrett, W.,
Rousseau, G.M., Moineau, S., and
Pritchard, D.G. (2006) Peptidoglycan hydrolase fusions maintain their
parental specificities. Appl. Environ.
Microbiol., 72, 2988–2996.
Rashel, M., Uchiyama, J., Ujihara, T.,
Uehara, Y., Kuramoto, S., Sugihara, S.
et al. (2007) Efficient elimination of
multidrug-resistant Staphylococcus
aureus by cloned lysin derived from
bacteriophage phi MR11. J. Infect. Dis.,
196, 1237–1247.
Gu, J., Xu, W., Lei, L., Huang, J.,
Feng, X., Sun, C. et al. (2011) LysGH15,
a novel bacteriophage lysin, protects
a murine bacteremia model efficiently
against lethal methicillin-resistant
Staphylococcus aureus infection. J. Clin.
Microbiol., 49, 111–117.
Lukacik, P., Barnard, T.J., Keller, P.W.,
Chaturvedi, K.S., Seddiki, N.,
Fairman, J.W. et al. (2012) Structural engineering of a phage lysin
that targets Gram-negative pathogens.
Proc. Natl. Acad. Sci. U.S.A., 109,
9857–9862.
Cotter, P.D., Ross, R.P., and Hill, C.
(2013) Bacteriocins - a viable alternative to antibiotics? Nat. Rev. Microbiol.,
11, 95–105.
Rakin, A., Saken, E., Harmsen, D., and
Heesemann, J. (1994) The pesticin
receptor of Yersinia enterocolitica:
a novel virulence factor with dual
function. Mol. Microbiol., 13, 253–263.
Schubert, S., Rakin, A., Karch, H.,
Carniel, E., and Heesemann, J. (1998)
Prevalence of the high-pathogenicity
island of Yersinia species among
Escherichia coli strains that are
pathogenic to humans. Infect. Immun.,
66, 480–485.
241. Wang, Y., Geer, L.Y., Chappey, C.,
242.
243.
244.
245.
246.
247.
248.
249.
250.
Kans, J.A., and Bryant, S.H. (2000)
Cn3D: sequence and structure views
for Entrez. Trends Biochem. Sci., 25,
300–302.
Michel-Briand, Y. and Baysse, C. (2002)
The pyocins of Pseudomonas aeruginosa. Biochimie, 84, 499–510.
Scholl, D., Gebhart, D., Williams, S.R.,
Bates, A., and Mandrell, R. (2012)
Genome sequence of E. coli O104:H4
leads to rapid development of a targeted antimicrobial agent against this
emerging pathogen. PLoS One, 7,
e33637.
Ritchie, J.M., Greenwich, J.L.,
Davis, B.M., Bronson, R.T., Gebhart, D.,
Williams, S.R. et al. (2011) An
Escherichia coli O157-specific engineered pyocin prevents and ameliorates
infection by E. coli O157:H7 in an
animal model of diarrheal disease.
Antimicrob. Agents Chemother., 55,
5469–5474.
Scholl, D., Cooley, M., Williams, S.R.,
Gebhart, D., Martin, D., Bates, A. et al.
(2009) An engineered R-type pyocin is
a highly specific and sensitive bactericidal agent for the food-borne pathogen
Escherichia coli O157:H7. Antimicrob.
Agents Chemother., 53, 3074–3080.
Williams, S.R., Gebhart, D.,
Martin, D.W., and Scholl, D. (2008)
Retargeting R-type pyocins to generate novel bactericidal protein
complexes. Appl. Environ. Microbiol.,
74, 3868–3876.
Gebhart, D., Williams, S.R.,
Bishop-Lilly, K.A., Govoni, G.R.,
Willner, K.M., Butani, A. et al. (2012)
Novel high-molecular-weight, R-type
bacteriocins of Clostridium difficile.
J. Bacteriol., 194, 6240–6247.
Weinbauer, M.G. (2004) Ecology of
prokaryotic viruses. FEMS Microbiol.
Rev., 28, 127–181.
Wood, W.B. and Revel, H.R. (1976) The
genome of bacteriophage T4. Bacteriol.
Rev., 40, 847–868.
Fu, T.J., Geiduschek, E.P., and
Kassavetis, G.A. (1998) Abortive initiation of transcription at a hybrid
promoter. An analysis of the sliding
clamp activator of bacteriophage T4
References
251.
252.
253.
254.
255.
256.
257.
258.
259.
260.
late transcription, and a comparison
of the sigma70 and T4 gp55 promoter
recognition proteins. J. Biol. Chem.,
273, 34042–34048.
Savalia, D., Westblade, L.F., Goel, M.,
Florens, L., Kemp, P., Akulenko, N.
et al. (2008) Genomic and proteomic analysis of phiEco32, a novel
Escherichia coli bacteriophage. J. Mol.
Biol., 377, 774–789.
Jamalludeen, N., Kropinski, A.M.,
Johnson, R.P., Lingohr, E., Harel, J., and
Gyles, C.L. (2008) Complete genomic
sequence of bacteriophage phiEcoMGJ1, a novel phage that has myovirus
morphology and a podovirus-like RNA
polymerase. Appl. Environ. Microbiol.,
74, 516–525.
Studier, F.W. (1972) Bacteriophage T7.
Science, 176, 367–376.
Yuzenkova, J., Nechaev, S., Berlin, J.,
Rogulja, D., Kuznedelov, K., Inman, R.
et al. (2003) Genome of Xanthomonas
oryzae bacteriophage Xp10: an odd Todd phage. J. Mol. Biol., 330, 735–748.
Kulakov, L.A., Ksenzenko, V.N.,
Shlyapnikov, M.G., Kochetkov, V.V.,
Del Casale, A., Allen, C.C. et al. (2009)
Genomes of phiKMV-like viruses
of Pseudomonas aeruginosa contain
localized single-strand interruptions.
Virology, 391, 1–4.
Fischbach, M.A., Bluestone, J.A., and
Lim, W.A. (2013) Cell-based therapeutics: the next pillar of medicine. Sci.
Transl. Med., 5, 179ps7.
Guerrero, R., Pedros-Alio, C., Esteve, I.,
Mas, J., Chase, D., and Margulis, L.
(1986) Predatory prokaryotes: predation
and primary consumption evolved in
bacteria. Proc. Natl. Acad. Sci. U.S.A.,
83, 2138–2142.
Jurkevitch, E. (2012) Isolation and
classification of Bdellovibrio and like
organisms. Curr. Protoc. Microbiol., 26,
7B.1.1–7B.1.20.
Pasternak, Z., Njagi, M.,
Shani, Y., Chanyi, R., Rotem, O.,
Lurie-Weinberger, M.N. et al. (2013)
In and out: an analysis of epibiotic
vs periplasmic bacterial predators.
ISME J., 8 (3), 625–635.
Richardson, I.R. (1990) The incidence
of Bdellovibrio spp. in man-made water
261.
262.
263.
264.
265.
266.
267.
268.
269.
systems: coexistence with legionellas.
J. Appl. Bacteriol., 69, 134–140.
Richards, G.P., Fay, J.P., Dickens, K.A.,
Parent, M.A., Soroka, D.S., and
Boyd, E.F. (2012) Predatory bacteria
as natural modulators of Vibrio parahaemolyticus and Vibrio vulnificus in
seawater and oysters. Appl. Environ.
Microbiol., 78, 7455–7466.
Iebba, V., Santangelo, F., Totino, V.,
Nicoletti, M., Gagliardi, A.,
De Biase, R.V. et al. (2013) Higher
prevalence and abundance of Bdellovibrio bacteriovorus in the human gut of
healthy subjects. PLoS One, 8, e61608.
Kadouri, D.E., To, K., Shanks, R.M.,
and Doi, Y. (2013) Predatory bacteria:
a potential ally against multidrugresistant Gram-negative pathogens.
PLoS One, 8, e63397.
Chu, W.H. and Zhu, W. (2010) Isolation of Bdellovibrio as biological
therapeutic agents used for the
treatment of Aeromonas hydrophila
infection in fish. Zoonoses Public
Health, 57, 258–264.
Atterbury, R.J., Hobley, L., Till, R.,
Lambert, C., Capeness, M.J.,
Lerner, T.R. et al. (2011) Effects
of orally administered Bdellovibrio
bacteriovorus on the well-being and
Salmonella colonization of young
chicks. Appl. Environ. Microbiol., 77,
5794–5803.
Steyert, S.R. and Pineiro, S.A. (2007)
Development of a novel genetic system
to create markerless deletion mutants
of Bdellovibrio bacteriovorus. Appl.
Environ. Microbiol., 73, 4717–4724.
Sockett, R.E. and Lambert, C. (2004)
Bdellovibrio as therapeutic agents:
a predatory renaissance? Nat. Rev.
Microbiol., 2, 669–675.
Nermes, M., Salminen, S., and
Isolauri, E. (2013) Is there a role for
probiotics in the prevention or treatment of food allergy? Curr. Allergy
Asthma Rep., 13, 622–630.
Friedman, G. (2012) The role of probiotics in the prevention and treatment
of antibiotic-associated diarrhea and
Clostridium difficile colitis. Gastroenterol. Clin. North Am., 41, 763–779.
171
172
7 Synthetic Biology and Therapies for Infectious Diseases
270. Veerappan, G.R., Betteridge, J., and
271.
272.
273.
274.
275.
276.
277.
278.
279.
Young, P.E. (2012) Probiotics for the
treatment of inflammatory bowel disease. Curr. Gastroenterol. Rep., 14,
324–333.
Jonkers, D., Penders, J., Masclee, A.,
and Pierik, M. (2012) Probiotics in the
management of inflammatory bowel
disease: a systematic review of intervention studies in adult patients. Drugs,
72, 803–823.
Carey, C.M. and Kostrzynska, M.
(2013) Lactic acid bacteria and
bifidobacteria attenuate the proinflammatory response in intestinal
epithelial cells induced by Salmonella
enterica serovar Typhimurium. Can. J.
Microbiol., 59, 9–17.
Carey, C.M., Kostrzynska, M., Ojha, S.,
and Thompson, S. (2008) The effect
of probiotics and organic acids on
Shiga-toxin 2 gene expression in
enterohemorrhagic Escherichia coli
O157:H7. J. Microbiol. Methods, 73,
125–132.
Adams, C.A. (2010) The probiotic paradox: live and dead cells are biological
response modifiers. Nutr. Res. Rev., 23,
37–46.
Aggarwal, J., Swami, G., and Kumar, M.
(2013) Probiotics and their effects on
metabolic diseases: an update. J. Clin.
Diagn. Res., 7, 173–177.
Szachta, P., Ignys, I., and Cichy, W.
(2011) An evaluation of the ability
of the probiotic strain Lactobacillus rhamnosus GG to eliminate
the gastrointestinal carrier state of
vancomycin-resistant enterococci in
colonized children. J. Clin. Gastroenterol., 45, 872–877.
Sanders, M.E., Heimbach, J.T., Pot, B.,
Tancredi, D.J., Lenoir-Wijnkoop, I.,
Lahteenmaki-Uutela, A. et al. (2011)
Health claims substantiation for probiotic and prebiotic products. Gut
Microbes, 2, 127–133.
Sanders, M.E. and Levy, D.D. (2011)
The science and regulations of probiotic
food and supplement product labeling.
Ann. N. Y. Acad. Sci., 1219 (Suppl. 1),
E1–E23.
Duan, F. and March, J.C. (2010) Engineered bacterial communication
280.
281.
282.
283.
284.
285.
286.
287.
288.
289.
290.
prevents Vibrio cholerae virulence
in an infant mouse model. Proc. Natl.
Acad. Sci. U.S.A., 107, 11260–11264.
Schuster, M., Sexton, D.J., Diggle, S.P.,
and Greenberg, E.P. (2013) Acylhomoserine lactone quorum sensing:
from evolution to application. Annu.
Rev. Microbiol., 67, 43–63.
Fuqua, W.C., Winans, S.C., and
Greenberg, E.P. (1994) Quorum sensing
in bacteria: the LuxR-LuxI family of
cell density-responsive transcriptional
regulators. J. Bacteriol., 176, 269–275.
Saeidi, N., Wong, C.K., Lo, T.M.,
Nguyen, H.X., Ling, H., Leong, S.S.
et al. (2011) Engineering microbes
to sense and eradicate Pseudomonas
aeruginosa, a human pathogen. Mol.
Syst. Biol., 7, 521.
Gupta, S., Bram, E.E., and Weiss, R.
(2013) Genetically programmable
pathogen sense and destroy. ACS Synth.
Biol., 2 (12), 715–723.
Hwang, I.Y., Tan, M.H., Koh, E.,
Ho, C.L., Poh, C.L., and Chang, M.W.
(2013) Reprogramming microbes to be
pathogen-seeking killers. ACS Synth.
Biol., 3 (4), 228–237.
Daniel, R., Rubens, J., Sarpeshkar, R.,
and Lu, T.K. (2013) Synthetic analog
computation in living cells. Nature,
1497 (7451), 619–623.
Siuti, P., Yazbek, J., and Lu, T.K. (2013)
Synthetic circuits integrating logic and
memory in living cells. Nat. Biotechnol.,
31 (5), 448–452.
Benenson, Y. (2012) Biomolecular computing systems: principles, progress
and potential. Nat. Rev. Genet., 13,
455–468.
Lu, T.K., Khalil, A.S., and Collins, J.J.
(2009) Next-generation synthetic
gene networks. Nat. Biotechnol., 27,
1139–1150.
Gough, E., Shaikh, H., and
Manges, A.R. (2011) Systematic review
of intestinal microbiota transplantation
(fecal bacteriotherapy) for recurrent
Clostridium difficile infection. Clin.
Infect. Dis., 53, 994–1002.
Lemon, K.P., Armitage, G.C.,
Relman, D.A., and Fischbach, M.A.
(2012) Microbiota-targeted therapies:
References
291.
292.
293.
294.
295.
296.
297.
298.
299.
300.
301.
an ecological perspective. Sci. Transl.
Med., 4, 137rv5.
Demain, A.L. (1999) Pharmaceutically active secondary metabolites
of microorganisms. Appl. Microbiol.
Biotechnol., 52, 455–463.
Alexander, F. (1929) On the antibacetrial action of cultures of a penicillium,
with special references to their use in
the isolation of B. influenzae. Br. J. Exp.
Pathol., 10, 226–236.
Bentley, R. (2005) The development of
penicillin: genesis of a famous antibiotic. Perspect. Biol. Med., 48, 444–452.
Fleming, A. (2001) On the antibacterial
action of cultures of a penicillium, with
special reference to their use in the
isolation of B. influenzæ. Bull. World
Health Organ., 79, 780–790.
Bergmann, S., Schumann, J.,
Scherlach, K., Lange, C.,
Brakhage, A.A., and Hertweck, C.
(2007) Genomics-driven discovery of
PKS-NRPS hybrid metabolites from
Aspergillus nidulans. Nat. Chem. Biol.,
3, 213–217.
Corre, C. and Challis, G.L. (2009) New
natural product biosynthetic chemistry
discovered by genome mining. Nat.
Prod. Rep., 26, 977–986.
Ganesan, A. (2008) ChemInform
abstract: natural products and combinatorial chemistry — An uneasy past
but a glorious future. ChemInform, 39,
37–52.
Zakeri, B. and Lu, T.K. (2012) Synthetic
biology of antimicrobial discovery. ACS
Synth. Biol., 2 (7), 358–372.
Medema, M.H., Breitling, R., and
Takano, E. (2011) Synthetic biology
in Streptomyces bacteria. Methods
Enzymol., 497, 485–502.
Clardy, J., Fischbach, M.A., and
Walsh, C.T. (2006) New antibiotics
from bacterial natural products. Nat.
Biotechnol., 24, 1541–1550.
Fischbach, M.A. and Walsh, C.T.
(2006) Assembly-line enzymology
for polyketide and nonribosomal Peptide antibiotics: logic, machinery,
and mechanisms. Chem. Rev., 106,
3468–3496.
302. Hertweck, C. (2009) The biosynthetic
303.
304.
305.
306.
307.
308.
309.
310.
311.
logic of polyketide diversity. Angew.
Chem. Int. Ed., 48, 4688–4716.
Mootz, H.D., Schwarzer, D., and
Marahiel, M.A. (2002) Ways of assembling complex natural products on
modular nonribosomal peptide synthetases. ChemBioChem, 3, 490–504.
Weissman, K.J. (2004) Polyketide
biosynthesis: understanding and
exploiting modularity. Philos. Trans. A
Math. Phys. Eng. Sci., 362, 2671–2690.
Zerikly, M. and Challis, G.L. (2009)
Strategies for the discovery of new
natural products by genome mining.
ChemBioChem, 10, 625–633.
Schroeckh, V., Scherlach, K.,
Nutzmann, H.W., Shelest, E.,
Schmidt-Heck, W., Schuemann, J.
et al. (2009) Intimate bacterial-fungal
interaction triggers biosynthesis of
archetypal polyketides in Aspergillus
nidulans. Proc. Natl. Acad. Sci. U.S.A.,
106, 14558–14563.
Starcevic, A., Zucko, J., Simunkovic, J.,
Long, P.F., Cullum, J., and Hranueli, D.
(2008) ClustScan: an integrated program package for the semi-automatic
annotation of modular biosynthetic
gene clusters and in silico prediction
of novel chemical structures. Nucleic
Acids Res., 36, 6882–6892.
Li, M.H., Ung, P.M., Zajkowski, J.,
Garneau-Tsodikova, S., and
Sherman, D.H. (2009) Automated
genome mining for natural products.
BMC Bioinformatics, 10, 185.
Medema, M.H., Blin, K.,
Cimermancic, P., de Jager, V.,
Zakrzewski, P., Fischbach, M.A. et al
(2011) antiSMASH: rapid identification,
annotation and analysis of secondary
metabolite biosynthesis gene clusters in
bacterial and fungal genome sequences.
Nucleic Acids Res., 39, W339–W346.
Marchler-Bauer, A., Anderson, J.B.,
Chitsaz, F., Derbyshire, M.K.,
DeWeese-Scott, C., Fong, J.H. et al.
(2009) CDD: specific functional annotation with the Conserved Domain
Database. Nucleic Acids Res., 37,
D205–D210.
Marchler-Bauer, A. and Bryant, S.H.
(2004) CD-Search: protein domain
173
174
7 Synthetic Biology and Therapies for Infectious Diseases
312.
313.
314.
315.
316.
317.
318.
319.
320.
321.
annotations on the fly. Nucleic Acids
Res., 32, W327–W331.
Marchler-Bauer, A., Lu, S.,
Anderson, J.B., Chitsaz, F.,
Derbyshire, M.K., DeWeese-Scott, C.
et al. (2011) CDD: a Conserved
Domain Database for the functional
annotation of proteins. Nucleic Acids
Res., 39, D225–D229.
Letunic, I., Doerks, T., and Bork, P.
(2009) SMART 6: recent updates and
new developments. Nucleic Acids Res.,
37, D229–D232.
Schultz, J., Milpetz, F., Bork, P., and
Ponting, C.P. (1998) SMART, a simple
modular architecture research tool:
identification of signaling domains.
Proc. Natl. Acad. Sci., 95, 5857–5864.
Rausch, C., Weber, T., Kohlbacher, O.,
Wohlleben, W., and Huson, D.H. (2005)
Specificity prediction of adenylation
domains in nonribosomal peptide synthetases (NRPS) using transductive
support vector machines (TSVMs).
Nucleic Acids Res., 33, 5799–5808.
Rottig, M., Medema, M.H.,
Blin, K., Weber, T., Rausch, C.,
and Kohlbacher, O.
(2011) NRPSpredictor2 –
a web server for predicting NRPS
adenylation domain specificity. Nucleic
Acids Res., 39, W362–W367.
Ansari, M.Z., Yadav, G., Gokhale, R.S.,
and Mohanty, D. (2004) NRPS-PKS: a
knowledge-based resource for analysis
of NRPS/PKS megasynthases. Nucleic
Acids Res., 32, W405–W413.
Anand, S., Prasad, M.V., Yadav, G.,
Kumar, N., Shehara, J., Ansari, M.Z.
et al. (2010) SBSPKS: structure
based sequence analysis of polyketide synthases. Nucleic Acids Res., 38,
W487–W496.
Florence, S. and Jürgen, B. (2006) Combinatorial Synthesis of Natural ProductBased Libraries, CRC Press, pp. 53–64.
Wilkinson, B. and Micklefield, J. (2007)
Mining and engineering naturalproduct biosynthetic pathways. Nat.
Chem. Biol., 3, 379–386.
Bucar, F., Wube, A., and Schmid, M.
(2013) Natural product isolation – how
to get from biological material to
322.
323.
324.
325.
326.
327.
328.
329.
330.
pure compounds. Nat. Prod. Rep., 30,
525–545.
Winter, J.M., Behnken, S., and
Hertweck, C. (2011) Genomics-inspired
discovery of natural products. Curr.
Opin. Chem. Biol., 15, 22–31.
Kersten, R.D., Yang, Y.L., Xu, Y.,
Cimermancic, P., Nam, S.J., Fenical, W.
et al. (2011) A mass spectrometryguided genome mining approach for
natural product peptidogenomics. Nat.
Chem. Biol., 7, 794–802.
Ibrahim, A., Yang, L., Johnston, C.,
Liu, X., Ma, B., and Magarvey, N.A.
(2012) Dereplicating nonribosomal
peptides using an informatic search
algorithm for natural products (iSNAP)
discovery. Proc. Natl. Acad. Sci. U.S.A.,
109, 19196–19201.
Grigoriev, I.V., Nordberg, H.,
Shabalov, I., Aerts, A., Cantor, M.,
Goodstein, D. et al. (2012) The genome
portal of the Department of Energy
Joint Genome Institute. Nucleic Acids
Res., 40, D26–D32.
Ward, B.B. (2002) How many species of
prokaryotes are there? Proc. Natl. Acad.
Sci. U.S.A., 99, 10234–10236.
Yin, J., Straight, P.D., Hrvatin, S.,
Dorrestein, P.C., Bumpus, S.B., Jao, C.
et al. (2007) Genome-wide highthroughput mining of natural-product
biosynthetic gene clusters by phage
display. Chem. Biol., 14, 303–312.
Hildebrand, M., Waggoner, L.E.,
Lim, G.E., Sharp, K.H., Ridley, C.P.,
and Haygood, M.G. (2004) Approaches
to identify, clone, and express symbiont
bioactive metabolite genes. Nat. Prod.
Rep., 21, 122–142.
Kalaitzis, J.A., Lauro, F.M., and
Neilan, B.A. (2009) Mining cyanobacterial genomes for genes encoding
complex biosynthetic pathways. Nat.
Prod. Rep., 26, 1447–1465.
Schirmer, A., Gadkari, R., Reeves, C.D.,
Ibrahim, F., DeLong, E.F., and
Hutchinson, C.R. (2005) Metagenomic
analysis reveals diverse polyketide synthase gene clusters in microorganisms
associated with the marine sponge
Discodermia dissoluta. Appl. Environ.
Microbiol., 71, 4840–4849.
References
331. Brady, S.F., Simmons, L., Kim, J.H., and
332.
333.
334.
335.
336.
337.
338.
339.
340.
341.
Schmidt, E.W. (2009) Metagenomic
approaches to natural products from
free-living and symbiotic organisms.
Nat. Prod. Rep., 26, 1488–1503.
Ehrenreich, I.M., Waterbury, J.B., and
Webb, E.A. (2005) Distribution and
diversity of natural product genes in
marine and freshwater cyanobacterial
cultures and genomes. Appl. Environ.
Microbiol., 71, 7401–7413.
Gontang, E.A., Gaudencio, S.P.,
Fenical, W., and Jensen, P.R. (2010)
Sequence-based analysis of secondarymetabolite biosynthesis in marine
actinobacteria. Appl. Environ. Microbiol., 76, 2487–2499.
Savic, M. and Vasiljevic, B. (2006) Targeting polyketide synthase gene pool
within actinomycetes: new degenerate
primers. J. Ind. Microbiol. Biotechnol.,
33, 423–430.
Walsh, C.T. (2002) Combinatorial
biosynthesis of antibiotics: challenges
and opportunities. ChemBioChem, 3,
124–134.
Wang, J., Xiong, Z., Meng, H.,
Wang, Y., and Wang, Y. (2012) Synthetic biology triggers new era of
antibiotics development. Subcell.
Biochem., 64, 95–114.
Winter, J.M. and Tang, Y. (2012) Synthetic biological approaches to natural
product biosynthesis. Curr. Opin.
Biotechnol., 23, 736–743.
Wong, F.T. and Khosla, C. (2012) Combinatorial biosynthesis of polyketides –
a perspective. Curr. Opin. Chem. Biol.,
16, 117–123.
Ziermann, R. and Betlach, M.C. (1999)
Recombinant polyketide synthesis in Streptomyces: engineering of
improved host strains. Biotechniques,
26, 106–110.
Chen, Y., Smanski, M.J., and Shen, B.
(2010) Improvement of secondary
metabolite production in Streptomyces
by manipulating pathway regulation.
Appl. Microbiol. Biotechnol., 86, 19–25.
Xue, Y., Zhao, L., Liu, H.W., and
Sherman, D.H. (1998) A gene cluster for macrolide antibiotic biosynthesis
342.
343.
344.
345.
346.
347.
348.
349.
350.
351.
in Streptomyces venezuelae: architecture of metabolic diversity. Proc. Natl.
Acad. Sci. U.S.A., 95, 12111–12116.
Fields, S. and Johnston, M. (2005)
Cell biology. Whither model organism
research? Science, 307, 1885–1886.
Nawy, T. (2011) Non–model organisms. Nat. Methods, 9, 37.
Ro, D.K., Paradise, E.M., Ouellet, M.,
Fisher, K.J., Newman, K.L.,
Ndungu, J.M. et al. (2006) Production of the antimalarial drug precursor
artemisinic acid in engineered yeast.
Nature, 440, 940–943.
Pfeifer, B.A., Admiraal, S.J.,
Gramajo, H., Cane, D.E., and Khosla, C.
(2001) Biosynthesis of complex
polyketides in a metabolically engineered strain of E. coli. Science, 291,
1790–1792.
Pickens, L.B., Tang, Y., and Chooi, Y.H.
(2011) Metabolic engineering for the
production of natural products. Annu.
Rev. Chem. Biomol. Eng., 2, 211–236.
Malla, S., Niraula, N.P., Liou, K., and
Sohng, J.K. (2010) Improvement in
doxorubicin productivity by overexpression of regulatory genes in
Streptomyces peucetius. Res. Microbiol.,
161, 109–117.
Smanski, M.J., Peterson, R.M.,
Rajski, S.R., and Shen, B. (2009) Engineered Streptomyces platensis strains
that overproduce antibiotics platensimycin and platencin. Antimicrob.
Agents Chemother., 53, 1299–1304.
Ehmann, D.E., Shaw-Reid, C.A.,
Losey, H.C., and Walsh, C.T. (2000)
The EntF and EntE adenylation
domains of Escherichia coli enterobactin synthetase: sequestration and
selectivity in acyl-AMP transfers to
thiolation domain cosubstrates. Proc.
Natl. Acad. Sci. U.S.A., 97, 2509–2514.
Dittmann, J., Wenger, R.M.,
Kleinkauf, H., and Lawen, A. (1994)
Mechanism of cyclosporin A biosynthesis. Evidence for synthesis via a single
linear undecapeptide precursor. J. Biol.
Chem., 269, 2841–2846.
Gibson, D.G., Young, L., Chuang, R.Y.,
Venter, J.C., Hutchison, C.A. III, and
Smith, H.O. (2009) Enzymatic assembly of DNA molecules up to several
175
176
7 Synthetic Biology and Therapies for Infectious Diseases
352.
353.
354.
355.
356.
357.
358.
359.
hundred kilobases. Nat. Methods, 6,
343–345.
Engler, C., Kandzia, R., and
Marillonnet, S. (2008) A one pot,
one step, precision cloning method
with high throughput capability. PLoS
One, 3, e3647.
Pelzer, S., Wohlert, S.E., and Vente, A.
(2005) Tool-box: tailoring enzymes for
bio-combinatorial lead development
and as markers for genome-based natural product lead discovery. Ernst
Schering Res. Found. Workshop,
233–259.
Zhang, W., Li, Y., and Tang, Y. (2008)
Engineered biosynthesis of bacterial
aromatic polyketides in Escherichia
coli. Proc. Natl. Acad. Sci. U.S.A., 105,
20683–20688.
Watanabe, K., Hotta, K., Nakaya, M.,
Praseuth, A.P., Wang, C.C., Inada, D.
et al. (2009) Escherichia coli allows
efficient modular incorporation of
newly isolated quinomycin biosynthetic
enzyme into echinomycin biosynthetic
pathway for rational design and synthesis of potent antibiotic unnatural
natural product. J. Am. Chem. Soc.,
131, 9347–9353.
Na, D., Kim, T.Y., and Lee, S.Y. (2010)
Construction and optimization of
synthetic pathways in metabolic engineering. Curr. Opin. Microbiol., 13,
363–370.
Westfall, P.J., Pitera, D.J., Lenihan, J.R.,
Eng, D., Woolard, F.X., Regentin, R.
et al. (2012) Production of amorphadiene in yeast, and its conversion to
dihydroartemisinic acid, precursor
to the antimalarial agent artemisinin.
Proc. Natl. Acad. Sci. U.S.A., 109,
E111–E118.
Hale, V., Keasling, J.D., Renninger, N.,
and Diagana, T.T. (2007) Microbially
derived artemisinin: a biotechnology
solution to the global problem of access
to affordable antimalarial drugs. Am. J.
Trop. Med. Hyg., 77, 198–202.
Reyburn, H. (2010) New WHO guidelines for the treatment of malaria. Br.
Med. J., 340, c2637.
360. Kindermans, J.M., Pilloy, J., Olliaro, P.,
361.
362.
363.
364.
365.
366.
367.
368.
and Gomes, M. (2007) Ensuring sustained ACT production and reliable
artemisinin supply. Malar. J., 6, 125.
Du, J., Shao, Z., and Zhao, H. (2011)
Engineering microbial factories for synthesis of value-added products. J. Ind.
Microbiol. Biotechnol., 38, 873–890.
Jackson, B.E., Hart-Wells, E.A., and
Matsuda, S.P. (2003) Metabolic engineering to produce sesquiterpenes in
yeast. Org. Lett., 5, 1629–1632.
Newman, J.D., Marshall, J., Chang, M.,
Nowroozi, F., Paradise, E., Pitera, D.
et al. (2006) High-level production of
amorpha-4,11-diene in a two-phase
partitioning bioreactor of metabolically
engineered Escherichia coli. Biotechnol.
Bioeng., 95, 684–691.
Anthony, J.R., Anthony, L.C.,
Nowroozi, F., Kwon, G., Newman, J.D.,
and Keasling, J.D. (2009) Optimization
of the mevalonate-based isoprenoid
biosynthetic pathway in Escherichia coli
for production of the anti-malarial drug
precursor amorpha-4,11-diene. Metab.
Eng., 11, 13–19.
Chang, M.C., Eachus, R.A., Trieu, W.,
Ro, D.K., and Keasling, J.D. (2007)
Engineering Escherichia coli for production of functionalized terpenoids
using plant P450s. Nat. Chem. Biol., 3,
274–277.
Ro, D.K., Ouellet, M., Paradise, E.M.,
Burd, H., Eng, D., Paddon, C.J. et al.
(2008) Induction of multiple pleiotropic
drug resistance genes in yeast engineered to produce an increased
level of anti-malarial drug precursor, artemisinic acid. BMC Biotechnol.,
8, 83.
Baltz, R.H. (2009) Daptomycin: mechanisms of action and resistance, and
biosynthetic engineering. Curr. Opin.
Chem. Biol., 13, 144–151.
Straus, S.K. and Hancock, R.E. (2006)
Mode of action of the new antibiotic
for Gram-positive pathogens daptomycin: comparison with cationic
antimicrobial peptides and lipopeptides. Biochim. Biophys. Acta, 1758,
1215–1223.
References
369. Penn, J., Li, X., Whiting, A., Latif, M.,
370.
371.
372.
373.
374.
375.
376.
377.
378.
Gibson, T., Silva, C.J. et al. (2006) Heterologous production of daptomycin in
Streptomyces lividans. J. Ind. Microbiol.
Biotechnol., 33, 121–128.
Nguyen, K.T., He, X., Alexander, D.C.,
Li, C., Gu, J.Q., Mascio, C. et al. (2010)
Genetically engineered lipopeptide
antibiotics related to A54145 and daptomycin with improved properties.
Antimicrob. Agents Chemother., 54,
1404–1413.
Fukuda, D.S., Du Bus, R.H., Baker, P.J.,
Berry, D.M., and Mynderse, J.S. (1990)
A54145, a new lipopeptide antibiotic
complex: isolation and characterization.
J. Antibiot. (Tokyo), 43, 594–600.
Watanabe, K., Hotta, K., Praseuth, A.P.,
Koketsu, K., Migita, A., Boddy, C.N.
et al. (2006) Total biosynthesis of
antitumor nonribosomal peptides in
Escherichia coli. Nat. Chem. Biol., 2,
423–428.
Low, C.M.L., Drew, H.R., and
Waring, M.J. (1984) Sequence-specific
binding of echinomycin to DNA: evidence for conformational changes
affecting flanking sequences. Nucleic
Acids Res., 12, 4865–4879.
Foster, B., Clagett-Carr, K.,
Shoemaker, D.D., Suffness, M.,
Plowman, J., Trissel, L. et al. (1985)
Echinomycin: the first bifunctional
intercalating agent in clinical trials.
Invest. New Drugs, 3, 403–410.
Formica, J.V. and Waring, M.J. (1983)
Effect of phosphate and amino acids
on echinomycin biosynthesis by Streptomyces echinatus. Antimicrob. Agents
Chemother., 24, 735–741.
Wilke, M.S., Lovering, A.L., and
Strynadka, N.C. (2005) Beta-lactam
antibiotic resistance: a current
structural perspective. Curr. Opin.
Microbiol., 8, 525–533.
Poole, K. (2004) Resistance to betalactam antibiotics. Cell. Mol. Life Sci.,
61, 2200–2223.
Li, R. and Townsend, C.A. (2006)
Rational strain improvement for
enhanced clavulanic acid production
by genetic engineering of the glycolytic
pathway in Streptomyces clavuligerus.
Metab. Eng., 8, 240–252.
379. Gomez-Escribano, J.P., Martin, J.F.,
380.
381.
382.
383.
384.
385.
386.
Hesketh, A., Bibb, M.J., and Liras, P.
(2008) Streptomyces clavuligerus relAnull mutants overproduce clavulanic
acid and cephamycin C: negative regulation of secondary metabolism by
(p)ppGpp. Microbiology, 154, 744–755.
Medema, M.H., Alam, M.T.,
Heijne, W.H., van den Berg, M.A.,
Muller, U., Trefzer, A. et al. (2011)
Genome-wide gene expression changes
in an industrial clavulanic acid overproduction strain of Streptomyces
clavuligerus. Microb. Biotechnol., 4,
300–305.
Burg, R.W., Miller, B.M., Baker, E.E.,
Birnbaum, J., Currie, S.A., Hartman, R.
et al. (1979) Avermectins, new family
of potent anthelmintic agents: producing organism and fermentation.
Antimicrob. Agents Chemother., 15,
361–367.
Yoon, Y.J., Kim, E.S., Hwang, Y.S.,
and Choi, C.Y. (2004) Avermectin:
biochemical and molecular basis of
its biosynthesis and regulation. Appl.
Microbiol. Biotechnol., 63, 626–634.
Ikeda, H., Nonomiya, T., Usami, M.,
Ohta, T., and Omura, S. (1999) Organization of the biosynthetic gene
cluster for the polyketide anthelmintic
macrolide avermectin in Streptomyces avermitilis. Proc. Natl. Acad. Sci.
U.S.A., 96, 9509–9514.
Zhuo, Y., Zhang, W., Chen, D., Gao, H.,
Tao, J., Liu, M. et al. (2010) Reverse
biological engineering of hrdB to
enhance the production of avermectins
in an industrial strain of Streptomyces avermitilis. Proc. Natl. Acad. Sci.
U.S.A., 107, 11250–11254.
Goudie, A.C., Evans, N.A.,
Gration, K.A., Bishop, B.F.,
Gibson, S.P., Holdom, K.S.
et al. (1993) Doramectin – a
potent novel endectocide. Vet. Parasitol., 49, 5–15.
Dutton, C.J., Gibson, S.P.,
Goudie, A.C., Holdom, K.S.,
Pacey, M.S., Ruddock, J.C. et al.
(1991) Novel avermectins produced by mutational biosynthesis.
J. Antibiot. (Tokyo), 44, 357–365.
177
178
7 Synthetic Biology and Therapies for Infectious Diseases
387. Wang, J.B., Pan, H.X., and Tang, G.L.
388.
389.
390.
391.
392.
393.
394.
395.
396.
(2011) Production of doramectin by
rational engineering of the avermectin
biosynthetic pathway. Bioorg. Med.
Chem. Lett., 21, 3320–3323.
Zhang, X., Chen, Z., Li, M., Wen, Y.,
Song, Y., and Li, J. (2006) Construction of ivermectin producer by domain
swaps of avermectin polyketide synthase in Streptomyces avermitilis. Appl.
Microbiol. Biotechnol., 72, 986–994.
Omura, S. and Crump, A. (2004) The
life and times of ivermectin - a success
story. Nat. Rev. Microbiol., 2, 984–989.
Niraula, N.P., Kim, S.H., Sohng, J.K.,
and Kim, E.S. (2010) Biotechnological
doxorubicin production: pathway and
regulation engineering of strains for
enhanced production. Appl. Microbiol.
Biotechnol., 87, 1187–1194.
Malla, S., Niraula, N.P., Singh, B.,
Liou, K., and Sohng, J.K. (2010) Limitations in doxorubicin production from
Streptomyces peucetius. Microbiol. Res.,
165, 427–435.
Hutchinson, C.R. and Colombo, A.L.
(1999) Genetic engineering of doxorubicin production in Streptomyces
peucetius: a review. J. Ind. Microbiol.
Biotechnol., 23, 647–652.
Malla, S., Niraula, N.P., Liou, K., and
Sohng, J.K. (2010) Self-resistance mechanism in Streptomyces peucetius:
overexpression of drrA, drrB and
drrC for doxorubicin enhancement.
Microbiol. Res., 165, 259–267.
Scotti, C. and Hutchinson, C.R. (1996)
Enhanced antibiotic production
by manipulation of the Streptomyces peucetius dnrH and dnmT
genes involved in doxorubicin (adriamycin) biosynthesis. J. Bacteriol., 178,
7316–7321.
Lomovskaya, N., Otten, S.L.,
Doi-Katayama, Y., Fonstein, L.,
Liu, X.C., Takatsu, T. et al. (1999)
Doxorubicin overproduction in
Streptomyces peucetius: cloning and
characterization of the dnrU ketoreductase and dnrV genes and the doxA
cytochrome P-450 hydroxylase gene. J.
Bacteriol., 181, 305–318.
Loftus, P. (2011) J&J Is Short
of Cancer Drug Doxil, 29 May,
397.
398.
399.
400.
401.
402.
403.
404.
405.
2013, http://online.wsj.com/article/
SB100014240531119035549045764
60290484704816.html (accessed 19
April 2014).
McDaniel, R., Ebert-Khosla, S.,
Hopwood, D., and Khosla, C. (1993)
Engineered biosynthesis of novel
polyketides. Science, 262, 1546–1550.
Takahashi, Y. (2002) Streptomyces
avermectinius sp. nov., an avermectinproducing strain. Int. J. Syst. Evol.
Microbiol., 52, 2163–2168.
Komatsu, M., Uchiyama, T., Omura, S.,
Cane, D.E., and Ikeda, H. (2010)
Genome-minimized Streptomyces
host for the heterologous expression
of secondary metabolism. Proc. Natl.
Acad. Sci. U.S.A., 107, 2646–2651.
Gomez-Escribano, J.P. and Bibb, M.J.
(2011) Engineering Streptomyces coelicolor for heterologous expression of
secondary metabolite gene clusters.
Microb. Biotechnol., 4, 207–215.
Shima, J., Hesketh, A., Okamoto, S.,
Kawamoto, S., and Ochi, K. (1996)
Induction of actinorhodin production
by rpsL (encoding ribosomal protein
S12) mutations that confer streptomycin resistance in Streptomyces
lividans and Streptomyces coelicolor
A3(2). J. Bacteriol., 178, 7276–7284.
Hu, H., Zhang, Q., and Ochi, K. (2002)
Activation of antibiotic biosynthesis
by specified mutations in the rpoB
gene (encoding the RNA polymerase
beta subunit) of Streptomyces lividans.
J. Bacteriol., 184, 3984–3991.
Wang, J., Zhang, Y., Chen, Y.,
Lin, M., and Lin, Z. (2012) Global
regulator engineering significantly
improved Escherichia coli tolerances
toward inhibitors of lignocellulosic
hydrolysates. Biotechnol. Bioeng., 109,
3133–3142.
Coeffet-Le Gal, M.F., Thurston, L.,
Rich, P., Miao, V., and Baltz, R.H.
(2006) Complementation of daptomycin
dptA and dptD deletion mutations in
trans and production of hybrid lipopeptide antibiotics. Microbiology, 152,
2993–3001.
Doekel, S. and Marahiel, M.A. (2000)
Dipeptide formation on engineered
References
406.
407.
408.
409.
410.
411.
hybrid peptide synthetases. Chem. Biol.,
7, 373–384.
Miyata, Y., Nakamoto, H., and
Neckers, L. (2013) The therapeutic
target Hsp90 and cancer hallmarks.
Curr. Pharm. Des., 19, 347–365.
Kapur, S., Lowry, B., Yuzawa, S.,
Kenthirapalan, S., Chen, A.Y.,
Cane, D.E. et al. (2012) Reprogramming
a module of the 6-deoxyerythronolide
B synthase for iterative chain elongation. Proc. Natl. Acad. Sci. U.S.A., 109,
4110–4115.
Vosburg, D.A. and Walsh, C.T. (2005)
Natural product biosynthetic assembly lines: prospects and challenges
for reprogramming. Ernst Schering Research Foundation Workshop,
pp. 261–284.
Durfahrt, T. and Marahiel, M.A. (2005)
Functional and structural basis for targeted modification of non-ribosomal
peptide synthetases. Ernst Schering Research Foundation Workshop,
pp. 79–106.
Udwary, D.W., Merski, M., and
Townsend, C.A. (2002) A method
for prediction of the locations of linker
regions within large multifunctional
proteins, and application to a type I
polyketide synthase. J. Mol. Biol., 323,
585–598.
D’Onofrio, A., Crawford, J.M.,
Stewart, E.J., Witt, K., Gavrish, E.,
Epstein, S. et al. (2010) Siderophores
from neighboring organisms promote
412.
413.
414.
415.
416.
417.
the growth of uncultured bacteria.
Chem. Biol., 17, 254–264.
Bok, J.W. and Keller, N.P. (2004) LaeA,
a regulator of secondary metabolism
in Aspergillus spp. Eukaryot. Cell, 3,
527–535.
Challis, G.L. (2008) Mining microbial
genomes for new natural products and
biosynthetic pathways. Microbiology,
154, 1555–1569.
Heide, L. (2009) Genetic engineering of antibiotic biosynthesis for the
generation of new aminocoumarins.
Biotechnol. Adv., 27, 1006–1014.
Menzella, H.G., Tran, T.T., Carney, J.R.,
Lau-Wee, J., Galazzo, J., Reeves, C.D.
et al. (2009) Potent non-benzoquinone
ansamycin heat shock protein 90
inhibitors from genetic engineering of
Streptomyces hygroscopicus. J. Med.
Chem., 52, 1518–1521.
Gregory, M.A., Petkovic, H., Lill, R.E.,
Moss, S.J., Wilkinson, B., Gaisser, S.
et al. (2005) Mutasynthesis of
rapamycin analogues through the
manipulation of a gene governing
starter unit biosynthesis. Angew. Chem.
Int. Ed., 44, 4757–4760.
Weist, S., Kittel, C., Bischoff, D.,
Bister, B., Pfeifer, V., Nicholson, G.J.
et al. (2004) Mutasynthesis of glycopeptide antibiotics: variations of
vancomycin’s AB-ring amino acid 3,5dihydroxyphenylglycine. J. Am. Chem.
Soc., 126, 5942–5943.
179
181
8
Nano-Antimicrobials Based on Metals
Maria Chiara Sportelli, Rosaria Anna Picca, and Nicola Cioffi
8.1
Introduction
Some metals and inorganic oxides have been recognized as antibacterial agents
since ancient times; especially Ag, Cu, and ZnO have found applications in
different fields [1–3]. For centuries, silver has been in use for the treatment of
burns and chronic wounds [4, 5]. In particular, AgNO3 was used in the treatment
of ulcers in the eighteenth century [6]. The use of silver dates back to Greeks and
Romans Empires, thanks to its medical, preservative, and restorative properties.
Analogously to Ag, the first reported medical use of copper is mentioned in
the Smith Papyrus (between 2600 and 2200 B.C.) about water disinfection and
wounds treatment [7]. Copper and its complexes in general spread through different populations (e.g., Romans) to heal different infections and to keep hygiene
until the advent of antibiotics (early twentieth century). ZnO was also reported
to heal skin wounds and to inhibit Escherichia coli growth, protecting intestinal
cells [8]. However, after the introduction of penicillin, antibiotics became the
standard treatment for bacterial infections and the use of such metals reduced
significantly [6]. Unfortunately, the development of molecular antibiotics and
their misuse have rapidly led to the emergence of the “antimicrobial resistance”
(AMR) phenomenon [9], as also highlighted by the World Health Organization
(WHO) [10]. A high percentage of hospital-acquired infections are caused by
highly resistant bacteria such as methicillin-resistant Staphylococcus aureus
(MRSA) and vancomycin or multidrug-resistant enterococci Gram-negative
bacteria [10]. As a result, microbial contamination and pathogenic infections
represent major threats to public health and environment in several fields,
ranging from food packaging [11] to textile [12] and paint [13] industries, just to
cite a few. To this end, nanotechnology has been exploited to create engineered
nanostructures /nanocomposites (NSs) based on well-known antimicrobial materials (e.g., silver, copper, zinc oxide), generally classified as nano-antimicrobials
[14], with improved bioactive performances and controlled toxicity to human
beings. It has already been accepted by the scientific community that nanostructuration of such materials is helpful in generating antimicrobial agents with
Novel Antimicrobial Agents and Strategies, First Edition.
Edited by David A. Phoenix, Frederick Harris and Sarah R. Dennison.
© 2015 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2015 by Wiley-VCH Verlag GmbH & Co. KGaA.
182
8 Nano-Antimicrobials Based on Metals
improved performance, thanks to a synergic contribution from both size and high
surface-to-volume ratio [4, 15, 16]. Moreover, the possibility to conjugate the surface of the NSs or to combine them with other materials (e.g., polymers, natural
and synthetic fibers, clays) allows to achieve nano-antimicrobials with tunable
properties in terms of efficient bioactivity against the targeted microorganisms
and limited, if any, toxicity toward human cells.
In this chapter, we present several examples of nanomaterials based on
three of the main inorganic materials with known antimicrobial action (i.e.,
silver – Section 8.2, copper – Section 8.3, and zinc oxide – Section 8.4). Considering the huge number of publications on this topic, we have focused our
attention on the typical synthetic strategies involved in recent years in the development of the presented nano-antimicrobials. Some details on characterization
and applications of the proposed antimicrobial agents are also shown. Some
considerations on the design and the development of the next generation of
nano-antimicrobials are provided in the conclusive paragraph.
8.2
Silver Nano-antimicrobials
Much attention has been paid to metal nanoparticles (NPs) and NSs because of
their fascinating chemical and physical properties. Silver is a noble metal, characterized by high thermal stability, low toxicity, and high antibacterial activity
[17]. A myriad of shapes of Ag NSs have been already synthesized: in fact, it is
known that the properties of metal NSs strongly depend on the involved sizes and
shapes [18]. In spite of its fundamental and technological importance (as well as a
long history of study), the challenge to synthetically and systematically control the
shape of metal nanostructures (Me NSs) is not completely addressed [19]. Hence,
a generic method for the preparation of morphologically defined Me NSs with
related specific applications is still in development. In the following, we are going
to synthetically list and describe the most common preparation methods of Ag
NPs, the most diffused strategies to characterize them, and their main applications, especially in the antimicrobial field.
8.2.1
Synthesis of Silver Nanostructures
A variety of synthesis routes have been developed for the preparation of Ag NSs,
which can be mostly divided into a few large categories: physical, chemical, and
electrochemical approaches [20]. As can be easily guessed, each method is characterized by its peculiar advantages and drawbacks; the choice of the most appropriate synthetic route is mainly dictated from the material’s final use. Here, we
provide an overview of the most common techniques to synthesize Ag NSs and
nanocomposites.
8.2
Silver Nano-antimicrobials
8.2.1.1 Physical Approaches
One of the first physical methods employed for the production of silver NPs was
a mechanical approach: the bulk metal underwent homogenization, grinding, or
milling processes to generate small-sized Ag powders [21]. This technique is characterized by poor morphological control, although still widely used in the industrial or commercial production of Ag nanopowders [22]. Recently, one of the most
used physical processes is the so-called evaporation–condensation method [23].
During this process, performed in a tube furnace, solid bulk Ag is heated in a
small temperature-regulated crucible to produce vapors, which are then rapidly
quenched on a cold surface (i.e., liquid nitrogen, metallic plates, etc.) in an inert
atmosphere of argon gas [24]. A good shape control can be achieved this way [20],
although many parameters have been found to affect the final material, such as
temperature, evaporation power and rate, and argon pressure in the quenching
phase [24, 25]. Other physical methods are based on the reduction of a silver saline
precursor, catalyzed by ultrasounds [26–28], microwaves [29–32], UV light [33,
34], and so on. This technique is mostly used to obtain metal–polymer nanocomposites, avoiding complications related to in situ polymerization processes [35].
Immobilization of Ag nanophases onto polymeric materials has been recently
reported by many research groups [36, 37]. Interestingly, Perkas et al. [38] modified industrial nylon 6,6 fibers with Ag NPs, capable to exert high antimicrobial
activity; this innovative material paved the way as one of the first examples of reallife-applicable products.
8.2.1.2 Laser Ablation in Liquids
In the past years, laser ablation synthesis in solution (LASiS) emerged as a reliable
alternative to traditional methods for obtaining noble metal NPs. LASiS is a
“green” technique for the synthesis of stable metal NPs in water or in organic
solvents, which does not need stabilizing molecules or other chemicals [39]. The
process is performed by irradiating a bulk metallic target immersed in solvents
using intense laser light [40]. To the best of our knowledge, one of the first reports
about LASiS of Ag NPs is the one from R.A. Ganeev et al. [41], who reported
the synthesis of Ag NPs in the absence of stabilizing agents and with different
solvents obtaining, however, small and well dispersed NPs. The use of water [42]
or organic solvents [43] has been reported to affect the size and morphology
of particles, as described in detail in some interesting studies [41, 44]. Laser
parameters, such as fluence, power, repetition rate, and wavelength, were found
to be useful tools to obtain good morphological control for synthesized NSs
[45]. The chance to directly disperse Ag NPs in aqueous solutions of polymeric
matrices was also envisaged, with the advantage of obtaining more stable and
dispersed colloids [46–48]. Natural and biodegradable polymers such as chitosan
(CS) were also used [49, 50].
8.2.1.3 Chemical Approaches
The synthesis of NSs by means of the chemical reduction approach is one of
the most used methodologies for shape-controlled metal NPs [51]. This is a
well-known technique: the principle was first reported in 1914 [52]; since then,
183
184
8 Nano-Antimicrobials Based on Metals
many research papers about chemical reduction of silver ions have been published
[53–55]. Both aqueous [53, 56] and organic [57] solvents can be used, often
containing surfactants or polymers. The use of these substances is fundamental
to achieve both colloidal stabilization and size control [58–61]. The reduction of
silver ion takes place with the help of reducing agents; commonly used reductants
are borohydride [62], citrate [63], ascorbate [64], and molecular hydrogen [65].
The use of a strong reductant such as borohydride results in small particles that are
somewhat monodispersed, while the use of citrate or ascorbate, which are weaker
reductants, results in a slower reduction rate and narrower size distribution [66].
The most important drawbacks of this method are related to agglomeration of
silver NPs during the synthesis [67–69]. To overcome the limitation of chemical
reduction and prevent particle aggregation, the reverse microemulsion (reverse
micelle) method was introduced to obtain uniform and size-controlled NPs
[70]. This method has the obvious advantage of synthesizing NPs with specific
diameter and morphology [57], since nucleation and growth are restricted within
the cores of the micelles [71]. The droplet dimension can be modulated by various
parameters, in particular the molar ratio of water and the surfactant [72].
8.2.1.4 Biological and Biotechnological Approaches
The chemical method allows the preparation of uniform and size-controllable
NPs; however, highly deleterious organic solvents with potential risks for environment and biological hazards are frequently employed. An increasing awareness
toward green chemistry and biological processes is developing in the recent
years, leading to the set-up of an environment-friendly approach for the synthesis
of NPs [73]. Unlike other physical and chemical processes, biosynthesis of NPs
is a completely eco-friendly approach [74, 75]. In 2002, a surprising discovery
regarding protein and amino acids involved in biomineralization processes paved
the way to the use of biomolecules as reducing agents of silver ions [76] for the
synthesis of Ag NSs. Natural extracts from fungi and spores are among the first
used for the synthesis of Ag NPs [77–79], acting as stabilizing and reducing
agents at the same time. Moreover, they are cheap and easily available substances
[80]. Some microorganisms have been explored as biofactories for the synthesis of
Ag NPs, thanks to the presence of reductase and other enzymes on the bacterial
cell wall [81–85]. Finally, natural extracts from leaves [86–90], flowers [91], and
fruits [92, 93], containing proteins, enzymes, and polysaccharides, were found to
give good results in the chemical synthesis of Ag NSs, although morphological
control is quite difficult to achieve.
8.2.1.5 Electrochemical Approaches
Electrochemical techniques are very interesting, since they allow obtaining metal
NPs with high purity and crystallinity, controlling particles’ size by means of few
experimental parameters [94]. Reetz and Helbig [95] were the first to develop
an electrochemical process for the synthesis of colloidal transition metal NPs.
In detail, a bulk metal sheet was anodically dissolved in an organic solution
(a mixture of acetonitrile (ACN) and tetrahydrofuran) and the intermediate
8.2
Silver Nano-antimicrobials
metal ions were reduced at the cathode, giving rise to metal particles stabilized
by tetraalkylammonium salts. This strategy was first successfully adopted for the
electrochemical synthesis of Ag NPs by Rodríguez-Sánchez and coworkers [96], in
pure ACN containing tetrabutylammonium bromide. The authors found Reetz’s
mixture unsuitable for Ag synthesis because the presence of tetrahydrofurane
induced particles aggregation. Following a similar approach, Ag Nps were also
obtained in ethanol [97], hexane [98], and N,N-dimethylformamide (DMF) [99].
The use of a zwitterionic surfactant in ACN, instead of an ammonium salt, was
found to be susceptible to the growth of branched hierarchical NSs (specifically
nanofractals NFs) [100]. Despite many advantages, such as size tuning and
inexpensive experimental set-up, this method has intrinsic limitations: the use
of potentially dangerous organic solvents – these substances may limit real-life
applications of the as-prepared colloids [101]. This is why an important current
issue in the electrochemical synthesis of metal NPs regards the preparation of
aqueous and long-lived colloids [94, 101]. V. ten Kortenaar and coworkers [102]
found that long-lived, subnanometer-sized silver oxide clusters could be prepared
by anodic dispersion of a silver electrode in basic aqueous solutions (pH 10.5–12),
free of stabilizing polymers, upon the application of a high DC voltage (65 V)
between two silver electrodes. Synthetic parameters made this process energy and
time consuming; this inconvenience can be overcome by employing electrochemical reduction of a silver salt in aqueous solution [103]. Electrochemical reduction
of silver ions involves two competitive processes: the formation of Ag NPs
and the deposition of a silver film onto the cathode surface. Usually, this second
occurrence dominates over the first one, hence the need for the presence of a good
stabilizer [104]. Typical stabilizing agents are both synthetic and natural polymers,
ranging from poly(N-vinylpyrrolidone) (PVP) [103, 104], polyphenylpyrrole (PPP)
[105], and polyethylene-glycol (PEG) [106, 107], to CS [108]. Supported metal NSs,
that is, electrochemically deposited onto different substrates, are one of the most
developed and important subjects of study in electrochemistry. To the best of our
knowledge, one of the first examples of supported Ag NSs was provided by Zoval
et al. [109], who employed a potentiostatic pulsed method to deposit ordered
silver NPs onto atomically flat graphite. The intrinsic flexibility of this technique
is responsible for its rapid technological development: many different substrates
such as metal sheets [110, 111], graphene and graphite [112–115], and polymeric
films [116], were used to deposit variously shaped NSs. A direct embedding of NPs
in polymeric matrices can be also performed in a one-pot approach, using this
technique: recently Zhitomirsky and Hashambhoy [117] proposed the deposition
of inorganic Ag NPs contextually with a CS film, paving the way for a fast and controllable preparation of cheap and antimicrobial films for biomedical applications.
8.2.2
Characterization of Silver Nanostructures
Characterization of nanosized materials is one of the challenges of the last
decades, since they have special properties stemming from their nanoscale
185
186
8 Nano-Antimicrobials Based on Metals
dimensions. Because of their high surface/volume ratio, their chemical and
physical properties are strictly correlated to the so-called quantum mechanical
effects [118]. The principal properties that have to be determined are dimensions,
morphology, crystallinity, and chemical composition. UV–Vis spectroscopy is
a useful tool to quickly obtain concentration and mean diameter of a Me–NP
colloid [119]. A tight relation between particles dimension and their plasmon
resonant spectral response is known to exist [120], supported by sophisticated
and detailed quantum mechanics theories [121, 122]. A typical UV–Vis spectrum of Ag NPs shows a well-defined surface plasmon band centered at around
420 nm [123]. This plasmon band is broad, with an absorption tail in the longer
wavelengths, which could be in practice due to the size distribution of the particle
[124]. Dynamic light scattering (DLS) is a convenient technique, frequently used
to determine the effective size and size distribution of silver particles suspended in
solution. An average effective size, or hydrodynamic diameter, is calculated for Ag
particles using the Stokes–Einstein relation between the diffusion coefficient and
particle diameter [125]. In the last decade, the development and the diffusion of
the electron microscopies, and particularly of transmission electron microscopy
(TEM) [126, 127], have made morphological characterization of NPs direct and
easy to understand. High-resolution TEM micrographs of Ag NSs are frequently
found in the literature [28, 128–132]. Although scanning electron microscopy
(SEM) and atomic force microscopy (AFM) can provide high-resolution images
of nanometric crystallites surfaces, they are unlikely to clearly resolve the atomic
lattices of NPs because of the unavoidable wobbling of nanocrystals under
the scanning probe [133]. SEM technique is particularly useful in the study of
supported NSs, in contrast to TEM, which mostly requires colloidal dispersions
or ultrathin sections. Hence, SEM is especially used for the morphological
characterization of Ag nanostructured surfaces, prepared by means of electrochemical [110–112, 115, 116] and physical [23, 24, 134, 135] methods. X-ray
diffraction (XRD) or selected-area electron diffraction (SAED) measurements are
necessary to assess the crystalline nature of the NSs. Nanometer-sized crystalline
materials have been proposed to represent a solid-state structure that exhibits
neither long-range order (like crystals) nor short-range order (like glasses).
Structurally, these materials consist of the following two components, the volume
fraction of which is about 50% each: a crystalline component, formed by all atoms
located in the lattice of the crystallites, and an interfacial component comprising
the atoms situated in the interfaces [136]. It is this interfacial component that
was proposed to exhibit an atomic arrangement without short or long range,
giving rise to peculiar features in XRD patterns [137]. Zero-valent nanosized
silver crystals present an XRD pattern related to a face-centered cubic (fcc) unit
cell, generally representing the 111, 200, 220, 311, and 222 crystal planes due
to Bragg’s reflections [138]. A mathematical analysis of these Bragg’s peaks is
undertaken to calculate the crystallite size using the Scherrer formula [139].
X-ray photoelectron spectroscopy (XPS) is an excellent tool for the investigation
of the electronic structures and chemical bonding of materials; nevertheless, its
application to the study of colloidal transition metal NPs has so far been limited,
8.2
Silver Nano-antimicrobials
partly because limited chemical speciation can be achieved from photoelectronic
binding energies (BEs) for most of the transition metals [140]. Moreover, the BEs
of the Ag 3d5/2 core electrons obtained from NSs are not significantly different
from those of bulk materials [141]. Anyway, studies regarding the size effect can
be performed using Auger emissions and mathematical calculation of modified
Auger parameters (α′ ) [100, 142].
8.2.3
Applications of Silver Nanostructures
Silver NPs are being used in numerous technologies and incorporated into a wide
array of consumer products that take advantage of their desirable optical, conductive, and antibacterial properties. Silver NPs are used in conductive inks and
integrated into composites to enhance thermal and electrical conductivity. Conductive films that are both stretchable and flexible found applications in electronic
devices [143], sensors [144], actuators [145], and so on. A substantial amount
of research has been carried out on conductive polymer composites containing
Ag NPs, also in combination with carbon nanotubes (CNTs) and other nanosized metals [146–148]. Silver NPs are used to efficiently harvest light and for
enhanced optical spectroscopies including surface-enhanced Raman scattering
(SERS) [149]. The technique of SERS is a particularly sensitive and selective analytical tool for the detection of low concentrations of analytes adsorbed on noble
metal NSs [150]. The morphology of the metallic NSs is a primary factor determining the magnitude of signal enhancement and sensitivity of detection. Ag nanorod
(NR) arrays [151], nanoflakes [152], nanocubes [153], and NPs [154], deposited
onto proper substrates, are the most used Ag NSs because, once deposited, they
can produce surfaces with controllable and reproducible roughness [155]. The
main application of Ag NPs is, however, as antimicrobial agent [14]: for decades,
silver-based nanocomposites have been used extensively as antimicrobial agents
in a number of areas including medical, pharmaceutical, textile, industrial food
storage, and environmental applications [156]. This topic is discussed in detail in
the following paragraph.
8.2.3.1 Silver-Based Nano-antimicrobials
Silver has been known for antibacterial activity since the times of ancient
Greece. Currently, the investigation of this phenomenon has regained importance owing to the increase of bacterial resistance to antibiotics, caused by
their overuse [9]. Ag NPs exhibit very strong antibacterial activity against both
Gram-positive and Gram-negative bacteria (Figure 8.1), including multiresistant
strains [157].
Contrary to the bactericide effects of ionic silver, the antimicrobial activity of
colloid silver particles is influenced by many parameters (Table 8.1), especially
by their dimensions: the smaller the particles, the greater the antimicrobial
effect [158].
187
188
8 Nano-Antimicrobials Based on Metals
B. thuringiensis
(Gram-positive)
Inhibition zones (mm)
Chitosan
24
6h
27
12 h
28
18 h
30
24 h
35
Chitosan
26
6h
30
12 h
32
18 h
34
24 h
35
P. aeruginosa
(Gram-negative)
Inhibition zones (mm)
Figure 8.1 Comparison between the bacterial inhibition zones (mm) of chitosan and the
chitosan-Ag NPs developed at different times (6, 12, 18, and 24 h). (Reproduced from Ref.
[108], Copyright (2012), with permission from Elsevier.)
Table 8.1 Factors affecting Ag NSs toxicity.
Factor
Tendency
Possible explanation
Particle size
Smaller particle size
tends to enhance
antibacterial properties
Particle stability
Higher stability
produces higher
antibacterial properties
Particle Shape
Particles with shapes
containing more <111>
facets like triangular
particles tend to have
strongest antibacterial
properties
Depending in a case to
case base
As size decreases, there is larger
number of atoms on the surface
available to interact with bacteria or to
release a higher amount of silver ions
Non-stable nanoparticles will tend to
form aggregates; thus surface area will
be reduced and the density of atoms
available on the surface will be lower
[160–162]
<111> facets would contain larger
atom densities; thus more atoms
available for interaction [163]
Water chemistry
Since water chemistry affects particle
suspension/solubility, particle size
distribution, as well as bacterial ability
to face environmental stresses, water
chemistry will affect the interaction
between nano-scaled silver and
bacteria thus influencing the resulting
toxicity
Reproduced from Ref. [159], Table 4, Copyright (2010), with kind permission from Springer
Science and Business Media.
8.2
Silver Nano-antimicrobials
Enhanced antimicrobial activities have been reported in silver NPs modified by
surfactants and in polymers [4]. This is probably due to the modulated ion release
that can be obtained by embedding Ag NSs in a proper matrix. Examples are
widely reported in literature: commonly used polymers are CS [128], polyvinyl
alcohol [129], polylactic acid (PLA) [164], and poly-(vinyl-methyl ketone) [130].
Ag colloidal NSs can also be used to drench wound dressings [128] (Figure 8.2),
to modify textiles [165], and so on.
(a)
(b)
(c)
(d)
(e)
(f)
Figure 8.2 TEM micrographs of fibers
containing Ag-NPs for wound dressing
applications; (a–d) micrographs show
individual fibers loaded with Ag-NPs;
(e) nanofiber mat top-view; and (f ) crosssection nanofibers. (Reproduced from Ref.
[128] Copyright (2013), with permission from
Elsevier.)
189
190
8 Nano-Antimicrobials Based on Metals
The use of proper surfactants, such as tetraalkylammonium salts, which are
strong disinfectants, leads to a synergic antimicrobial effect deriving from both
stabilizers and metal NPs [130]. Although silver has been used as a biocide for
hundreds of years, the antimicrobial mechanism of silver has not been fully elucidated [159]. A variety of mechanisms may be involved in the antimicrobial activity
of silver against a broad spectrum of organisms. Some of the commonly accepted
mechanisms include silver–amino acid interaction [166], silver–DNA interaction
[167], generation of reactive oxygen species (ROS) [168], and direct cell membrane
damage, such as cell internalization [157].
8.3
Copper Nano-antimicrobials
Copper species are well-known bioactive agents, which induce an oxidative stress
on proteins, nucleic acids, and lipids by producing reactive hydroxyl radicals with
consequent damages to membrane and other cell organelles [13, 169, 170].
Despite these considerations, copper-based nanomaterials have been less
studied in comparison to nanostructured silver [171, 172]. On the other hand,
in some cases copper nanoparticles (Cu NPs) have demonstrated superior
bioactivity [173]. It should be pointed out that either elemental copper or copper
oxides have been proposed as nanostructured antimicrobial agents [174, 175];
however in the following paragraphs we review mainly the case of Cu(0) NPs.
Firstly, we focus on the methods that are generally used for the synthesis of
copper-based nano-antimicrobials. A brief overview of the main applications of
Cu NPs as bioactive material is also given.
8.3.1
Preparation and Applications of Antimicrobial Cu Nanostructures
In this section, we summarize shortly some approaches to the development of
copper nano-antimicrobials. All the cited examples were chosen because the
antimicrobial application was clearly addressed, considering their novelty and/or
importance. It is beyond the scope of this chapter to give an extensive review of
all the possible methods for the synthesis of Cu NSs, which have been reviewed
elsewhere [176]. For clarity reasons, the reported approaches are classified in
physical, wet-chemical, and other alternative approaches for the development of
copper nano-antimicrobials, such as electrochemical syntheses, laser ablation in
liquid (LAL), and biological syntheses.
8.3.1.1 Physical Methods
These techniques are generally related to the single-step preparation of
nanocomposites based on the formation of Cu nanoclusters in (or onto) an
8.3
Copper Nano-antimicrobials
organic/inorganic coating. In many cases, no stabilization of Cu NPs occurs, leading to the formation of copper oxides shell. Plasma and sputtering approaches
represent the main methods proposed in literature for the preparation of
antimicrobial nanocomposites. Plasma immersion ion implantation (PIII) allows
the penetration (and incorporation) of copper in the first layers of polymeric films
(e.g., polyethylene, PE), thus conferring antimicrobial properties to the polymer
without causing dramatic damages to its structure [177–180]. In particular,
Zhang and coworkers [180] showed how a dual-plasma system (Cu/N2 ) was
more efficient in the regulation of the copper release rate improving the longterm bioactivity of low density PE against E. coli and S. aureus. Other works
employing unusual plasma approaches were also reported. A procedure based on
a thermionic vacuum arc plasma (TVA process) for the deposition of Cu or Ag
nanofilms on stainless steel and polymer spheres to treat nosocomial bacterial
species was presented in [181]. An industrial method using thermal plasma
(Tesima™) was employed to prepare several antimicrobial (e.g., CuO, Cu, Cu2 O,
ZnO, and Ag) nanopowders in a continuous gas phase process [182].
Among sputtering techniques, magnetron sputtering and ion beam sputtering
(IBS) are more frequently applied. Direct current (pulsed) magnetron sputtering
has been proposed to modify TiN hard and wear-resistant coatings [183] as
well as polyester surfaces [184]. In particular, in [183] co-deposition from two
high purity targets, keeping constant the power for Ti and varying that for Cu to
achieve different copper loadings in the nanocomposites in Ar–N2 atmosphere is
reported. These nanocoatings showed very good performances against S. aureus.
Cu-sputtered polyester surfaces exerted excellent bioactivity against the same
methicillin-resistant bacteria [184]. Advantages of sputtering methods lie also
in the deposition of mechanically durable and relatively well-adhered Cu layers
directly on fabrics (e.g., cotton) [185].
Copper-doped diamond-like carbon films (useful to develop surgical devices)
have also been produced by radio-frequency magnetron sputtering, also combining plasma-enhanced chemical vapor deposition (CVD) under various Ar/CH4
gas mixtures [186, 187].
Alternatively, nanocomposites incorporating copper nanoclusters (with a
mixed CuO/CuF2 nature) in a fluoropolymeric matrix were prepared by simultaneous IBS of pure Cu and polytetrafluoroethylene (PTFE) targets. Copper
content was adjusted by simply tuning the growth rates of the two materials
[188]. Obviously, CVD technique was also applied to the development of
Cu-based antimicrobial agents (in the form of thin films) as reported in [189],
with particular focus on low cost ambient pressure approaches [190].
Recently, a strategy (the mixed melt method) based on the use of commercially
available Cu (or CuO) NPs and polymer (polypropylene) was proposed by Palza
and coworkers [191, 192] to produce novel bioactive nanocomposites. They found
that NP content and NP composition (CuO compared to Cu) could greatly affect
the release kinetics (and thus the bioactivity). Interestingly, despite other works
considering only the first layers effective for Cu2+ release, they could not quantify
the Cu surface concentration. However, they correlated the high bioactivity of the
191
192
8 Nano-Antimicrobials Based on Metals
nanocomposites to the diffusion of water through the percolation network of the
polymer, which allowed the “corrosion” of Cu (CuO) NPs.
A procedure based on a flame aerosol reactor was described to synthesize
Cu-doped TiO2 NPs using titanium tetra-isopropoxide and copper (II) ethyl hexanoate as precursors [193]. Some authors employed a mixed chemical–physical
method – the molecular level mixing method (MLM), coupled with high energy
ball milling (BM) – to fabricate Cu/CNTs nanocomposites [194, 195]. The
electrospinning process was employed for the synthesis of Cu NPs-polyurethane
nanofibers without adding any chemical precursors [196]. Another approach was
introduced by Karlsson et al. [197], who synthesized Cu NPs by the electrical wire
explosion technique [198], though this method generally leads to the formation
of aggregates of Cu NPs.
8.3.1.2 Wet-Chemical Methods
When Cu NPs are needed as antimicrobials or modifiers for advanced bioactive
composites, approaches based on the chemical reduction of Cu(II) salts (chloride
or sulfate) are the most widely applied. In general, sodium borohydride [199–208]
and hydrazine [208–213] are used as reducing agents. The reaction is carried out
under alkaline conditions (NaOH addition, pH ∼ 11) and in an inert atmosphere
to prevent fast oxidation of Cu(0) NPs [201, 208]. Instead of nitrogen, some works
reported the use of some “protecting/stabilizing” agents to overcome this issue
and avoid aggregation of NPs. For example, poly(vinylalcohol) [203], CS [210, 212],
ethylenediamine tetraacetic acid (EDTA) [194], or gelatin [211] were proposed
with this aim. It should be pointed out however that, even in suitable conditions,
this approach leads generally to the formation of Cu(0) core – Cu(I) shell NPs as
also stated in some works [204]. The main advantage of this method resides in
the possibility to modify in a single step some natural fibers (e.g., cellulose), which
were considered effective in protecting NPs and avoiding their direct release in
aqueous media [200, 202, 205–208]. In fact, the possible leaching of NPs may
pose critical issues about their safe use [16] and it is then important to get a
controlled ion release [188, 200]. As a result, many examples about the preparation of Cu NP/fabric composites were reported in the literature, as also recently
reviewed [214]. A very elegant strategy was described by Cady et al. [200], who
combined a layer-by-layer (LBL) electrostatic self-assembly process to make cellulose chelating Cu2+ ions for further chemical reduction to Cu NPs. In particular,
natural cotton fiber was converted into anionic carboxymethyl cotton fiber with
chloroacetic acid (Figure 8.3). The as-prepared materials showed good bioactivity
against Acinetobacter baumannii, without toxicity against mammalian cells.
Other fibers were also treated by the reduction method, such as nylon
[215] or bamboo rayon [216]. In particular, in [215] Cu NPs were synthesized
on/within polyamide chains of nylon fabric using ascorbic acid as natural
antioxidant/reducing agent and cetyl-trimethylammonium bromide (CTAB) as
capping/assembling agent. It was also proposed to synthesize Cu NPs by chemical
reduction with ascorbic acid and then transfer them to sodium alginate-treated
cotton by microwave heating [217].
8.3
Copper Nano-antimicrobials
193
OH
OH
O
HO
O
O
O
OH
OH
OH O
O
HO
O
Cu
OH
n
O
O
CI
−
O Na+
OH
O
HO
O
O
NaOH
n
O
O
O
CuSO4
−
O
HO
NaBH4
O HO
O HO
HO
O
O
O
O
O
Anionic
cotton
fiber
O
O
O
Cu
O
O
O
OH
O
HO
O
O
O
HO
n
O
Nanoparticlecotton complex
O
O
Figure 8.3 Scheme of the production of Cu NPs-modified cotton fibers. (Reproduced from
Ref. [200], Scheme 1, Copyright (2011), with permission from Wiley-VCH Verlag GmbH & Co.
KGaA, Weinheim.)
Controlled-release Cu-based nano-antimicrobials were also produced by
chemical reduction using clays as embedding matrices [209, 213, 218, 219].
These materials possess high surface area, chemical inertness, and nontoxicity,
making them promising choices in the biomedical field. In some cases, NPs were
prepared before mixing with clays [213] whereas their ion-exchange properties
were employed to capture cupric ions for their further reduction directly inside
the matrix [218].
Chemical reduction may also be combined with physical approaches as reported
in [194], where high-energy BM follows the reduction step with NaBH4 to fabricate Cu/CNT nanocomposites as potential antimicrobial coatings on catheter
linings, hospital trays, and food containers. A similar combination was also presented by Kamrupi and Dolui [220] to fabricate Cu NPs/polystyrene nanocomposites. Colloidal Cu NPs were prepared by NaBH4 reduction of CuCl2 into styrene
droplets for its following polymerization using water in supercritical carbon dioxide (water-in-sc-CO2 ) at 70 ∘ C.
The sol–gel approach was also widely employed, especially in works regarding
the preparation of bioactive glasses [221–223] and silica-based materials [224,
225]. Bioactive glasses find applications in different clinical fields (e.g., dental
devices), thanks to their high surface area and drug delivery properties [222].
In a simple manner, the precursors of glass (tetraethyl orthosilicate, triethyl
phosphate, and calcium nitrate) and CuO (copper(II) salt) were used to carry
out a sol–gel synthesis. The gels could undergo thermal treatments to obtain
different composites (e.g., aged gels at 60 ∘ C, gel glasses at 600 ∘ C, and glass
ceramics at 1050 ∘ C) [221, 222], which showed the presence of Cu2+ species,
homogeneously distributed in the glass matrix, with low copper release (5 ppm)
[221]. Alternatively, the same group proposed the so-called “impregnation”
process based on the soaking into a solution of Cu(II) precursor of sol–gel
prepared glasses (treated at 600 ∘ C). The modified glasses were then aged at
60 ∘ C followed by heat treatment at 700 ∘ C in N2 /H2 atmosphere, resulting in the
reduction of Cu2+ to Cu(0) NPs [221, 222]. Cu NPs were mainly distributed on
194
8 Nano-Antimicrobials Based on Metals
the glass surface, thus exerting higher copper release (65 ppm) [221]. The possible
application of Cu NPs as effective antifungi additives for architectural paints and
impregnates was demonstrated by Zielecka and collaborators [224] (Figure 8.4).
They described the sol–gel synthesis of novel silica nanospheres immobilizing
copper NPs (or Ag NPs). The most important advantage of these additives lies in
the stability of copper NPs as well as in the harmlessness to human beings.
Other protocols based on wet-chemical methods were also reported, for
example, polyol reduction [226, 227], hydrothermal synthesis [228], microwave
heating in the presence [229, 230] or absence of stabilizers (e.g., CS or starch)
[231], or UV photoreduction [232].
Despite the diffusion and versatility of wet-chemical approaches, some “alternative” methodologies to produce efficient nano-antimicrobials were also considered by researchers for several reasons (e.g., reducing costs, environmental
sustainability, controlled release, chemical speciation, and low toxicity of the final
products). In the following section, three important approaches are reviewed as
valid alternatives to purely physical and chemical ones.
F Cu 0,1
2
F Cu 0,5
0
N
5
Figure 8.4 Activity of Cu NPs against algae after 48 h incubation. (Reproduced from Ref.
[224], Copyright (2011), with permission from Elsevier.)
8.3
Copper Nano-antimicrobials
8.3.1.3 Electrochemical Syntheses
Cathodic
reduction
Anodic
dissolution
Workin
electro g
de
(Cu)
Reference
Electrochemical routes to the synthesis of nanomaterials are generally considered
highly versatile and up scalable for industrial applications. In particular, they may
allow the fine-tuning of size and shape control by operating either on chemical
(pH, temperature, electrolytes, and solvents) or electrochemical (current density
and potential) parameters [95]. Such features were considered appealing also in
the field of nanostructured antimicrobials, as demonstrated by the publications on
this topic [130, 233–238]. In particular, our group developed a suitable procedure
to synthesize tetra-alkylammonium-stabilized Cu NPs (as well as Ag NPs) based
on the “sacrificial anode electrolysis” [95] in organic solvents [130, 233–236]. Very
stable colloidal dispersion of Cu NPs (spherical-shaped) was prepared by applying
a constant potential (chosen accordingly) and using a three-electrode conventional cell. In particular, a high purity copper sheet, acting as working electrode,
was used as the source of cupric ions. The presence of tetraoctylammonium chloride (considered one of the best performing species) allowed stabilization of the
Cu(0) nuclei formed at the cathode (Pt sheet) and control of their growth in solution as NPs (Figure 8.5).
Colloidal Cu NPs were effectively dispersed in commercial polymers (e.g.,
polyvinylmethylketone, polyvinylchloride) [130, 233–235] and successively
deposited as thin nanocomposite films by a simple spin coating procedure on
glass slides to test their bioactivity against different microorganisms. It was
shown how ion release (as studied by atomic absorption spectroscopy) could
be quantitatively correlated to their good antimicrobial performance. Later,
a similar approach was also employed to mix Cu NPs with a silicon-based
product, commonly used as a water-repellent/consolidant, to obtain bioactive
nanocoatings to be applied on stone artworks [236].
Counte
r
electro
de
(Pt)
+
N
+
N
+
+
N
+
N
N
+
N
Cu(0)
+
N
Nanoparticle
growth
+
N
Core-shell
nanoparticle
Figure 8.5 Electrochemical synthesis of Cu NPs by sacrificial anode electrolysis.
195
8 Nano-Antimicrobials Based on Metals
Colloid
500 nm Silicone
resin
Ethyl acetate
200 nm
Ethyl acetate
500 nm
Laser ablation
of silver and
copper in ethyl
acetate and
hexane
n-hexane
200 nm
Mixing with
silicone resin,
evaporating
solvent under
vacuum
n-hexane
Silicone
polymer
200 nm
Mixing with curing
Ethyl acetate
agent,
polymerisation, ΔT
200 nm
Degree of
dispersion (%)
196
100
80
60
40
20
0
Ethyl
n-
acetate hexane
n-hexane
Figure 8.6 Laser ablation synthesis of Ag and Cu NPs in the presence of silicone matrix.
(Reproduced from Ref. [243], Figure 4, Copyright (2010), with permission from Wiley-VCH
Verlag GmbH & Co. KGaA, Weinheim.)
Recently, CuO NPs were electrosynthesized at 12 V under ambient atmosphere
by a template-free, low-temperature approach [238]. Copper bars were used
as both cathode and anode and the electrolytic solution consisted of aqueous
KCl/hydrogen peroxide. H2 O2 was used as oxidant in this process. In the absence
of stabilizers, the electrochemical deposition of copper (in the presence hydrogen
peroxide) led to CuO/Cu(OH)2 mixed NPs. A calcination step at 150 ∘ C allowed
complete conversion to CuO NPs. CuO NPs showed fast deactivation kinetics
against E. coli as well as other Gram-positive (Bacillus subtilis, S. aureus) and
-negative (Salmonella typhi) bacteria.
8.3.1.4 Laser Ablation in Liquids
In the last years, LAL has been receiving increasing attention thanks to the
possibility of physically reducing a solid target to nano-colloids, without the
use of chemical precursors [39, 239, 240]. Other advantages are represented by
very short reaction times, mild temperature conditions, no by-products, and
outstanding purity. Many researchers have envisaged the possibility to apply
this technology to prepare nano-antimicrobial agents [241–247]. Since their
early work on the generation of Ag NPs in ethylacetate doped with silicone resin
[241], Hahn and collaborators [242–244] focused on the LA synthesis of Cu NPs
embedded in a polymer matrix to get a controlled ion release. They showed how
targeting the “therapeutic window” might allow inhibition of the proliferation of
some cells (e.g., fibroblasts) without being toxic for others (e.g., neuronal cells).
This approach was proposed to fabricate Cu NPs-treated cochlear implant (CI)
electrodes. In particular, they prepared laser-ablated stable colloidal Cu NPs in
n-hexane doped with 1% of silicone. n-Hexane behaved better than ethyl acetate
leading to a homogenous distribution of NPs in the silicone matrix (Figure 8.6).
LAL-synthesized copper antimicrobials were also considered in food packaging
applications [245–247]. PLA materials filled with Cu NPs were prepared by either
one- or two-step protocol in our laboratories [246]. The two-step preparation was
8.4
Zinc Oxide Nano-antimicrobials
developed earlier [245] and consisted in the generation of Cu NPs in acetone using
picosecond-pulsed laser ablation followed by their mixing with a PLA solution in
an ultrasonic bath to get the composites. The one-pot synthesis allowed the laser
ablation of the copper target in acetone in the presence of PLA as stabilizer [246].
In both cases, the composites could be deposited as thin films by drop casting. The
as-prepared nanocomposites were used in preservation tests for fiordilatte cheese
[247]. The bioactive films showed good antibacterial activity. In certain conditions,
it was found that the proliferation of main spoilage microorganisms was delayed
with a consequent preservation of sensory attributes.
8.3.1.5 Biological Syntheses
All the presented protocols have pros and cons, which are often related also to
limited eco-compatibility of the synthesis; therefore, attention toward “green”
methods using natural substances has grown exponentially in recent times.
Obviously, the expanding market of antimicrobials based on nanometals/oxides
was also involved. In fact, recent reviews addressed the importance of biological
synthesis [171, 175, 248] to develop copper-based nanomaterials – either simple
natural compounds (e.g., glutathione) [249] or natural extracts [250–253]. An
interesting approach by Barua et al. [250] was based on the chemical extraction of
cellulose nanofibrils from an abundant source (Colocasia esculenta) followed by
the reduction of Cu(II) acetate operated by the alcoholic extracts of Terminalia
chebula fruit. The stem latex of a medicinally important plant, Euphorbia nivulia,
was also successfully used to induce the synthesis of Cu (or Ag) NPs in high yield
by microwave assistance. Euphol (a major latex component) was responsible
for reduction whereas NP stabilization was ensured by certain peptides and
terpenoids present within the latex (as assessed by FT-IR analysis) [251].
8.4
Zinc Oxide Nano-antimicrobials
The application of metal oxides in the form of nanostructured materials is continuously expanding. Among them, in the last decades, ZnO-based nanomaterials
have attracted the interest of scientists and technologists as assessed by the huge
amount of reports on the synthesis of nanosized zinc oxide materials. In fact, ZnO
is a versatile wide band-gap semiconductor with tunable optoelectronic and piezoelectric properties. Furthermore, ZnO is chemically stable and almost harmless to
humans showing also good bioactivity toward several microorganisms.
8.4.1
Synthesis of Zinc Oxide Nanostructures
Various methods have been employed to synthesize ZnO NPs, including physical,
chemical, and electrochemical techniques. Conventionally, these methods are
grouped into two classes: top–down and bottom–up procedures [254]. The first
197
198
8 Nano-Antimicrobials Based on Metals
are generally physical processes [255–257], reducing bulk materials to the
nano-scale; large quantities of NPs can be produced, whereas the synthesis
of monodispersed and size-controlled colloids might be difficult to achieve.
Bottom–up methods, mainly chemical/electrochemical ones, on the other hand,
can be used to produce, in laboratory scale, uniform and size-controlled NPs
[258, 259].
8.4.1.1 Physical Approaches
Among physical approaches for the production of transition metal oxide NSs,
CVD is one of the most used, as it can be considered time and cost effective
[260]. This process may be defined as the deposition of a solid on a heated surface, from a chemical reaction in the vapor phase. It is a very versatile approach,
because by easily controlling the experimental parameters, the morphological and
chemical properties of the final product can be tuned [261]. Nevertheless, it is
scarcely used for the production of antimicrobial coatings and surfaces, because
of the need for high operating temperature that could degrade polymeric matrices and/or substrates [262]. Deposition techniques, such as radio-frequency (RF)
magnetron sputtering [263, 264], IBS [265, 266], and so on, are mostly used for the
production of ZnO-based antibacterial coatings. They are investigated to modify
biomedical implants, in order to reduce bacterial adhesion and viability and overthrow implant-associated infections [267]. Deposition techniques have also been
used for the functionalization of textiles and cotton fabrics, since they can easily provide uniform coverage of the surface [268]. Prospects of use of these thin
films in sterile packaging [269, 270], wound dressing [271], and non-fouling textiles [272] have been envisaged.
8.4.1.2 Chemical Approaches
The sol–gel chemical method has been widely used for producing metal oxide
powders with high purity and homogeneity [259, 273]. The process involves the
preparation of a colloidal suspension (sol), which is subsequently converted into
a viscous gel using principles of precipitation, condensation, and hydrolyzation
[274]. Surfactants can be useful for an accurate particle size tuning [275, 276].
ZnO colloids prepared with this approach have been used for the impregnation
of manufactured goods, in order to confer them with antifungal/antimicrobial
properties [277–279]. Recently, a report about in situ synthesized ZnO NSs,
on the surface of cotton fabrics, was provided [280]. This modified material
exerted high UV protection and antimicrobial activity against Gram-positive
and Gram-negative bacteria. Hybrid organic–inorganic composites, prepared
using sol–gel-produced ZnO colloids dispersed into polymeric matrices, have
been employed to preserve food from degradation [281]. Most of these materials
are characterized by multifunctionality, ease of processability, and the potential
for large-scale manufacturing [282]. Biocompatible and biodegradable polymers
have been used for this purpose, such as starch [269, 283–285], methylcellulose
(MA) [282], PLA [285], polyvinyl chloride (PVC) [286], and poly-(propylene
carbonate) (PPC) [287] (Figure 8.7).
8.4
Zinc Oxide Nano-antimicrobials
A (c)
A (b)
A (a)
B (c)
B (b)
Figure 8.7 Growth of A. flavus (A) and P.
citrinum (B) after the treatment by the sample films with ZnO NPs; A (a, b) and B (a, b),
ZnO NPs + PVC composite films; A (c) and
B (a)
B (c), blank PVC film as the control. (Reproduced from Ref. [286], Figure 3, Copyright
(2009), with permission from Wiley-VCH
Verlag GmbH & Co. KGaA, Weinheim.)
199
200
8 Nano-Antimicrobials Based on Metals
Another typical chemical route is the hydrothermal process, which involves
thermal decomposition of a precursor in the liquid phase [288]. Conveniently,
particle morphology and size, capable of affecting the final antimicrobial activity
[289, 290], can be modified by using different zinc precursors (nitrate, acetate,
chloride, etc.) [288]. NSs produced this way are usually coated on different
substrates: metallic sheets for biomedical implants [291], air filters [292], and
wound dressing materials [293].
8.4.1.3 Electrochemical Approaches
Interesting electrochemical approaches to unsupported dispersible ZnO NSs
are herein reviewed and classified into three main categories, according to
the composition of the electrolytic bath: electrochemical deposition under
oxidizing conditions (EDOC)-like, aqueous-based electrolyses, and alcoholbased electrolyses. EDOC-like processes, developed from the principle firstly
illustrated by Reetz et al. [294] for the synthesis of CoO colloids, was firstly
reported by Natter and coworkers [295, 296] for the preparation of ZnO NSs.
Anyway, no antimicrobial applications are known for this synthetic approach.
To the best of our knowledge, the very first example of aqueous electrochemical synthesis of ZnO colloids was reported in 2004 [297]; from there,
the fabrication of ZnO NPs by the potentiostatic anodization (20–40 V) of a
Zn sheet in 0.1 M NaOH solution at 15 ∘ C has been demonstrated [298]. The
use of slightly high potentials to synthesize unsupported ZnO NPs has been
overcome in the works by Venkatesha and coworkers [299–301]. They presented
a hybrid electrochemical–thermal method using a NaHCO3 aqueous solution
as alkaline medium under galvanostatic control at 10 mA cm−2 to generate a
mixed species of zinc carbonate and hydroxide, which can be decomposed
to ZnO by calcination at temperatures above 300 ∘ C [299, 300]. According
to the electrochemical–thermal method, we have recently developed a novel
approach based on the use of anionic stabilizers to better control the final
particle size and morphology of ZnO NPs galvanostatically grown from aqueous
media [302]. Few examples have been reported in the literature on the use
of electrolytic baths containing alcohols as the main solvent for the synthesis
of ZnO nano-colloids [303, 304]; high potentials and mixed ZnO/Zn colloids
are obtained this way, unless a small percentage of water is introduced in the
electrochemical media [305]. Finally, electrochemical approaches to ZnO nanomaterials are still mainly focused on the deposition of supported phases and/or
well-ordered nano-arrays. We found few reports about antimicrobial activity
of electrochemically deposited ZnO NSs [306]: Gupta and coworkers deposited
ZnO NRs onto F-SnO2 substrate by electrochemical deposition method from
aqueous solution, with a 15 mA/cm2 current density. The antibacterial activity
of ZnO NRs was studied for E. coli, showing a high inhibition of bacterial
growth.
References
8.5
Conclusions
As described in previous sections, several approaches have been employed to
prepare metal or metal oxide nano-antimicrobials, all of them presenting both
advantages and drawbacks. It should be considered that the final application of
the bioactive nanomaterials drives the choice of the synthesis. Moreover, the
proper design of effective nano-antimicrobials cannot exclude the knowledge
of bioactivity mechanisms, generally associated to NSs. Some extensive works
have already been presented on this topic, although some points have not been
completely clarified [3, 4, 16, 159, 307]. Generation of ROS, ion release, DNA
damage, oxidative stress as well as direct cell membrane damage [7, 172] were
accounted among the possible mechanisms. However, a complete treatise on
bioactivity of these agents is beyond the scope of the chapter.
For perspective uses, besides nano-antimicrobials based on Ag, Cu, and ZnO,
other appealing inorganic nanomaterials (such as TiO2 , Ga, Se, CNTs, etc.) are
emerging as antimicrobial agents among the scientific community [308]. TiO2
antimicrobials possess the peculiar property of being photoactivated [309, 310],
and are widely used in biomedical applications [311]. Ga(III), structurally similar
to Fe(III), cannot be reduced under physiological conditions, having therefore the
potential to serve as a Fe analog, capable of exerting high antibacterial activity, by
means of cellular uptake [312]. Se NPs, mostly used for antibacterial biomedical
coatings [313, 314], possess enormous antibacterial activity because of the massive
ROS generation and growth inhibition [315]. The unique and tunable properties
of carbon-based nanomaterials enable new technologies for the development of
novel antimicrobial strategies: fast membrane perturbation or disruption occurs,
leading bacteria in contact with CNTs [316], paving the way for the application of
these NSs as building blocks for novel antimicrobial materials [317].
As a last remark, along the development of metal-based nano-antimicrobials,
the study of their toxicity mechanisms toward important proteins, cells, tissues,
and organisms is receiving increasing attention as nanometer-size objects can
cause serious damages. As a result, the next generation of nano-antimicrobials
should be developed according to a multidisciplinary approach involving biology,
engineering, chemistry, and physics to ensure their large-scale production for safe
use and real life applications.
References
1. Silver, S., Phung, L.T., and Silver, G.
(2006) Silver as biocides in burn and
wound dressings and bacterial resistance to silver compounds. J. Ind.
Microbiol. Biotechnol., 33, 627–634.
2. Borkow, G. and Gabbay, J. (2009)
Copper, an ancient remedy returning to fight microbial, fungal and
viral infections. Curr. Chem. Biol., 3,
272–278.
3. Jones, N., Ray, B., Ranjit, K.T., and
Manna, A.C. (2008) Antibacterial activity of ZnO nanoparticle suspensions
on a broad spectrum of microorganisms. FEMS Microbiol. Lett., 279,
71–76.
201
202
8 Nano-Antimicrobials Based on Metals
4. Rai, M., Yadav, A., and Gade, A. (2009)
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
Silver nanoparticles as a new generation of antimicrobials. Biotechnol. Adv.,
27, 76–83.
Klasen, H.J. (2000) Historical review
of the use of silver in the treatment
of burns. I. Early uses. Burns, 26,
117–130.
Chopra, I. (2007) The increasing use of
silver-based products as antimicrobial
agents: a useful development or a cause
for concern? J. Antimicrob. Chemother.,
59, 587–590.
Grass, G., Rensing, C., and Solioz, M.
(2011) Metallic copper as an antimicrobial surface. Appl. Environ. Microbiol.,
77, 1541–1547.
Roselli, M., Finamore, A., Garaguso, I.,
Britti, M.S., and Mengheri, E. (2003)
Zinc oxide protects cultured enterocytes from the damage induced
by Escherichia coli. J. Nutr., 133,
4077–4082.
Kolář, M., Urbánek, K., and Látal, T.
(2001) Antibiotic selective pressure and
development of bacterial resistance. Int.
J. Antimicrob. Agents, 17, 357–363.
WHO (2013) Antimicrobial Resistance Fact Sheet N. 194, WHO
http://www.who.int/mediacentre/
factsheets/fs194/en/.
Llorens, A., Lloret, E., Picouet, P.A.,
Trbojevich, R., and Fernandez, A.
(2012) Metallic-based micro and
nanocomposites in food contact materials and active food packaging. Trends
Food Sci. Technol., 24, 19–29.
Dastjerdi, R. and Montazer, M. (2010)
A review on the application of inorganic nano-structured materials in the
modification of textiles: focus on antimicrobial properties. Colloids Surf. B
Biointerfaces, 79, 5–18.
Kaiser, J.-P., Zuin, S., and Wick, P.
(2013) Is nanotechnology revolutionizing the paint and lacquer industry?
A critical opinion. Sci. Total Environ.,
442, 282–289.
Cioffi, N. and Rai, M. (2012) NanoAntimicrobials: Progress and Prospects
Springer, Springer-Verlag, Berlin, Heidelberg.
15. Taylor, E. and Webster, T.J. (2011)
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
Reducing infections through nanotechnology and nanoparticles. Int. J.
Nanomed., 6, 1463.
Ingle, A., Duran, N., and Rai, M.
(2013) Bioactivity, mechanism of
action, and cytotoxicity of copperbased nanoparticles: a review. Appl.
Microbiol. Biotechnol., 98, 1–9.
Meng, X.K., Tang, S.C., and Vongehr,
S. (2010) A review on diverse silver
nanostructures. J. Mater. Sci. Technol.,
26, 487–522.
Wiley, B., Sun, Y., Mayers, B., and Xia,
Y. (2005) Shape-controlled synthesis of
metal nanostructures: the case of silver.
Chem. Eur. J., 11, 454–463.
Sau, K.T. and Rogach, A.L. (2012)
Complex-Shaped Metal Nanoparticles:
Bottom-Up Syntheses and Applications,
Wiley-VCH Verlag GmbH.
Kan, C.-X., Zhu, J.-J., and Zhu, X.G. (2008) Silver nanostructures with
well-controlled shapes: synthesis, characterization and growth mechanisms.
J. Phys. D: Appl. Phys., 41, 155304.
Pimpang, P., Sutham, W.,
Mangkorntong, N., Mangkorntong,
P., and Choopun, S. (2008) Effect of
stabilizer on preparation of silver
and gold nanoparticle using grinding
method. Chiang Mai J. Sci., 40 (2008),
250–257.
Tsuzuki, T. (2009) Commercial scale
production of inorganic nanoparticles.
Int. J. Nanotechnol., 6, 567–578.
Förster, H., Wolfrum, C., and Peukert,
W. (2012) Experimental study of
metal nanoparticle synthesis by an
arc evaporation/condensation process.
J. Nanopart. Res., 14, 926 (16 pp).
Simchi, A., Ahmadi, R., Reihani, S.M.S.,
and Mahdavi, A. (2007) Kinetics and
mechanisms of nanoparticle formation
and growth in vapor phase condensation process. Mater. Des., 28, 850–856.
Harra, J., Mäkitalo, J., Siikanen, R.,
Virkki, M., Genty, G., Kobayashi,
T., Kauranen, M., and Mäkelä, J.M.
(2012) Size-controlled aerosol synthesis of silver nanoparticles for
plasmonic materials. J. Nanopart.
Res., 14, 870–880.
References
26. Goharshadi, E.K. and Azizi-
27.
28.
29.
30.
31.
32.
33.
34.
Toupkanloo, H. (2013) Silver colloid
nanoparticles: ultrasound-assisted
synthesis, electrical and rheological properties. Powder Technol., 237,
97–101.
Vasileva, P., Stefanova, L., Vaklev, N.,
and Dushkin, C. (2008) Ultrasoundassisted synthesis of silver nanoparticles
via sodium borohydride reduction.
Nanosci. Nanotechnol., 8, 76.
He, C., Liu, L., Fang, Z., Li, J., Guo,
J., and Wei, J. (2014) Formation and
characterization of silver nanoparticles in aqueous solution via ultrasonic
irradiation. Ultrason. Sonochem., 21,
542–548.
Chen, J., Wang, J., Zhang, X., and Jin,
Y. (2008) Microwave-assisted green
synthesis of silver nanoparticles by
carboxymethyl cellulose sodium and
silver nitrate. Mater. Chem. Phys., 108,
421–424.
Hu, B., Wang, S.-B., Wang, K., Zhang,
M., and Yu, S.-H. (2008) Microwaveassisted rapid facile “green” synthesis
of uniform silver nanoparticles: selfassembly into multilayered films and
their optical properties. J. Phys. Chem.
C, 112, 11169–11174.
Kudle, K.R., Donda, M.R., Merugu, R.,
Kudle, M.R., and Pratap, M.P.R. (2013)
Microwave assisted green synthesis of
silver nanoparticles using Boswellia
Serrata flower extract and evaluation of
their antimicrobial activity. Int. Res. J.
Pharm., 4, 197–200.
Manikprabhu, D. and Lingappa, K.
(2013) Microwave assisted rapid and
green synthesis of silver nanoparticles
using a pigment produced by streptomyces coelicolor klmp33. Bioinorg.
Chem. Appl., 2013, 1–5.
Jia, H., Xu, W., An, J., Li, D., and Zhao,
B. (2006) A simple method to synthesize triangular silver nanoparticles by
light irradiation. Spectrochim. Acta,
Part A Mol. Biomol. Spectrosc., 64,
956–960.
Zhuo, Y., Du, C., Li, X., Sun, W., and
Chu, Y. (2013) One-step synthesis and
photoluminescence properties of polycarbazole spheres and Ag/polycarbazole
35.
36.
37.
38.
39.
40.
41.
42.
43.
core/shell composites. Eur. Polym. J.,
49, 1365–1372.
Weimer, M.W., Chen, H., Giannelis,
E.P., and Sogah, D.Y. (1999) Direct synthesis of dispersed nanocomposites by
in situ living free radical polymerization using a silicate-anchored initiator.
J. Am. Chem. Soc., 121, 1615–1616.
Gao, F., Lu, Q., and Komarneni, S.
(2005) Interface reaction for the selfassembly of silver nanocrystals under
microwave-assisted solvothermal conditions. Chem. Mater., 17, 856–860.
Perkas, N., Shuster, M., Amirian, G.,
Koltypin, Y., and Gedanken, A. (2008)
Sonochemical immobilization of silver
nanoparticles on porous polypropylene.
J. Polym. Sci., Part A: Polym. Chem., 46,
1719–1729.
Perkas, N., Amirian, G., Dubinsky, S.,
Gazit, S., and Gedanken, A. (2007)
Ultrasound-assisted coating of nylon
6,6 with silver nanoparticles and its
antibacterial activity. J. Appl. Polym.
Sci., 104, 1423–1430.
Amendola, V. and Meneghetti, M.
(2009) Laser ablation synthesis in solution and size manipulation of noble
metal nanoparticles. Phys. Chem. Chem.
Phys., 11, 3805.
Singh, S.C. (2012) Nanomaterials,
Wiley-VCH Verlag GmbH & Co.
KGaA, Weinheim, pp. 317–437.
Ganeev, R.A., Baba, M., Ryasnyansky,
A.I., Suzuki, M., and Kuroda, H. (2004)
Characterization of optical and nonlinear optical properties of silver
nanoparticles prepared by laser ablation
in various liquids. Opt. Commun., 240,
437–448.
Dell’Aglio, M., Gaudiuso, R., ElRashedy,
R., Pascale, O.D., Palazzo, G., and
Giacomo, A.D. (2013) Collinear double
pulse laser ablation in water for the
production of silver nanoparticles. Phys.
Chem. Chem. Phys., 15, 20868–20875.
Amendola, V., Polizzi, S., and
Meneghetti, M. (2007) Free silver
nanoparticles synthesized by laser
ablation in organic solvents and their
easy functionalization. Langmuir, 23,
6766–6770.
203
204
8 Nano-Antimicrobials Based on Metals
44. Tilaki, R.M., Iraji zad, A., and Mahdavi,
45.
46.
47.
48.
49.
50.
51.
52.
53.
S.M. (2006) Stability, size and optical properties of silver nanoparticles
prepared by laser ablation in different carrier media. Appl. Phys. A, 84,
215–219.
Sreeja, R., Reshmi, R., Aneesh, P.M.,
and Jayaraj, M.K. (2012) Liquid phase
pulsed laser ablation of metal nanoparticles for nonlinear optical applications.
Sci. Adv. Mater., 4, 439–448.
Tsuji, T., Thang, D.-H., Okazaki,
Y., Nakanishi, M., Tsuboi, Y., and
Tsuji, M. (2008) Preparation of silver nanoparticles by laser ablation in
polyvinylpyrrolidone solutions. Appl.
Surf. Sci., 254, 5224–5230.
Tsuji, T., Mizuki, T., Ozono, S.,
and Tsuji, M. (2009) Laser-induced
silver nanocrystal formation in
polyvinylpyrrolidone solutions. J. Photochem. Photobiol., A Chem., 206,
134–139.
Zamiri, R., Zakaria, A., Sadrolhosseini,
A.R., Ahangar, H.A., and Drummen,
G.P.C. (2013) Laser-Ablation synthesis
and evaluation of thermal non-linear
optical properties of silver nanoparticles in monoolein. Sci. Adv. Mater., 5,
748–757.
Zamiri, R., Azmi, B.Z., Naseri, M.G.,
Ahangar, H.A., Darroudi, M., and
Nazarpour, F.K. (2011) Laser based
fabrication of chitosan mediated silver nanoparticles. Appl. Phys. A, 105,
255–259.
Zamiri, R., Azmi, B.Z., Ahangar, H.A.,
Zamiri, G., Husin, M.S., and Wahab,
Z.A. (2012) Preparation and characterization of silver nanoparticles in natural
polymers using laser ablation. Bull.
Mater. Sci., 35, 727–731.
Basu, S. and Pal, T. (2009) in Advanced
Wet-Chemical Synthetic Approaches
to Inorganic Nanostructures (ed D.P.
Cozzoli), Research Signpost, p. 219.
Gibbons, V.L. and Getman, F.H. (1914)
The potential of silver in non aqueous
solutions of silver nitrat. J. Am. Chem.
Soc., 36, 2091.
Caswell, K.K., Bender, C.M., and
Murphy, C.J. (2003) Seedless, surfactantless wet chemical synthesis of silver
nanowires. Nano Lett., 3, 667–669.
54. Nagabhushana, K.S. and Bönnemann,
55.
56.
57.
58.
59.
60.
61.
62.
63.
64.
H. (2004) in Nanotechnology in Catalysis (eds B. Zhou, S. Hermans, and G.A.
Somorjai), Boston, MA, Springer, pp.
51–82.
Pulit, J., Banach, M., and Kowalski,
Z. (2013) Chemical reduction as the
main method for obtaining nanosilver. J. Comput. Theor. Nanosci., 10,
276–284.
Wang, W. and Gu, B. (2004) Concentrated Dispersions, ACS Symposium
Series. American Chemical Society, pp.
1–14, 10.1021/bk-2004-0878.ch001.
Li, D., Hong, B., Fang, W., Guo, Y.,
and Lin, R. (2010) Preparation of
well-dispersed silver nanoparticles for
oil-based nanofluids. Ind. Eng. Chem.
Res., 49, 1697–1702.
García-Barrasa, J., López-de-Luzuriaga,
J.M., and Monge, M. (2010) Silver
nanoparticles: synthesis through chemical methods in solution and biomedical
applications. Cent. Eur. J. Chem., 9,
7–19.
Punitha, N. and Ramesh, P.S. (2013)
Characterization of polymer assisted
nano silver by chemical reduction
method. Asian J. Chem., 25, S157.
Tiwari, A.D., Mishra, A.K., Mishra,
S.B., Kuvarega, A.T., and Mamba, B.B.
(2013) Stabilisation of silver and copper
nanoparticles in a chemically modified
chitosan matrix. Carbohydr. Polym., 92,
1402–1407.
Yuan, W., Fu, J., Su, K., and Ji, J. (2010)
Self-assembled chitosan/heparin multilayer film as a novel template for
in situ synthesis of silver nanoparticles. Colloids Surf. B Biointerfaces, 76,
549–555.
Xin, J., Yin, X., Chen, S., and Wu, A.
(2012) Synthesis of uniform and stable
silver nanoparticles by a gold seedmediated growth approach in a buffer
system. J. Exp. Nanosci., 9, 1–9.
Sathiya Priya, R., Geetha, D., and
Ramesh, P.S. (2013) Antibacterial
activity of nano-silver capped by βcyclodextrin. Carbon Sci. Technol., 5,
197.
Tavallai, H. and Pouresmaeil, S. (2012)
Determination of ascorbic acid by
References
65.
66.
67.
68.
69.
70.
71.
72.
73.
modified method based on photoluminescence of silver nanoparticles. Int. J.
Chem. Tech. Res., 4, 304.
Kwon, J.W., Yoon, S.H., Lee, S.S., Seo,
K.W., and Shim, I.W. (2005) Preparation of silver nanoparticles in cellulose
acetate polymer and the reaction chemistry of silver complexes in the polymer.
Bull. Korean Chem. Soc., 26, 837–840.
Oluwafemi, O.S., Lucwaba, Y., Gura, A.,
Masabeya, M., Ncapayi, V., Olujimi,
O.O., and Songca, S.P. (2013) A
facile completely “green” size tunable synthesis of maltose-reduced
silver nanoparticles without the use
of any accelerator. Colloids Surf. B
Biointerfaces, 102, 718–723.
Kuo, P.-L., Chen, C.-C., and Yuen,
S.-M. (2004) Protection effects of
hydrophile-grafted silicone copolymers
on the formation of colloidal silver
nanoparticles. J. Phys. Chem. B, 108,
5541–5546.
Wang, H., Qiao, X., Chen, J., Wang, X.,
and Ding, S. (2005) Mechanisms
of PVP in the preparation of silver
nanoparticles. Mater. Chem. Phys., 94,
449–453.
Wang, H., Qiao, X., Chen, J., and
Ding, S. (2005) Preparation of silver
nanoparticles by chemical reduction
method. Colloids Surf., A Physicochem.
Eng. Aspects, 256, 111–115.
Solanki, J.N. and Murthy, Z.V.P. (2010)
Highly monodisperse and sub-nano
silver particles synthesis via microemulsion technique. Colloids Surf., A
Physicochem. Eng. Aspects, 359, 31–38.
Solanki, J.N. and Murthy, Z.V.P. (2011)
Reduction of nitro aromatic compounds over Ag/Al2 O3 nanocatalyst
prepared in water-in-oil microemulsion:
effects of water-to-surfactant mole ratio
and type of reducing agent. Ind. Eng.
Chem. Res., 50, 7338–7344.
Zhang, W., Qiao, X., and Chen,
J. (2007) Synthesis of silver
nanoparticles—Effects of concerned
parameters in water/oil microemulsion.
Mater. Sci. Eng., B, 142, 1–15.
Verma, V. (2009) Biosynthesis of noble
metal nanoparticles and their application. CAB Rev. Perspect. Agric. Vet. Sci.
Nutr. Nat. Resour., 4 (026), 1–17.
74. Iravani, S. (2011) Green synthesis of
75.
76.
77.
78.
79.
80.
81.
82.
83.
metal nanoparticles using plants. Green
Chem., 13, 2638.
Pavani, K.V., Kumar, N.S., and
Gayathramma, K. (2012) Plants as
ecofriendly nanofactories. J. Bionanosci.,
6, 1–6.
Naik, R.R., Stringer, S.J., Agarwal, G.,
Jones, S.E., and Stone, M.O. (2002)
Biomimetic synthesis and patterning
of silver nanoparticles. Nat. Mater., 1,
169–172.
Ahmad, A., Mukherjee, P., Senapati, S.,
Mandal, D., Khan, M.I., Kumar, R., and
Sastry, M. (2003) Extracellular biosynthesis of silver nanoparticles using the
fungus Fusarium oxysporum. Colloids
Surf. B Biointerfaces, 28, 313–318.
Bhainsa, K.C. and D’Souza, S.F. (2006)
Extracellular biosynthesis of silver nanoparticles using the fungus
Aspergillus fumigatus. Colloids Surf. B
Biointerfaces, 47, 160–164.
Wei, X., Zhou, H., Xu, L., Luo, M., and
Liu, H. (2014) Sunlight-induced biosynthesis of silver nanoparticles by animal
and fungus biomass and their characterization. J. Chem. Technol. Biotechnol.,
89, 305–311.
Mukherjee, P., Ahmad, A., Mandal,
D., Senapati, S., Sainkar, S.R., Khan,
M.I., Parishcha, R., Ajaykumar, P.V.,
Alam, M., Kumar, R. et al. (2001)
Fungus-mediated synthesis of silver
nanoparticles and their immobilization
in the mycelial matrix: a novel biological approach to nanoparticle synthesis.
Nano Lett., 1, 515–519.
Gurunathan, S., Kalishwaralal, K.,
Vaidyanathan, R., Venkataraman,
D., Pandian, S.R.K., Muniyandi, J.,
Hariharan, N., and Eom, S.H. (2009)
Biosynthesis, purification and characterization of silver nanoparticles
using Escherichia coli. Colloids Surf. B
Biointerfaces, 74, 328–335.
Kalimuthu, K., Suresh Babu, R.,
Venkataraman, D., Bilal, M., and
Gurunathan, S. (2008) Biosynthesis of
silver nanocrystals by Bacillus licheniformis. Colloids Surf. B Biointerfaces,
65, 150–153.
Malarkodi, C., Rajeshkumar, S.,
Paulkumar, K., Vanaja, M., Jobitha,
205
206
8 Nano-Antimicrobials Based on Metals
84.
85.
86.
87.
88.
89.
90.
91.
G.D.G., and Annadurai, G. (2013) Bacselected human pathogens. J. Exp.
Nanosci., 9, 197–209.
tericidal activity of bio mediated silver
92. Roy, K., Biswas, B., and Banerjee,
nanoparticles synthesized by Serratia
P.C. (2013) “Green” synthesis of silver
nematodiphila. Drug Invent. Today, 5,
nanoparticles by using grape (Vitis
119–125.
vinifera) fruit extract: characterization
Malhotra, A., Dolma, K., Kaur, N.,
of the particles and study of antibacteRathore, Y.S., Ashish, Mayilraj, S.,
rial activity. Res. J. Pharm. Biol. Chem.
and Choudhury, A.R. (2013) BiosynSci., 4, 1271.
thesis of gold and silver nanoparti93. Vidhu, V.K. and Philip, D. (2014) Speccles using a novel marine strain of
troscopic, microscopic and catalytic
Stenotrophomonas. Bioresour. Technol.,
properties of silver nanoparticles syn142, 727–731.
thesized using Saraca indica flower.
Vidhu, V.K. and Philip, D. (2014) CatSpectrochim. Acta, Part A Mol. Biomol.
alytic degradation of organic dyes using
Spectrosc., 117, 102–108.
biosynthesized silver nanoparticles.
94.
Khaydarov, R.A., Khaydarov, R.R.,
Micron, 56, 54–62.
Gapurova, O., Estrin, Y., and Scheper,
Annamalai, A., Christina, V.L.P.,
T. (2009) Electrochemical method for
Christina, V., and Lakshmi, P.T.V. (2014)
the synthesis of silver nanoparticles.
Green synthesis and characterisation
J. Nanopart. Res., 11, 1193–1200.
of Ag NPs using aqueous extract of
95. Reetz, M.T. and Helbig, W. (1994) SizePhyllanthus maderaspatensis L. J. Exp.
selective synthesis of nanostructured
Nanosci., 9, 113–119.
transition metal clusters. J. Am. Chem.
Ghaffari-Moghaddam, M. and
Soc., 116, 7401–7402.
Hadi-Dabanlou, R. (2013) Plant medi96. Rodríguez-Sánchez, L., Blanco, M.C.,
ated green synthesis and antibacterial
and López-Quintela, M.A. (2000)
activity of silver nanoparticles using
Electrochemical synthesis of silver
Crataegus douglasii fruit extract. J. Ind.
nanoparticles. J. Phys. Chem. B, 104,
Eng. Chem., 20, 739–744.
9683–9688.
Nazeruddin, G.M., Prasad, N.R.,
97. Starowicz, M., Stypuła, B., and Banaś,
Waghmare, S.R., Garadkar, K.M.,
J. (2006) Electrochemical synthesis
and Mulla, I.S. (2014) Extracellular
of silver nanoparticles. Electrochem.
biosynthesis of silver nanoparticle using
Commun., 8, 227–230.
Azadirachta indica leaf extract and its
98. Jian, Z., Xiang, Z., and Yongchang, W.
anti-microbial activity. J. Alloys Compd.,
(2005) Electrochemical synthesis and
583, 272–277.
fluorescence spectrum properties of
Shankar, S.S., Ahmad, A., and Sastry,
silver nanospheres. Microelectron. Eng.,
M. (2003) Geranium leaf assisted
77, 58–62.
biosynthesis of silver nanoparticles.
99. Rabinal, M.K., Kalasad, M.N.,
Biotechnol. Prog., 19, 1627–1631.
Praveenkumar, K., Bharadi, V.R.,
Shanmugam, N., Rajkamal, P.,
and Bhikshavartimath, A.M. (2013)
Cholan, S., Kannadasan, N.,
Electrochemical synthesis and optical
Sathishkumar, K., Viruthagiri, G., and
properties of organically capped silver
Sundaramanickam, A. (2014) Biosynnanoparticles. J. Alloys Compd., 562,
thesis of silver nanoparticles from the
43–47.
marine seaweed Sargassum wightii
100. Cioffi, N., Colaianni, L., Pilolli, R.,
and their antibacterial activity against
Calvano, C., Palmisano, F., and
some human pathogens. Appl. Nanosci.,
Zambonin, P. (2009) Silver nanofractals:
September 2013, 1–8.
electrochemical synthesis, XPS characDaniel, S.C.G.K., Banu, B.N., Harshiny,
terization and application in LDI-MS.
M., Nehru, K., Ganesh, P.S., Kumaran,
Anal. Bioanal. Chem., 394, 1375–1383.
S., and Sivakumar, M. (2014) Ipomea
101. Lee, C.-L. and Syu, C.-M. (2010)
carnea-based silver nanoparticle synElectrochemical synthesis of
thesis for antibacterial activity against
hexadecyltrimethylammonium-coated
References
102.
103.
104.
105.
106.
107.
108.
109.
110.
111.
112.
Ag nanopeanuts and their self-assembly
to nanonets. Colloids Surf., A Physicochem. Eng. Aspects, 358, 158–162.
ten Kortenaar, M.V., Kolar, Z.I., and
Tichelaar, F.D. (1999) Formation of
long-lived silver clusters in aqueous
solution by anodic dispersion. J. Phys.
Chem. B, 103, 2054–2060.
Yin, B., Ma, H., Wang, S., and Chen,
S. (2003) Electrochemical synthesis of
silver nanoparticles under protection
of poly( N -vinylpyrrolidone). J. Phys.
Chem. B, 107, 8898–8904.
Ma, H., Yin, B., Wang, S., Jiao, Y.,
Pan, W., Huang, S., Chen, S., and
Meng, F. (2004) Synthesis of silver and
gold nanoparticles by a novel electrochemical method. ChemPhysChem, 5,
68–75.
Johans, C. (2002) Electrosynthesis
of polyphenylpyrrole coated silver
particles at a liquid–liquid interface.
Electrochem. Commun., 4, 227–230.
Roldán, M.V., Pellegri, N., and de
Sanctis, O. (2013) Electrochemical
method for Ag-PEG nanoparticles
synthesis. J. Nanopart., 2013, 1–7.
Zhu, J.-J., Liao, X.-H., Zhao, X.-N., and
Chen, H.-Y. (2001) Preparation of silver
nanorods by electrochemical methods.
Mater. Lett., 49, 91–95.
Reicha, F.M., Sarhan, A., Abdel-Hamid,
M.I., and El-Sherbiny, I.M. (2012)
Preparation of silver nanoparticles in
the presence of chitosan by electrochemical method. Carbohydr. Polym.,
89, 236–244.
Zoval, J.V., Stiger, R.M., Biernacki, P.R.,
and Penner, R.M. (1996) Electrochemical deposition of silver nanocrystallites
on the atomically smooth graphite basal
plane. J. Phys. Chem., 100, 837–844.
Yuan, G., Chang, X., and Zhu, G.
(2011) Electrosynthesis and catalytic
properties of silver nano/microparticles
with different morphologies. Particuology, 9, 644–649.
Guan, D. and Wang, Y. (2012) Electrodeposition of Ag nanoparticles onto
bamboo-type TiO2 nanotube arrays to
improve their lithium-ion intercalation
performance. Ionics, 19, 879–885.
Fang, Y.-M., Lin, Z.-B., Zeng, Y.-M.,
Chen, W.-K., Chen, G.-N., Sun, J.-J.,
113.
114.
115.
116.
117.
118.
119.
120.
121.
122.
Ren, B., and Tian, Z.-Q. (2010) Facile
electrochemical preparation of Ag nanothorns and their growth mechanism.
Chem. Eur. J., 16, 6766–6770.
Mazur, M. (2004) Electrochemically prepared silver nanoflakes and
nanowires. Electrochem. Commun., 6,
400–403.
Tang, Z., Liu, S., Dong, S., and Wang,
E. (2001) Electrochemical synthesis of
Ag nanoparticles on functional carbon
surfaces. J. Electroanal. Chem., 502,
146–151.
Zhong, L., Gan, S., Fu, X., Li, F., Han,
D., Guo, L., and Niu, L. (2013) Electrochemically controlled growth of silver
nanocrystals on graphene thin film and
applications for efficient nonenzymatic
H2 O2 biosensor. Electrochim. Acta, 89,
222–228.
Tang, S., Meng, X., Wang, C., and
Cao, Z. (2009) Flowerlike Ag microparticles with novel nanostructure
synthesized by an electrochemical
approach. Mater. Chem. Phys., 114,
842–847.
Zhitomirsky, I. and Hashambhoy,
A. (2007) Chitosan-mediated electrosynthesis of organic–inorganic
nanocomposites. J. Mater. Process.
Technol., 191, 68–72.
Hache, F., Ricard, D., and Flytzanis, C.
(1986) Optical nonlinearities of small
metal particles: surface-mediated resonance and quantum size effects. J. Opt.
Soc. Am. B, 3, 1647–1655.
Lee, K.-C., Lin, S.-J., Lin, C.-H., Tsai,
C.-S., and Lu, Y.-J. (2008) Size effect of
Ag nanoparticles on surface plasmon
resonance. Surf. Coat. Technol., 202,
5339–5342.
Haiss, W., Thanh, N.T.K., Aveyard, J.,
and Fernig, D.G. (2007) Determination of size and concentration of gold
nanoparticles from UV − vis spectra.
Anal. Chem., 79, 4215–4221.
Evanoff, D.D. and Chumanov, G.
(2004) Size-controlled synthesis of
nanoparticles. 2. Measurement of
extinction, scattering, and absorption
cross sections. J. Phys. Chem. B, 108,
13957–13962.
Mock, J.J., Barbic, M., Smith, D.R.,
Schultz, D.A., and Schultz, S. (2002)
207
208
8 Nano-Antimicrobials Based on Metals
123.
124.
125.
126.
127.
128.
129.
130.
131.
Shape effects in plasmon resonance of
individual colloidal silver nanoparticles.
J. Chem. Phys., 116, 6755–6759.
Prema, P. (2011) Chemical mediated
synthesis of silver nanoparticles and its
potential antibacterial application, in
Progress in Molecular and Environmental Bioengineering – From Analysis and
Modeling to Technology Applications (ed
A. Carpi), InTech. doi: 10.5772/22114
Sarkar, A., Kapoor, S., and
Mukherjee, T. (2010) Oleic acidassisted phase transfer of nanosized
silver colloids. Res. Chem. Intermed.,
36, 403–410.
Barth, H.G. and Flippen, R.B. (1995)
Particle size analysis. Anal. Chem., 67,
257–272.
Williams, D.B. and Carter, C.B. (2009)
Transmission Electron Microscopy,
2nd edn, Springer Science & Business
Media.
Wang, Z.L. (2000) Transmission electron microscopy of shape-controlled
nanocrystals and their assemblies.
J. Phys. Chem. B, 104, 1153–1175.
Abdelgawad, A.M., Hudson, S.M.,
and Rojas, O.J. (2014) Antimicrobial
wound dressing nanofiber mats from
multicomponent (chitosan/silverNPs/polyvinyl alcohol) systems.
Carbohydr. Polym., 100, 166–178.
Ataeefard, M. and Sharifi, S. (2014)
Antibacterial flexographic ink containing silver nanoparticles. Prog. Org.
Coat., 77, 118–123.
Cioffi, N., Ditaranto, N., Torsi, L.,
Picca, R.A., Giglio, E.D., Sabbatini, L.,
Novello, L., Tantillo, G., Bleve-Zacheo,
T., and Zambonin, P.G. (2005) Synthesis, analytical characterization and
bioactivity of Ag and Cu nanoparticles
embedded in poly-vinyl-methyl-ketone
films. Anal. Bioanal. Chem., 382,
1912–1918.
Panáček, A., Kvítek, L., Prucek, R.,
Kolář, M., Večeřová, R., Pizúrová,
N., Sharma, V.K., Nevěčná, T., and
Zbořil, R. (2006) Silver colloid nanoparticles: synthesis, characterization, and
their antibacterial activity. J. Phys.
Chem. B, 110, 16248–16253.
132. Shahverdi, A.R., Fakhimi, A., Shahverdi,
133.
134.
135.
136.
137.
138.
139.
140.
141.
H.R., and Minaian, S. (2007) Synthesis and effect of silver nanoparticles
on the antibacterial activity of different antibiotics against Staphylococcus
aureus and Escherichia coli. Nanomed.
Nanotechnol. Biol. Med., 3, 168–171.
Joy, D.C. (2006) Scanning electron
microscopy, in Materials Science and
Technology, Wiley-VCH Verlag GmbH
& Co. KGaA.
Brook, L.A., Evans, P., Foster, H.A.,
Pemble, M.E., Steele, A., Sheel, D.W.,
and Yates, H.M. (2007) Highly bioactive
silver and silver/titania composite films
grown by chemical vapour deposition.
J. Photochem. Photobiol., A Chem., 187,
53–63.
Rill, M.S., Plet, C., Thiel, M., Staude,
I., von Freymann, G., Linden, S., and
Wegener, M. (2008) Photonic metamaterials by direct laser writing and
silver chemical vapour deposition. Nat.
Mater., 7, 543–546.
Tzeng Lue, J. (2007) in Encyclopedia
of Nanoscience and Nanotechnology
(ed H.S. Nalwa), American Scientific
Publishers, pp. 1–46.
Zhu, X., Birringer, R., Herr, U., and
Gleiter, H. (1987) X-ray diffraction
studies of the structure of nanometersized crystalline materials. Phys. Rev. B,
35, 9085–9090.
Khanna, P.K., Singh, N., Kulkarni,
D., Deshmukh, S., Charan, S., and
Adhyapak, P.V. (2007) Water based
simple synthesis of re-dispersible silver nano-particles. Mater. Lett., 61,
3366–3370.
Uvarov, V. and Popov, I. (2007) Metrological characterization of X-ray
diffraction methods for determination
of crystallite size in nano-scale materials. Mater. Charact., 58, 883–891.
Wagner, C.D. (1979) Handbook of
X-Ray Photoelectron Spectroscopy:
A Reference Book of Standard Data
for Use in X-Ray Photoelectron Spectroscopy, Physical Electronics Division,
Perkin-Elmer Corp.
Wertheim, G.K., DiCenzo, S.B., and
Buchanan, D.N.E. (1986) Noble- and
transition-metal clusters: the d bands of
References
142.
143.
144.
145.
146.
147.
148.
149.
150.
silver and palladium. Phys. Rev. B, 33,
5384–5390.
Shin, H.S., Choi, H.C., Jung, Y.,
Kim, S.B., Song, H.J., and Shin, H.J.
(2004) Chemical and size effects of
nanocomposites of silver and polyvinyl
pyrrolidone determined by X-ray photoemission spectroscopy. Chem. Phys.
Lett., 383, 418–422.
Schumm, B., Wisser, F.M., Mondin,
G., Hippauf, F., Fritsch, J., Grothe, J.,
and Kaskel, S. (2012) Semi-transparent
silver electrodes for flexible electronic
devices prepared by nanoimprint
lithography. J. Mater. Chem. C, 1,
638–645.
Zhou, W. (2011) A label-free biosensor
based on silver nanoparticles array for
clinical detection of serum p53 in head
and neck squamous cell carcinoma. Int.
J. Nanomed., 6, 381–386.
Chung, C.K., Fung, P.K., Hong, Y.Z., Ju,
M.S., Lin, C.C.K., and Wu, T.C. (2006)
A novel fabrication of ionic polymermetal composites (IPMC) actuator with
silver nano-powders. Sens. Actuators, B
Chem., 117, 367–375.
Chun, K.-Y., Oh, Y., Rho, J.,
Ahn, J.-H., Kim, Y.-J., Choi, H.R.,
and Baik, S. (2010) Highly conductive,
printable and stretchable composite
films of carbon nanotubes and silver.
Nat. Nanotechnol., 5, 853–857.
Li, Y., Wu, Y., and Ong, B.S. (2005)
Facile synthesis of silver nanoparticles useful for fabrication of highconductivity elements for printed
electronics. J. Am. Chem. Soc., 127,
3266–3267.
Wu, Y., Li, Y., and Ong, B.S. (2007)
A simple and efficient approach to a
printable silver conductor for printed
electronics. J. Am. Chem. Soc., 129,
1862–1863.
Wijesundera, D.N., Rajapaksa, I., Wang,
X., Liu, J.-R., Rusakova, I., and Chu,
W.-K. (2013) Ion beam engineered
nano silver silicon substrates for surface enhanced Raman spectroscopy.
J. Raman Spectrosc., 44, 1014–1017.
Pinkhasova, P., Puccio, B., Chou, T.,
Sukhishvili, S., and Du, H. (2012) Noble
metal nanostructure both as a SERS
151.
152.
153.
154.
155.
156.
157.
158.
159.
nanotag and an analyte probe. Chem.
Commun., 48, 9750–9752.
Shanmukh, S., Jones, L., Driskell, J.,
Zhao, Y., Dluhy, R., and Tripp, R.A.
(2006) Rapid and sensitive detection
of respiratory virus molecular signatures using a silver nanorod array SERS
substrate. Nano Lett., 6, 2630–2636.
Abid, M.I., Wang, L., Zhang, X., and
Xu, Y. (2013) Silver nano islands
enhanced Raman scattering on large
area grating substrates fabricated by
two beam laser interference. Chem. Res.
Chin. Univ., 29, 1006–1010.
Fu, Q., Zhang, D., Chen, Y., Wang, X.,
Han, L., Zhu, L., Wang, P., and Ming,
H. (2013) Surface enhanced Raman
scattering arising from plasmonic interaction between silver nano-cubes and
a silver grating. Appl. Phys. Lett., 103,
041122.
Hossain, W., Ghosh, M., Sinha, C.,
Debnath, D.K., and Sarkar, U.K. (2013)
SERS and DFT study of silver nano
particle induced dark isomerisation
in 1H-2(Phenylazo) imidazole. Chem.
Phys. Lett., 586, 132–137.
Orivnáková, R., Skantárová, L., Orivnák,
A., Demko, J., Kupková, M., and
Andersson, J.T. (2013) Electrochemical
deposition of SERS active nanostructured silver films. Int. J. Electrochem.
Sci., 8, 80–99.
Kim, J.S., Kuk, E., Yu, K.N., Kim, J.-H.,
Park, S.J., Lee, H.J., Kim, S.H., Park,
Y.K., Park, Y.H., Hwang, C.-Y. et al.
(2007) Antimicrobial effects of silver
nanoparticles. Nanomed. Nanotechnol.
Biol. Med., 3, 95–101.
Sondi, I. and Salopek-Sondi, B. (2004)
Silver nanoparticles as antimicrobial
agent: a case study on E. coli as a
model for Gram-negative bacteria.
J. Colloid Interface Sci., 275, 177–182.
Lu, Z., Rong, K., Li, J., Yang, H.,
and Chen, R. (2013) Size-dependent
antibacterial activities of silver nanoparticles against oral anaerobic pathogenic
bacteria. J. Mater. Sci. Mater. Med., 24,
1465–1471.
Marambio-Jones, C. and Hoek, E.M.V.
(2010) A review of the antibacterial
effects of silver nanomaterials and
potential implications for human health
209
210
8 Nano-Antimicrobials Based on Metals
160.
161.
162.
163.
164.
165.
166.
167.
168.
and the environment. J. Nanopart. Res.,
12, 1531–1551.
Kvítek, L., Panáček, A., Soukupová, J.,
Kolář, M., Večeřová, R., Prucek, R.,
Holecová, M., and Zbořil, R. (2008)
Effect of Surfactants and Polymers on
Stability and Antibacterial Activity of
Silver Nanoparticles (NPs). J. Phys.
Chem. C, 112, 5825–5834
Shrivastava, S., Bera, T., Roy, A., Singh,
G., Ramachandrarao, P., and Dash, D.
(2007) Characterization of enhanced
antibacterial effects of novel silver
nanoparticles. Nanotechnology, 18,
225103 (9pp).
Teeguarden, J., Hinderliter, P., Orr,
G., Thrall, B., and Pounds, J. (2007)
Particokinetics in vitro: dosimetry considerations for in vitro nanoparticle
toxicity assessments. Toxicol. Sci., 95,
300–312.
Morones, J., Elechiguerra, J.,
Camacho, A., Holt, K., Kouri, J.,
Ramirez, J., and Yacaman, M. (2005)
The bactericidal effect of silver
nanoparticles. Nanotechnology, 16,
2346–2353.
Stevanović, M., Bračko, I., Milenković,
M., Filipović, N., Nunić, J., Filipič,
M., and Uskoković, D.P. (2014) Multifunctional PLGA particles containing
poly(l-glutamic acid)-capped silver
nanoparticles and ascorbic acid with
simultaneous antioxidative and prolonged antimicrobial activity. Acta
Biomater., 10, 151–162.
Dubas, S.T., Kumlangdudsana, P., and
Potiyaraj, P. (2006) Layer-by-layer
deposition of antimicrobial silver
nanoparticles on textile fibers. Colloids
Surf., A Physicochem. Eng. Aspects, 289,
105–109.
Chappell, J.B. and Greville, G.D. (1954)
Effect of silver ions on mitochondrial
adenosine triphosphatase. Nature, 174,
930–931.
Jensen, R.H. and Davidson, N. (1966)
Spectrophotometric, potentiometric,
and density gradient ultracentrifugation
studies of the binding of silver ion by
DNA. Biopolymers, 4, 17–32.
Stohs, S. (1995) Oxidative mechanisms in the toxicity of metal ions. Free
Radical Biol. Med., 18, 321–336.
169. Halliwell, B. and Gutteridge, J.M. (1984)
170.
171.
172.
173.
174.
175.
176.
177.
178.
Oxygen toxicity, oxygen radicals, transition metals and disease. Biochem. J.,
219, 1–14.
Peña, M.M.O., Koch, K.A., and Thiele,
D.J. (1998) Dynamic regulation of copper uptake and detoxification genes in
saccharomyces cerevisiae. Mol. Cell.
Biol., 18, 2514–2523.
Varshney, R., Bhadauria, S., and
Gaur, M.S. (2012) A review: biological synthesis of silver and copper
nanoparticles. Nano Biomed. Eng., 4
(2), 99–106.
Ruparelia, J.P., Chatterjee, A.K.,
Duttagupta, S.P., and Mukherji, S.
(2008) Strain specificity in antimicrobial
activity of silver and copper nanoparticles. Acta Biomater., 4, 707–716.
Yoon, K.-Y., Hoon Byeon, J., Park,
J.-H., and Hwang, J. (2007) Susceptibility constants of Escherichia coli and
Bacillus subtilis to silver and copper
nanoparticles. Sci. Total Environ., 373,
572–575.
Longano, D., Ditaranto, N., Sabbatini,
L., Torsi, L., and Cioffi, N. (2012) in
Nano-Antimicrobials: Progress and
Prospects (eds N. Cioffi and M. Rai),
Springer-Verlag, Berlin, Heidelberg, pp.
85–117.
Rubilar, O., Rai, M., Tortella, G., Diez,
M., Seabra, A., and Durán, N. (2013)
Biogenic nanoparticles: copper, copper
oxides, copper sulphides, complex
copper nanostructures and their
applications. Biotechnol. Lett., 35,
1365–1375.
Cioffi, N., Ditaranto, N., Torsi, L., and
Sabbatini, L. (2009) in Metallic Nanomaterials Nanotechnologies for the Life
Sciences (ed. C.S.S.R. Kumar), WileyVCH Verlag GmbH & Co. KGaA,
Weinheim, pp. 3–70.
Cheruthazhekatt, S., Černák, M.,
Slavíček, P., and Havel, J. (2010) Gas
plasmas and plasma modified materials
in medicine. J. Appl. Biomed., 8, 55–66.
Zhang, W. and Chu, P.K. (2008)
Enhancement of antibacterial properties
and biocompatibility of polyethylene
by silver and copper plasma immersion
ion implantation. Surf. Coat. Technol.,
203, 909–912.
References
179. Zhang, W., Zhang, Y.-H., Ji, J.-H.,
180.
181.
182.
183.
184.
185.
186.
Zhao, J., Yan, Q., and Chu, P.K. (2006)
Antimicrobial properties of copper
plasma-modified polyethylene. Polymer,
47, 7441–7445.
Zhang, W., Zhang, Y., Ji, J., Yan, Q.,
Huang, A., and Chu, P.K. (2007)
Antimicrobial polyethylene with controlled copper release. J. Biomed. Mater.
Res. A, 83A, 838–844.
Codiã, I., Caplan, D.M., Drãgulescu,
E.-C., Lixandru, B.-E., Coldea, I.L.,
Dragomirescu, C.C., Surdu-Bob, C.,
and Bãdulescu, M. (2010) Antimicrobial
activity of copper and silver nanofilms
on nosocomial bacterial species. Rom.
Arch., 18, 204.
Ren, G., Hu, D., Cheng, E.W.C.,
Vargas-Reus, M.A., Reip, P., and
Allaker, R.P. (2009) Characterisation of
copper oxide nanoparticles for antimicrobial applications. Int. J. Antimicrob.
Agents, 33, 587–590.
Kelly, P.J., Li, H., Benson, P.S.,
Whitehead, K.A., Verran, J., Arnell,
R.D., and Iordanova, I. (2010) Comparison of the tribological and antimicrobial properties of CrN/Ag, ZrN/Ag,
TiN/Ag, and TiN/Cu nanocomposite
coatings. Surf. Coat. Technol., 205,
1606–1610.
Rio, L., Kusiak-Nejman, E., Kiwi, J.,
Bétrisey, B., Pulgarin, C., Trampuz,
A., and Bizzini, A. (2012) Comparison
of methods for evaluation of the bactericidal activity of copper-sputtered
surfaces against methicillin-resistant
staphylococcus aureus. Appl. Environ.
Microbiol., 78, 8176–8182.
Osorio-Vargas, P., Sanjines, R.,
Ruales, C., Castro, C., Pulgarin, C.,
Rengifo-Herrera, A.-J., Lavanchy, J.C., and Kiwi, J. (2011) Antimicrobial
Cu-functionalized surfaces prepared by
bipolar asymmetric DC-pulsed magnetron sputtering (DCP). J. Photochem.
Photobiol., A Chem., 220, 70–76.
Tsai, M.-Y., Huang, M.-S., Chen, L.-K.,
Shen, Y.-D., Lin, M.-H., Chiang, Y.-C.,
Ou, K.-L., and Ou, S.-F. (2013) Surface properties of copper-incorporated
diamond-like carbon films deposited by
hybrid magnetron sputtering. Ceram.
Int., 39, 8335–8340.
187. Chan, Y.-H., Huang, C.-F., Ou, K.-L.,
188.
189.
190.
191.
192.
193.
194.
195.
and Peng, P.-W. (2011) Mechanical
properties and antibacterial activity
of copper doped diamond-like carbon films. Surf. Coat. Technol., 206,
1037–1040.
Cioffi, N., Ditaranto, N., Torsi, L.,
Picca, R.A., Sabbatini, L., Valentini, A.,
Novello, L., Tantillo, G., Bleve-Zacheo,
T., and Zambonin, P.G. (2005) Analytical characterization of bioactive
fluoropolymer ultra-thin coatings modified by copper nanoparticles. Anal.
Bioanal. Chem., 381, 607–616.
Varghese, S., ElFakhri, S., Sheel, D.,
Sheel, P., Bolton, F., and Foster, H.
(2013) Antimicrobial activity of novel
nanostructured Cu-SiO2 coatings prepared by chemical vapour deposition
against hospital related pathogens.
AMB Express, 3, 1–8.
Hodgkinson, J.L., Massey, D., and
Sheel, D.W. (2013) The deposition
of copper-based thin films via atmospheric pressure plasma-enhanced
CVD. 19th European Conference
on Chemical Vapor Deposition
EuroCVD19, Varna, Bulgaria, September 1–6, 2013, vol. 230, pp. 260–265.
Palza, H., Gutiérrez, S., Delgado, K.,
Salazar, O., Fuenzalida, V., Avila, J.I.,
Figueroa, G., and Quijada, R. (2010)
Toward tailor-made biocide materials based on poly(propylene)/copper
nanoparticles. Macromol. Rapid Commun., 31, 563–567.
Delgado, K., Quijada, R., Palma, R.,
and Palza, H. (2011) Polypropylene
with embedded copper metal or copper oxide nanoparticles as a novel
plastic antimicrobial agent. Lett. Appl.
Microbiol., 53, 50–54.
Wu, B., Huang, R., Sahu, M., Feng,
X., Biswas, P., and Tang, Y.J. (2010)
Bacterial responses to Cu-doped TiO2
nanoparticles. Sci. Total Environ., 408,
1755–1758.
Singhal, S.K., Lal, M., Lata, Kabi, S.R.,
and Mathur, R.B. (2012) Synthesis of
Cu/CNTs nanocomposites for antimicrobial activity. Adv. Nat. Sci.: Nanosci.
Nanotechnol., 3, 045011 (10 pp).
Lal, M., Singhal, S.K., Sharma, I., and
Mathur, R.B. (2013) An alternative
211
212
8 Nano-Antimicrobials Based on Metals
196.
197.
198.
199.
200.
201.
202.
203.
204.
improved method for the homogeneous
dispersion of CNTs in Cu matrix for
the fabrication of Cu/CNTs composites.
Appl. Nanosci., 3, 29–35.
Sheikh, F.A., Kanjwal, M.A., Saran,
S., Chung, W.-J., and Kim, H. (2011)
Polyurethane nanofibers containing
copper nanoparticles as future materials. Appl. Surf. Sci., 257, 3020–3026.
Karlsson, H.L., Cronholm, P., Hedberg,
Y., Tornberg, M., De Battice, L.,
Svedhem, S., and Wallinder, I.O. (2013)
Cell membrane damage and protein
interaction induced by copper containing nanoparticles—Importance of the
metal release process. Nanotoxicology,
313, 59–69.
Kwon, Y.S., Ilyin, A.P., Tikhonov, D.V.,
Yablunovsky, G.V., and An, V.V. (2008)
Characteristics of nanopowders produced by wire electrical explosion of
tinned copper conductor in argon.
Mater. Lett., 62, 3143–3145.
Ozay, O., Akcali, A., Otkun, M.T.,
Silan, C., Aktas, N., and Sahiner, N.
(2010) P(4-VP) based nanoparticles
and composites with dual action as
antimicrobial materials. Colloids Surf. B
Biointerfaces, 79, 460–466.
Cady, N.C., Behnke, J.L., and
Strickland, A.D. (2011) Copper-based
nanostructured coatings on natural
cellulose: nanocomposites exhibiting rapid and efficient inhibition of a
multi-drug resistant wound pathogen,
A. baumannii, and mammalian cell
biocompatibility in vitro. Adv. Funct.
Mater., 21, 2506–2514.
Chattopadhyay, D.P. and Patel, B.H.
(2010) Effect of nanosized colloidal
copper on cotton fabric. J. Eng. Fibers
Fabr., 5, 1–6.
Zhong, T., Oporto, G., Jaczynski, J.,
Tesfai, A., and Armstrong, J. (2013)
Antimicrobial properties of the hybrid
copper nanoparticles-carboxymethyl
cellulose. Wood Fiber Sci., 45, 215–222.
Chowdhury, M.N.K., Beg, M.D.H.,
Khan, M.R., and Mina, M.F. (2013) Synthesis of copper nanoparticles and their
antimicrobial performances in natural
fibres. Mater. Lett., 98, 26–29.
Llorens, A., Lloret, E., Picouet, P., and
Fernandez, A. (2012) Study of the
205.
206.
207.
208.
209.
210.
211.
212.
antifungal potential of novel cellulose/copper composites as absorbent
materials for fruit juices. Int. J. Food
Microbiol., 158, 113–119.
Jia, B., Mei, Y., Cheng, L., Zhou, J., and
Zhang, L. (2012) Preparation of copper
nanoparticles coated cellulose films
with antibacterial properties through
one-step reduction. ACS Appl. Mater.
Interfaces, 4, 2897–2902.
Grace, M., Chand, N., and Bajpai, S.K.
(2009) Copper alginate-cotton cellulose
(CACC) fibers with excellent antibacterial properties. J. Eng. Fibers Fabr., 4,
1–14.
Grace, M., Bajpai, S.K., and Chand,
N. (2009) Copper (II) ions and copper
nanoparticles-loaded chemically modified cotton cellulose fibers with fair
antibacterial properties. J. Appl. Polym.
Sci., 113, 757–766.
Pinto, R.J.B., Neves, M.C., Neto, C.P.,
and Trindade, T. (2012) Growth and
chemical stability of copper nanostructures on cellulosic fibers. Eur. J. Inorg.
Chem., 2012, 5043–5049.
Bagchi, B., Kar, S., Dey, S.K., Bhandary,
S., Roy, D., Mukhopadhyay, T.K., Das,
S., and Nandy, P. (2013) In situ synthesis and antibacterial activity of
copper nanoparticle loaded natural
montmorillonite clay based on contact
inhibition and ion release. Colloids Surf.
B Biointerfaces, 108, 358–365.
Mallick, S., Sharma, S., Banerjee, M.,
Ghosh, S.S., Chattopadhyay, A., and
Paul, A. (2012) Iodine-stabilized Cu
nanoparticle chitosan composite for
antibacterial applications. ACS Appl.
Mater. Interfaces, 4, 1313–1323.
Chatterjee, A.K., Sarkar, R.K.,
Chattopadhyay, A.P., Aich, P.,
Chakraborty, R., and Basu, T. (2012)
A simple robust method for synthesis
of metallic copper nanoparticles of high
antibacterial potency against E. coli.
Nanotechnology, 23, 085103.
Usman, M.S., Zowalaty, M.E.E.,
Shameli, K., Zainuddin, N., Salama,
M., and Ibrahim, N.A. (2013) Synthesis,
characterization, and antimicrobial
properties of copper nanoparticles. Int.
J. Nanomed., 8, 4467–4479.
References
213. Bagchi, B., Dey, S., Bhandary, S., Das,
214.
215.
216.
217.
218.
219.
220.
221.
S., Bhattacharya, A., Basu, R., and
Nandy, P. (2012) Antimicrobial efficacy and biocompatibility study of
copper nanoparticle adsorbed mullite aggregates. Mater. Sci. Eng. C, 32,
1897–1905.
Giannossa, L.C., Longano, D.,
Ditaranto, N., Nitti, M.A., Paladini,
F., Pollini, M., Rai, M., Sannino, A.,
Valentini, A., and Cioffi, N. (2013)
Metal nanoantimicrobials for textile
applications. Nanotechnol. Rev., 2,
307–331.
Komeily-Nia, Z., Montazer, M., and
Latifi, M. (2013) Synthesis of nano
copper/nylon composite using ascorbic acid and CTAB. Colloids Surf.,
A Physicochem. Eng. Aspects, 439,
167–175.
Teli, M.D. and Sheikh, J. (2013)
Modified bamboo rayon–copper
nanoparticle composites as antibacterial textiles. Int. J. Biol. Macromol.,
61, 302–307.
Díaz-Visurraga, J., Daza, C., Pozo, C.,
Becerra, A., von Plessing, C., and
García, A. (2012) Study on antibacterial
alginate-stabilized copper nanoparticles by FT-IR and 2D-IR correlation
spectroscopy. Int. J. Nanomed., 7,
3597–3612.
Drelich, J., Li, B., Bowen, P.,
Hwang, J.-Y., Mills, O., and Hoffman, D.
(2011) Vermiculite decorated with copper nanoparticles: novel antibacterial
hybrid material. Appl. Surf. Sci., 257,
9435–9443.
Li, B., Bowen, P., Drelich, J., and
Villeneuve, B. (2012) Inexpensive mineral copper materials with antibacterial
surfaces. Surf. Innov., 1, 15–26.
Kamrupi, I.R. and Dolui, S.K. (2011)
Synthesis of copper–polystyrene
nanocomposite particles using water
in supercritical carbon dioxide medium
and its antimicrobial activity. J. Appl.
Polym. Sci., 120, 1027–1033.
Aina, V., Cerrato, G., Martra, G.,
Malavasi, G., Lusvardi, G., and
Menabue, L. (2013) Towards the controlled release of metal nanoparticles
from biomaterials: physico-chemical,
morphological and bioactivity features
222.
223.
224.
225.
226.
227.
228.
229.
230.
of Cu-containing sol–gel glasses. Appl.
Surf. Sci., 283, 240–248.
Bonici, A., Lusvardi, G., Malavasi, G.,
Menabue, L., and Piva, A. (2012) Synthesis and characterization of bioactive
glasses functionalized with Cu nanoparticles and organic molecules. J. Eur.
Ceram. Soc., 32, 2777–2783, Special
Issue ECerS XII, 12th Conference of
the European Ceramic Society.
Fritsche, A., Haenle, M., Zietz, C.,
Mittelmeier, W., Neumann, H.-G.,
Heidenau, F., Finke, B., and Bader, R.
(2009) Mechanical characterization of
anti-infectious, anti-allergic, and bioactive coatings on orthopedic implant
surfaces. J. Mater. Sci., 44, 5544–5551.
Zielecka, M., Bujnowska, E., Ke˛pska, B.,
Wenda, M., and Piotrowska, M. (2011)
Antimicrobial additives for architectural paints and impregnates. Prog. Org.
Coat., 72, 193–201.
Zhang, N., Gao, Y., Zhang, H., Feng,
X., Cai, H., and Liu, Y. (2010) Preparation and characterization of core–shell
structure of SiO2 @Cu antibacterial
agent. Colloids Surf. B Biointerfaces, 81,
537–543.
Chapman, J., Weir, E., and Regan, F.
(2010) Period four metal nanoparticles
on the inhibition of biofouling. Colloids
Surf. B Biointerfaces, 78, 208–216.
Ramyadevi, J., Jeyasubramanian, K.,
Marikani, A., Rajakumar, G., and
Rahuman, A.A. (2012) Synthesis
and antimicrobial activity of copper nanoparticles. Mater. Lett., 71,
114–116.
Singh, A., Krishna, V., Angerhofer, A.,
Do, B., MacDonald, G., and Moudgil,
B. (2010) Copper coated silica nanoparticles for odor removal. Langmuir, 26,
15837–15844.
Valodkar, M., Rathore, P.S., Jadeja, R.N.,
Thounaojam, M., Devkar, R.V., and
Thakore, S. (2012) Cytotoxicity evaluation and antimicrobial studies of starch
capped water soluble copper nanoparticles. J. Hazard. Mater., 201–202,
244–249.
Cárdenas, G., Díaz, V.J.,
Meléndrez, M.F., Cruzat, C.C., and
García Cancino, A. (2009) Colloidal Cu
nanoparticles/chitosan composite film
213
214
8 Nano-Antimicrobials Based on Metals
231.
232.
233.
234.
235.
236.
237.
238.
obtained by microwave heating for food
package applications. Polym. Bull., 62,
511–524.
Raspolli Galletti, A.M., Antonetti,
C., Marracci, M., Piccinelli, F., and
Tellini, B. (2013) Novel microwavesynthesis of Cu nanoparticles in the
absence of any stabilizing agent and
their antibacterial and antistatic applications. Appl. Surf. Sci., 280, 610–618.
Schwarz, F., Thorwarth, G., and
Stritzker, B. (2009) Synthesis of silver and copper nanoparticle containing
a-C:Hby ion irradiation of polymers.
E-MRS Spring Meeting 2008, vol. 11,
pp. 1819–1823.
Cioffi, N., Torsi, L., Ditaranto, N.,
Tantillo, G., Ghibelli, L., Sabbatini, L.,
Bleve-Zacheo, T., D’Alessio, M.,
Zambonin, P.G., and Traversa, E.
(2005) Copper nanoparticle/polymer
composites with antifungal and bacteriostatic properties. Chem. Mater., 17,
5255–5262.
Cioffi, N., Torsi, L., Ditaranto, N.,
Sabbatini, L., Zambonin, P.G.,
Tantillo, G., Ghibelli, L., D’Alessio,
M., Bleve-Zacheo, T., and Traversa, E.
(2004) Antifungal activity of polymerbased copper nanocomposite coatings.
Appl. Phys. Lett., 85, 2417–2419.
Cioffi, N., Torsi, L., Ditaranto, N.,
Picca, R.A., Tantillo, G., Sabbatini, L.,
and Zambonin, P.G. (2004) Synthesis analytical characterization and
bio-activity of Ag-and Cu-containing
nano-structured membranes. J. Appl.
Biomater. Biomech., 2, 1912–1918.
Ditaranto, N., Loperfido, S.,
van der Werf, I., Mangone, A., Cioffi,
N., and Sabbatini, L. (2011) Synthesis and analytical characterisation of
copper-based nanocoatings for bioactive stone artworks treatment. Anal.
Bioanal. Chem., 399, 473–481.
Theivasanthi, T. and Alagar, M. (2011)
Nano sized copper particles by electrolytic synthesis and characterizations.
Int. J. Phys. Sci., 6, 3662–3671.
Pandey, P., Merwyn, S., Agarwal,
G.S., Tripathi, B.K., and Pant, S.C.
(2012) Electrochemical synthesis
of multi-armed CuO nanoparticles
and their remarkable bactericidal
239.
240.
241.
242.
243.
244.
245.
246.
potential against waterborne bacteria.
J. Nanopart. Res., 14, 1–13.
Bärsch, N., Jakobi, J., Weiler, S., and
Barcikowski, S. (2009) Pure colloidal
metal and ceramic nanoparticles from
high-power picosecond laser ablation in
water and acetone. Nanotechnology, 20,
445603.
Barcikowski, S. and Compagnini, G.
(2013) Advanced nanoparticle generation and excitation by lasers in
liquids. Phys. Chem. Chem. Phys., 15,
3022–3026.
Hahn, A. and Barcikowski, S. (2009)
Production of bioactive nanomaterial
using laser generated nanoparticles.
JLMN-J. Laser Micro/Nanoeng., 4,
51–54.
Hahn, A., Brandes, G., Wagener, P.,
and Barcikowski, S. (2011) Metal ion
release kinetics from nanoparticle silicone composites. J. Controlled Release,
154, 164–170.
Hahn, A., Stöver, T., Paasche, G.,
Löbler, M., Sternberg, K., Rohm, H.,
and Barcikowski, S. (2010) Therapeutic
window for bioactive nanocomposites
fabricated by laser ablation in polymerdoped organic liquids. Adv. Eng. Mater.,
12, B156–B162.
Hahn, A., Gunther, S., Wagener,
P., and Barcikowski, S. (2011)
Electrochemistry-controlled metal
ion release from silicone elastomer
nanocomposites through combination of different metal nanoparticles.
J. Mater. Chem., 21, 10287–10289.
Longano, D., Ditaranto, N., Cioffi, N.,
DiNiso, F., Sibillano, T., Ancona, A.,
Conte, A., Nobile, M.A., Sabbatini, L.,
and Torsi, L. (2012) Analytical characterization of laser-generated copper
nanoparticles for antibacterial composite food packaging. Anal. Bioanal.
Chem., 403, 1179–1186.
Longano, D., Ditaranto, N., Cioffi, N.,
Di Niso, F., Sibillano, T., Mezzapesa,
F.P., Ancona, A., Conte, A., Del Nobile,
M.A., Sabbatini, L. et al. (2013) One- vs
two-step preparation of antimicrobial
coatings composed of laser ablated copper nanoparticles and poly-lactic acid.
MRS Online Proc. Lib., 1453, DOI:
10.1557/opl.2012.1044 (6 pp).
References
247. Conte, A., Longano, D., Costa, C.,
248.
249.
250.
251.
252.
253.
254.
255.
256.
257.
Ditaranto, N., Ancona, A., Cioffi, N.,
Scrocco, C., Sabbatini, L., Contò, F.,
and Del Nobile, M.A. (2013) A novel
preservation technique applied to
fiordilatte cheese. Innovative Food Sci.
Emerg. Technol., 19, 158–165.
Rai, M., Yadav, A., and Gade, A. (2008)
CRC 675—current trends in phytosynthesis of metal nanoparticles. Crit. Rev.
Biotechnol., 28, 277–284.
Wang, C. and Huang, Y. (2013) Green
route to prepare biocompatible and
near infrared thiolate-protected copper
nanoclusters for cellular imaging. Nano,
08, 1350054 (10 pp).
Barua, S., Das, G., Aidew, L.,
Buragohain, A.K., and Karak, N. (2013)
Copper-copper oxide coated nanofibrillar cellulose: a promising biomaterial.
RSC Adv., 3, 14997–15004.
Valodkar, M., Nagar, P.S., Jadeja, R.N.,
Thounaojam, M.C., Devkar, R.V., and
Thakore, S. (2011) Euphorbiaceae
latex induced green synthesis of
non-cytotoxic metallic nanoparticle solutions: a rational approach to
antimicrobial applications. Colloids
Surf., A Physicochem. Eng. Aspects, 384,
337–344.
Varshney, R., Bhadauria, S., Gaur, M.S.,
and Pasricha, R. (2010) Characterization of copper nanoparticles
synthesized by a novel microbiological
method. JOM, 62, 102–104.
Silva, A.R. and Unali, G. (2011)
Controlled silver delivery by
silver–cellulose nanocomposites prepared by a one-pot green synthesis
assisted by microwaves. Nanotechnology, 22, 315605 (6 pp).
Whitesides, G.M. (2005) Nanoscience,
nanotechnology, and chemistry. Small,
1, 172–179.
Fan, Z. and Lu, J.G. (2005) Zinc oxide
nanostructures: synthesis and properties. J. Nanosci. Nanotechnol., 5,
1561–1573.
Wang, Z.L. (2004) Zinc oxide nanostructures: growth, properties and
applications. J. Phys. Condens. Matter,
16, R829.
Wang, Z.L. (2004) Nanostructures of
zinc oxide. Mater. Today, 7, 26–33.
258. Meulenkamp, E.A. (1998) Synthesis and
259.
260.
261.
262.
263.
264.
265.
266.
267.
268.
growth of ZnO nanoparticles. J. Phys.
Chem. B, 102, 5566–5572.
Spanhel, L. and Anderson, M.A. (1991)
Semiconductor clusters in the solgel process: quantized aggregation,
gelation, and crystal growth in concentrated zinc oxide colloids. J. Am. Chem.
Soc., 113, 2826–2833.
Vijayakumar, P.S., Blaker, K.A., Wieting,
R.D., Wong, B., and Halani, A.T. (1988)
Chemical vapor deposition of zinc
oxide films and products. US 4751149
A.
Pierson, H.O. (1999) Handbook of
Chemical Vapor Deposition: Principles,
Technology and Applications, 2nd edn,
William Andrew.
Panigrahi, J., Behera, D., Mohanty, I.,
Subudhi, U., Nayak, B.B., and Acharya,
B.S. (2011) Radio frequency plasma
enhanced chemical vapor based ZnO
thin film deposition on glass substrate:
a novel approach towards antibacterial
agent. Appl. Surf. Sci., 258, 304–311.
Carcia, P.F., McLean, R.S., Reilly, M.H.,
and Nunes, G. (2003) Transparent ZnO
thin-film transistor fabricated by RF
magnetron sputtering. Appl. Phys. Lett.,
82, 1117–1119.
Bidmeshkipour, S. and Shahtahmasebi,
N. (2013) Different properties of
aluminum doped zinc oxide nanostructured thin films prepared by radio
frequency magnetron sputtering. Semiconductors, 47, 787–790.
Tsai, H.-Y. (2007) Characteristics of
ZnO thin film deposited by ion beam
sputter. J. Mater. Process. Technol.,
192-193, 55–59.
Qu, Y. (1993) Electrical and optical properties of ion beam sputtered
ZnO:Al films as a function of film
thickness. J. Vac. Sci. Technol. Vac. Surf.
Films, 11, 996–1000.
Jansson, T., Clare-Salzler, Z.J., Zaveri,
T.D., Mehta, S., Dolgova, N.V., Chu,
B.-H., Ren, F., and Keselowsky, B.G.
(2012) Antibacterial effects of zinc
oxide nanorod surfaces. J. Nanosci.
Nanotechnol., 12, 7132–7138.
Athauda, T.J., Ozer, R.R., and Chalker,
J.M. (2013) Investigation of cotton
functionalized with ZnO nanorods and
215
216
8 Nano-Antimicrobials Based on Metals
269.
270.
271.
272.
273.
274.
275.
276.
277.
its interaction with E. coli. RSC Adv., 3,
10662.
Nafchi, A.M., Alias, A.K., Mahmud, S.,
and Robal, M. (2012) Antimicrobial,
rheological, and physicochemical properties of sago starch films filled with
nanorod-rich zinc oxide. J. Food Eng.,
113, 511–519.
Paisoonsin, S., Pornsunthorntawee, O.,
and Rujiravanit, R. (2013) Preparation
and characterization of ZnO-deposited
DBD plasma-treated PP packaging film
with antibacterial activities. Appl. Surf.
Sci., 273, 824–835.
Sudheesh Kumar, P.T., Lakshmanan, V.K., Anilkumar, T.V., Ramya, C., Reshmi,
P., Unnikrishnan, A.G., Nair, S.V., and
Jayakumar, R. (2012) Flexible and
microporous chitosan hydrogel/nano
ZnO composite bandages for wound
dressing: in vitro and in vivo evaluation. ACS Appl. Mater. Interfaces, 4,
2618–2629.
Al-Balakocy, N.G., El-Badry, K.H., and
Hassan, T.M. (2013) Multi-finishing
of polyester and polyester cotton
blend fabrics activated by enzymatic
treatment and loaded with zinc oxide
nanoparticles. J. Appl. Sci. Res., 9,
2767–2776.
Znaidi, L. (2010) Sol–gel-deposited
ZnO thin films: a review. Mater. Sci.
Eng., B, 174, 18–30.
Yang, D. and Zhang, H. (2004) Fabrication and characterization of ZnO
nanostructures. Focus Nanotechnol.
Res., 1, 173–215.
Divya, M.J., Sowmia, C., Joona, K., and
Dhanya, K.P. (2013) Synthesis of zinc
oxide nanoparticle from hibiscus rosasinensis leaf extract and investigation of
its antimicrobial activity. Res. J. Pharm.
Biol. Chem. Sci., 4, 1137–1142.
Rajiv Gandhi, R., Gowri, S., Suresh,
J., and Sundrarajan, M. (2013) Ionic
liquids assisted synthesis of ZnO
nanostructures: controlled size, morphology and antibacterial properties.
J. Mater. Sci. Technol., 29, 533–538.
Subash, A.A., Chandramouli, K.V.,
Ramachandran, T., Rajendran, R., and
Muthusamy, M. (2012) Preparation,
characterization, and functional analysis of zinc oxide nanoparticle-coated
278.
279.
280.
281.
282.
283.
284.
285.
cotton fabric for antibacterial efficacy.
J. Text. Inst., 103, 298–303.
Vigneshwaran, N., Kumar, S., Kathe,
A.A., Varadarajan, P.V., and Prasad, V.
(2006) Functional finishing of cotton
fabrics using zinc oxide–soluble starch
nanocomposites. Nanotechnology, 17,
5087.
Uğur, Ş.S., Sarıışık, M., Aktaş, A.H.,
Uçar, M.Ç., and Erden, E. (2010) Modifying of cotton fabric surface with
Nano-ZnO multilayer films by layerby-layer deposition method. Nanoscale
Res. Lett., 5, 1204–1210.
Shateri-Khalilabad, M. and
Eh-Yazdanshenas, M. (2013) Biofunctionalization of cotton textiles by ZnO
nanostructures: antimicrobial activity
and ultraviolet protection. Text. Res. J.,
83, 993–1004.
Vigneshwaran, N., Bharimalla,
A.K., Prasad, V., Kathe, A.A., and
Balasubramanya, R.H. (2008) Functional behaviour of polyethylene-ZnO
nanocomposites. J. Nanosci. Nanotechnol., 8, 4121–4126.
Espitia, P.J.P., Soares, N.d.F.F., Teófilo,
R.F., Coimbra, J.S.d.R., Vitor, D.M.,
Batista, R.A., Ferreira, S.O., de
Andrade, N.J., and Medeiros, E.A.A.
(2013) Physical–mechanical and
antimicrobial properties of nanocomposite films with pediocin and ZnO
nanoparticles. Carbohydr. Polym., 94,
199–208.
Nafchi, A.M., Nassiri, R., Sheibani,
S., Ariffin, F., and Karim, A.A. (2013)
Preparation and characterization of
bionanocomposite films filled with
nanorod-rich zinc oxide. Carbohydr.
Polym., 96, 233–239.
Bajpai, S.K., Chand, N., and Chaurasia,
V. (2010) Investigation of water vapor
permeability and antimicrobial property of zinc oxide nanoparticles-loaded
chitosan-based edible film. J. Appl.
Polym. Sci., 115, 674–683.
Pantani, R., Gorrasi, G., Vigliotta, G.,
Murariu, M., and Dubois, P. (2013)
PLA-ZnO nanocomposite films: water
vapor barrier properties and specific
end-use characteristics. Eur. Polym. J.,
49, 3471–3482.
References
286. Li, X., Xing, Y., Jiang, Y., Ding, Y., and
287.
288.
289.
290.
291.
292.
293.
294.
Li, W. (2009) Antimicrobial activities
of ZnO powder-coated PVC film to
inactivate food pathogens. Int. J. Food
Sci. Technol., 44, 2161–2168.
Seo, J., Jeon, G., Jang, E.S.,
Bahadar Khan, S., and Han, H. (2011)
Preparation and properties of poly
(propylene carbonate) and nanosized
ZnO composite films for packaging
applications. J. Appl. Polym. Sci., 122,
1101–1108.
Baruah, S. and Dutta, J. (2009)
Hydrothermal growth of ZnO nanostructures. Sci. Technol. Adv. Mater., 10,
013001.
Raghupathi, K.R., Koodali, R.T., and
Manna, A.C. (2011) Size-dependent
bacterial growth inhibition and mechanism of antibacterial activity of zinc
oxide nanoparticles. Langmuir, 27,
4020–4028.
Ma, J., Liu, J., Bao, Y., Zhu, Z., Wang,
X., and Zhang, J. (2013) Synthesis of
large-scale uniform mulberry-like ZnO
particles with microwave hydrothermal
method and its antibacterial property.
Ceram. Int., 39, 2803–2810.
Femi, V., Prabha, P.H., Sudha, P., Jerald,
A.L., and Reni, A. (2011) Evaluation
of antibacterial potential of ZnO-Au
nanocomposites. South Asian J. Exp.
Biol., 1, 36–39.
İpeksaç, T., Kaya, F., and Kaya, C.
(2013) Hydrothermal synthesis of Zinc
oxide (ZnO) nanotubes and its electrophoretic deposition on nickel filter.
Mater. Lett., 100, 11–14.
Wysokowski, M., Motylenko, M.,
Stöcker, H., Bazhenov, V.V., Langer, E.,
Dobrowolska, A., Czaczyk, K., Galli,
R., Stelling, A.L., Behm, T. et al. (2013)
An extreme biomimetic approach:
hydrothermal synthesis of β-chitin/ZnO
nanostructured composites. J. Mater.
Chem. B, 1, 6469.
Reetz, M.T., Quaiser, S.A., Winter, M.,
Becker, J.A., Schäfer, R., Stimming, U.,
Marmann, A., Vogel, R., and Konno,
T. (1996) Nanostructured metal oxide
clusters by oxidation of stabilized metal
clusters with air. Angew. Chem. Int. Ed.
Engl., 35, 2092–2094.
295. Dierstein, A., Natter, H., Meyer,
296.
297.
298.
299.
300.
301.
302.
303.
F., Stephan, H.-O., Kropf, C., and
Hempelmann, R. (2001) Electrochemical deposition under oxidizing
conditions (EDOC): a new synthesis
for nanocrystalline metal oxides. Scr.
Mater., 44, 2209–2212.
Natter, H. and Hempelmann, R. (2003)
Tailor-made nanomaterials designed by
electrochemical methods. Electrochim.
Acta, 49, 51–61.
Shkurankov, A., Natter, H., and
Hempelmann, R. (2004) GDCh
Grundlagen und Anwendungen der
Elektrochemischen Oberflächentechnik,
GDCh-Monographien, Gesellschaft
Deutscher Chemiker (GDCh), pp.
247–255.
Huang, G.S., Wu, X.L., Cheng, Y.C.,
Shen, J.C., Huang, A.P., and Chu, P.K.
(2007) Fabrication and characterization
of anodic ZnO nanoparticles. Appl.
Phys. A, 86, 463–467.
Chandrappa, K.G. and Venkatesha, T.V.
(2012) Electrochemical synthesis and
photocatalytic property of zinc oxide
nanoparticles. Nano-Micro Lett., 4,
14–24.
Chandrappa, K., Venkatesha,
T., Vathsala, K., and
Shivakumara, C. (2010) A hybrid
electrochemical–thermal method for
the preparation of large ZnO nanoparticles. J. Nanopart. Res., 12, 2667–2678.
Venkatesha, T.G., Arthoba Nayaka, Y.,
Viswanatha, R., Vidyasagar, C.C., and
Chethana, B.K. (2012) Electrochemical
synthesis and photocatalytic behavior
of flower shaped ZnO microstructures.
Powder Technol., 225, 232–238.
Picca, R.A., Sportelli, M.C., Sabbatini,
L., and Cioffi, N. (2013) New insights
on the electrochemical development
of II generation nano-antimicrobials.
64th Annual Meeting of the International Society of Electrochemistry in
Queretaro, Mexico, p. 37.
Stypuła, B., Starowicz, M., Hajos, M.,
and Olejnik, E. (2011) Electrochemical synthesis of ZnO nanoparticles
during anodic dissolution of zinc in
alcohols solvents. Arch. Metall. Mater.,
56, 286–292.
217
218
8 Nano-Antimicrobials Based on Metals
304. Starowicz, M. and Stypuła, B. (2008)
305.
306.
307.
308.
309.
310.
Electrochemical synthesis of ZnO
nanoparticles. Eur. J. Inorg. Chem.,
2008, 869–872.
Banaś, J., Stypuła, B., Banaś, K.,
Światowska-Mrowiecka, J., Starowicz,
M., and Lelek-Borkowska, U. (2009)
Corrosion and passivity of metals in
methanol solutions of electrolytes.
J. Solid State Electrochem., 13,
1669–1679.
Gupta, A.K., Kashyap, V., Gupta, B.K.,
Nandi, S.P., Saxena, K., and Khare, N.
(2013) Synthesis of ZnO nanorods by
electrochemical deposition method and
its antibacterial activity. J. Nanoeng.
Nanomanuf., 3, 348–352.
Schrand, A.M., Rahman, M.F., Hussain,
S.M., Schlager, J.J., Smith, D.A., and
Syed, A.F. (2010) Metal-based nanoparticles and their toxicity assessment.
Wiley Interdiscip. Rev. Nanomed.
Nanobiotechnol., 2, 544–568.
Hajipour, M.J., Fromm,
K.M., Akbar Ashkarran, A.,
Jimenez de Aberasturi, D., de
Larramendi, I.R., Rojo, T., Serpooshan,
V., Parak, W.J., and Mahmoudi, M.
(2012) Antibacterial properties of
nanoparticles. Trends Biotechnol., 30,
499–511.
Foster, H.A., Ditta, I.B., Varghese,
S., and Steele, A. (2011) Photocatalytic disinfection using titanium
dioxide: spectrum and mechanism of
antimicrobial activity. Appl. Microbiol.
Biotechnol., 90, 1847–1868.
Allahverdiyev, A.M., Abamor, E.S.,
Bagirova, M., and Rafailovich, M.
(2011) Antimicrobial effects of TiO2
311.
312.
313.
314.
315.
316.
317.
and Ag2 O nanoparticles against
drug-resistant bacteria and leishmania parasites. Future Microbiol., 6,
933–940.
Visai, L., De Nardo, L., Punta, C.,
Melone, L., Cigada, A., Imbriani, M.,
and Arciola, C.R. (2011) Titanium
oxide antibacterial surfaces in biomedical devices. Int. J. Artif. Organs, 34,
929–946.
Kelson, A.B., Carnevali, M., and
Truong-Le, V. (2013) Gallium-based
anti-infectives: targeting microbial
iron-uptake mechanisms. Curr. Opin.
Pharmacol., 13, 707–716.
Wang, Q. and Webster, T.J. (2013)
Introducing antibacterial properties to
paper towels through the use of selenium nanoparticles. MRS Online Proc.
Lib., 1498, 21–26.
Tran, P.A. and Webster, T.J. (2013)
Antimicrobial selenium nanoparticle
coatings on polymeric medical devices.
Nanotechnology, 24, 155101.
Dutta, R.K., Nenavathu, B.P., and
Talukdar, S. (2014) Anomalous antibacterial activity and dye degradation by
selenium doped ZnO nanoparticles.
Colloids Surf. B Biointerfaces, 114,
218–224.
Mauter, M.S. and Elimelech, M. (2008)
Environmental applications of carbonbased nanomaterials. Environ. Sci.
Technol., 42, 5843–5859.
Kang, S., Pinault, M., Pfefferle, L.D.,
and Elimelech, M. (2007) Singlewalled carbon nanotubes exhibit strong
antimicrobial activity. Langmuir, 23,
8670–8673.
219
9
Natural Products as Antimicrobial Agents – an Update
Muhammad Saleem
9.1
Introduction
In the late 1920s, Alexander Fleming accidently discovered an antibiotic penicillin
from the fungus Penicillium notatum that completely killed the bacterium Staphylococcus aureus. Afterwards, several analogs of penicillin or other antibiotics were
discovered from natural sources or were synthesized in chemistry laboratories.
When natural or synthetic antibiotics became commercially available half a century ago, they were treated as miracle drugs. However, with the passage of time,
the antibiotics lost their effectiveness as bacteria evolved resistance against them.
Today, there is a variety of bacterial pathogens, which are resistant to most of the
available antibiotics. Old antibiotics are now not as effective at killing the microbes
as they were when first introduced. Moreover, since newer antibiotics are so similar to the older varieties, resistance in bacteria evolves very quickly. Therefore,
to overcome this life-threatening problem, continuous research to discover new
molecules with novel structural features is crucially important. Fortunately, pharmaceutical companies have already started to respond to this need and are focusing on natural products to develop new antibiotics.
Through evolutionary pressure, Nature has guided the production of an
immense diversity of organic molecules for a variety of biological purposes.
These molecules were the basis of most early traditional medicines and have
played an important role in drug discovery. Natural products continue to play
an important role in the discovery and development of new pharmaceuticals, as
clinically useful drugs, as starting materials to produce synthetic drugs, and or as
lead compounds from which a total synthetic drug is designed [1]. The majority
of new antibiotics were generated from natural products and/or from compounds
derived from natural products. Unfortunately, a couple of decades back, natural
product drug discovery programs declined as pharmaceutical companies shifted
to using synthetic chemical libraries because of the business risk in the field of
natural product drug development [2, 3].
Now, over the last decade, it has become clear that the popularity of synthetic
drugs is decreasing because of their toxicity and undesirable side effects. Further,
Novel Antimicrobial Agents and Strategies, First Edition.
Edited by David A. Phoenix, Frederick Harris and Sarah R. Dennison.
© 2015 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2015 by Wiley-VCH Verlag GmbH & Co. KGaA.
220
9 Natural Products as Antimicrobial Agents – an Update
the existing antimicrobial agents are losing their effectiveness as pathogens evolve
resistance against them. On the other hand, the new drugs only rarely reach the
market. Moreover, microbial pathogens can acquire drug resistance in a multitude
of ways, so getting around the resistance problem is not a straightforward matter.
To address these issues, pharmaceutical companies have revived efforts to develop
new antibiotics from natural sources. Mishra and Tiwari [4] have recently published a minireview on natural products under clinical trials including a number
of natural and semi-synthetic antimicrobial agents.
The discovery and commercial uses of streptomycin, gentamicin, omegamycin
griseofulvin (Likuden M®), the antibacterial terpenoid fusidic acid (Fucidine®),
semi-synthetic or synthetic penicillins and cephalosporins, ergotamine (ErgoKranit®) and chloramphenicol [5, 6], and many under clinical trial such as
antibacterial ceftaroline acetate (PPI-0903, TAK-599, 31) [7] have motivated
researchers and pharmaceutical industries to focus on natural sources to develop
new antimicrobial agents.
Antibacterial natural products played an important role in life-saving medicines
during the twentieth century; besides, they played a pivotal role in the growth of
biological chemistry as a discipline [8]. Several thousands of antimicrobial agents
have so far been discovered from many natural sources but only fewer have been
developed into a drug. The aim of writing this chapter on natural antimicrobial
agents is to list the natural molecules as future antimicrobial candidates and to
explore the diverse and indigenous sources to get more effective and less toxic
antimicrobial chemicals [9]. This chapter covers the literature between 2000 and
2012.
9.2
Antimicrobial Natural Products from Plants
9.2.1
Antimicrobial Alkaloids from Plants
Probably the first unidentified alkaloid was narcotine isolated from opium in 1803,
and the first identified medicinally important alkaloid morphine opened a new
research area, which resulted in the discovery of strychnine, emetine, brucine,
piperine, caffeine, quinine, colchicine, and coniine. Probably, coniine was the first
alkaloid, which, after identification, was synthesized [10]. However, modern techniques have now eased the isolation and identification of more complex alkaloids
with potential biological activities.
A steroidal alkaloid solanopubamine (3β-amino-5α, 22αH, 25βH-solanidan23β-ol, 1), has recently been reported from Solanum schimperianum in significant
yield. Its structure was established by IR, positive electron impact mass spectrometry (ESI-MS), and 1D and 2D NMR techniques, whereas -3β-NH2 and
-23β-OH groups were identified through methylation and acetylation. Compound
1 showed good antifungal activity only against Candida albicans and Candida
9.2
Antimicrobial Natural Products from Plants
tenuis with minimum inhibitory concentration (MIC) of 12.5 μg ml−1 [11].
Another antibacterial alkaloid (−)-latifolian A (2), isolated from the leaves of
Chinese vine Gnetum montanum, inhibited the growth of Pseudomonas aeruginosa with an inhibitory concentration (IC)50 value of 9.8 μM (MIC = 35 μM) [12].
These observations revealed that the selective and potential action of compounds
1 and 2 makes the strongest candidates for the development of antimicrobial
agents.
HO
OH
H2N
HO
N
HO
N
OH
1
OH
2
OH
Benzo[c]phenanthridine-type alkaloids (3–6) 8-hydroxydihydrosanguinarine
(hhS), isolated from aerial parts of Chelidonium majus Linn., inhibited the
growth of clinical strains of methicillin-resistant Staphylococcus aureus (MRSA)
with MICs/MBCs (minimal bactericidal concentrations) ranging from 0.49 to
15.63/1.95 to 62.50 μg ml−1 . The potent anti-MRSA activity of these compounds
is encouraging for the development of new antibiotics; however, a study on the
toxicity of these compounds is required [13].
O
O
N
R2O
OR1 R3
CH3
3: R1
R2 = CH2, R3 = OH
4: R1 = R2 = CH3, R3 = OH
5: R1
R2 = CH2, R3 = H
6: R1 = R2 = CH3, R3 = H
Ileabethoxazole (7), a perhydroacenaphthene-type diterpene alkaloid
containing the uncommon benzoxazole moiety, was isolated from a marine
plant Pseudopterogorgia elisabethae. This compound showed 92% inhibition
of Mycobacterium tuberculosis (H37 Rv) at the concentration range of 128 to
64 μg ml−1 [14]. The secondary metabolites from the roots of Piper sarmentosum,
brachyamide B (8) and sarmentosine (9) exhibited antifungal activity against
C. albicans with IC50 values 41.82 and 32.82, (μg ml−1 ) respectively [15]. Alkaloid
10, isolated from the culture extract of Janibacter limosus, at 30 μg/paper disk,
showed moderate activity against Bacillus subtilis (12 mm zone of inhibition),
Streptomyces viridochromogenes Tü 57 (19 mm zone of inhibition), and S. aureus
(15 mm zone of inhibition) [16].
221
222
9 Natural Products as Antimicrobial Agents – an Update
O
O
N
O
H
O
N
7
8
HO
HO
H
O
O
H
N
O
N
H
OH
O
9
10
Oliva, et al. [17] evaluated some alkaloids from Rutaceous plants for their
activities against plant pathogenic fungi. Of the tested metabolites, 11 and 12
inhibited the growth of Colletotrichum acutatum by 50 and 57%, respectively,
at 300 μM. Compounds 11–14 were observed to be moderately active on
Colletotrichum fragariae but 12 exhibited the most potent activity against
Colletotrichum gloeosporioides, with 67.7% inhibition at 100 μM. Compound 13
showed 50% inhibition at 300 μM, whereas at 300 μM, 11 and 14 were moderately
active against Fusarium oxysporum. Compounds 12 and 15 showed activity
against Botrytis cinerea at 300 μM, but only compound 12 had high activity at
30 μM. Compounds 12 and 15 exhibited 100% inhibition of Phomopsis obscurans
at a concentration of 30 μM. These data suggest that the alkaloids are important to
the Rutaceous plant in defense against plant pathogens and these kinds of natural
products can be potential leads for natural product-based fungicides for use
against plant pathogens of economic significance in the horticultural industry.
Canthin-6-one (16) and 5-methoxycanthin-6-one (17), the secondary metabolites present in the stem barks of Zanthoxylum chiloperone var. angustifolium,
have been identified as antifungal compounds. Compound 16 and its hydroxyl
derivative [(8-hydroxy-canthin-6-one (18)] were also isolated from Allium
neapolitanum [18]. Compound 16 exhibited a broad spectrum of activities
against Aspergillus fumigatus, Aspergillus niger, Aspergillus terreus, C. albicans,
Candida tropicalis, Candida glabrata, Cryptococcus neoformans, Geotrichum
candidum, Saccharomyces cerevisiae, Trichosporon beigelii, Trichosporon cutaneum, and Trichophyton mentagrophytes var. interdigitale with MIC values
between 5.3 and 46 μM l−1 . 5-Methoxy-canthin-6-one (17) was active only
against T. mentagrophytes var. interdigitale with a MIC value of 12.3 μM l−1 [19].
Indole-derived alkaloids 16 and 18 were also very active against a panel of
fast-growing Mycobacterium species with MICs 8–32 and 8–64 μg ml−1 against
multidrug-resistant (MDR) and methicillin-resistant strains of S. aureus. Methyl1H-pyrimidine-2,4-dione (19) isolated from the flowers of Alangium salviifolium
showed significant antibacterial activities against a number of Gram-positive and
Gram-negative pathogens. The MIC is reported to be within 64–128 μg ml−1 [20].
9.2
O
N
H
11
Antimicrobial Natural Products from Plants
OMe
O
N
8
O
MeO
O
MeO
N
4
12
O
13
O
O
O
N
N
O
N
H
O
4
14
O
O
R
16 R = H
17 R = OMe
15
O
N
H
N
N
O
N
OH
O
18
N
CH3
19
9.2.2
Antimicrobial Alkaloids from Microbial Sources
The presence of many microbial metabolites in the pharmaceutical market
indicates the potential of microorganisms as valuable sources of lead drugs – for
example, the antibiotic polyketide griseofulvin (Likuden M), the antibacterial
terpenoid fusidic acid (Fucidine), semisynthetic or synthetic penicillins and
cephalosporins, chloramphenicol as well as ergot alkaloids such as ergotamine
(Ergo-Kranit) [5].
N-methyl-1-hydroxyphenazine (20) was isolated from the culture extract of
Pseudomonas. This compound was found to be bacteriostatic at 0.5 mg l−1 against
Vibrio harveyi [21]. V. harveyi has been reported as the most important aquaculture pathogen with multiple antibiotic resistance, causing mass mortalities in
shrimp/prawn hatcheries. However, in vitro tests suggest that 20 was a potential
inhibitor that possesses toxicity in Penaeus monodon hemocyte culture and the
IC50 value was found to be 1.4 ± 0.31 mg l−1 [21].
O−
O
N
N
N
CH3
N
CH3
20
Phenazine antibiotics are synthesized by a number of bacteria from diverse
genera including Streptomyces, Pseudomonas, Pelagiobacter, and Vibrio [22–24].
223
224
9 Natural Products as Antimicrobial Agents – an Update
These microbes produce a range of phenazine compounds that differ widely in
antibiotic properties, according to the nature and position of side groups attached
to a phenazine nucleus [25]. For example, D-alanylgriseoluteic acid (21) is a
potent antimicrobial phenazine compound produced by Pantoea agglomerans
(Erwinia herbicola) Eh1087 and was isolated from the culture supernatant. Its
structure was determined by modern NMR techniques [26]. Susceptibility tests
against a range of microbes indicated that 21 exhibited a broad spectrum of
antimicrobial activity, particularly against Gram-positive pathogens, including :
many pneumococcal and MDR isolates with an MIC range of 0.06–0.75 μg ml−1 .
It was further established that 21 induced the save our soul (SOS) response
in Escherichia coli and slightly increased the frequency of GC-AT transition
mutations. The potency and broad spectrum of 21 activity suggest that it may
warrant further investigation and in future could have an application as a topical
agent [27]. A novel antifungal compound YM-215343 (22) was isolated from the
culture of Phoma sp. QN04621, which showed close relation to apiosporamide
and fischerin. YM-215343 (22) inhibited the growth of pathogenic fungi C.
albicans, C. neoformans, and A. fumigatus with MIC values of 2–16 μg ml−1 [28].
However, this compound has also been reported to be cytotoxic.
O
OMe
OH
O
OH O
OH
H
N
O
N
O
⊕
H3N
H
N
H
O
22
O
21
Me
Ajudazols A (23) and B (24) are the constituents of Chondromyces crocatus,
which were reported to possess antimicrobial activity in an agar diffusion assay.
With a concentration of 40 μg/disc in 20 μl of methanol, 24 inhibited growth of
the fungi B. cinerea (10 mm), Trichoderma koningii (21 mm), Gibberella fujikuroi
(17 mm), and Ustilago maydis (13 mm). It was weakly active against a few Grampositive bacteria with the MIC for Micrococcus luteus being 12.5 μg ml−1 . Ajudazol
A (23) showed only minor activity against a few fungi and Gram-positive bacteria
when compared with its analog 24 [29]. This difference revealed that the olefinic
function close to the oxazole ring in 24 might play an important role in exhibiting
antimicrobial activity.
O
N
OH
O
O
HO
O
O
N
CH3
23
9.2
Antimicrobial Natural Products from Plants
O
N
OH
O
N
O
HO
O
CH2
24
O
The fungus Bionectra byssicola F120 is reported to produce epidithiodioxopiperazines, bionectins A (25), B (26), C (27), and verticillin D (28). Compounds
25, 26, and 28 inhibited the growth of S. aureus including MRSA and quinoloneresistant Staphylococcus aureus (QRSA), with MIC values of 10–30 μg ml−1 ,
whereas 27 was nearly inactive [30]. Structural differences revealed that compounds 25, 26, and 28 have a disulfide bridge in their dioxopiperazine ring,
while 27 contains a dioxopiperazine ring with two methylsulfanyl groups. Mostly,
this class of compounds such as the verticillins, the leptosins, and Sch52900
are known to have antibacterial, nematocidial, antifungal, and antitumor activities [31–34]. The comparison revealed that the antibacterial activity of the
epidithiodioxopiperazines is mediated in part by the disulfide bridge in the
epidithiodioxopiperazine moiety.
H
N
H
N
O
OH
OH
N
N
H H SS
N
O
O
N
N
H H
O
HO
SCH3
O
N
SCH3
R
CH(OH)CH3
O
N S
H
S
N
N
OH
N
N
H H S S
N
O
25 R = H
26 R = —CH(OH)CH3
O
CH(OH)CH3
27
28
9.2.3
Antimicrobial Alkaloids from Marine Sources
Terrestrial natural sources have been in the focus in the search for new drug
candidates, although nearly 70% of the earth’s surface is covered by oceans [35].
The isolation of valuable amounts of prostaglandins from the gorgonian Plexaura
homomalla by Weinheimer and Spraggnis is considered as the starting point of
the discovery of drugs from marine sources [35] and as a result, new discoveries
have been increasingly reported from marine sources in the area of antibiotic
development.
Two diketopiperazine rodriguesines, A (29) and B (30) as a mixture of
homologs, were isolated from a marine invertebrate of the Genus Didemnum,
which could be identified by analysis of spectroscopic data including MS/MS
225
226
9 Natural Products as Antimicrobial Agents – an Update
experiments. Interestingly, the mixture of 29 and 30 proved to be potentially
active (MIC = 4.3–125.0 μg ml−1 ) against antibiotic-resistant strains S. aureus,
E. coli, Enterococcus faecalis, Streptococcus sanguinis, Streptococcus sobrinus,
and C. albicans. These pathogens are all associated with mucobuccal diseases.
The mixture was found be most active against P. aeruginosa P1 with MIC value
of 4.3 μg ml−1 . Notable is that this mixture of 29 and 30 did not show any
cytotoxicity [36]. On the basis of the reported findings, these compounds can be
potential candidates to treat mucobuccal diseases.
The mixture of aqabamycin E (31 and 32) and aqabamycin F (33) was isolated
from Vibrio species isolated from the surface of the soft coral Sinularia polydactyla. These maleimide derivatives exhibited antibacterial activity against B.
subtilis, M. luteus, E. coli, and Proteus vulgaris with MIC values ranging between
3.15 and 25 μg ml−1 [37].
R2
O
R1
O
N
HN
R3
N
H
n
NHR
R4
N
NO2
O
29: n = 1, R = H
30: n = 2, R = H
N
H
O
31: R1 = NO2, R2 = OH, R3, R4 = H
32: R1 = R2 = R3 = OH, R4 = NO2
33: R1 = R4 = NO2, R2 = R3 = OH
9.3
Antimicrobial Natural Products Bearing an Acetylene Function
Acetylenic function is not very common in nature; however, a few reports have
been found on antimicrobial natural products bearing acetylenic function, which
are mostly from fungal sources. The antifungal and antimicrobial properties of
fatty acids have been known for centuries. Survey of the previous pharmaceutical
market revealed that unsaturated fatty acids with olefin and/or acetylene functions are, in general, more potent against fungal pathogens [38]. The representative
unsaturated fatty acid with a single double bond at C-10, undecylenic acid (UDA),
is still on the market as a cost-effective antifungal agent and the active ingredient
of many topical over-the-counter antifungal preparations [39].
Interestingly, 6-acetylenic acids (34–38) purified by reversed-phase high
performance liquid chromatography (HPLC) were identified as potential
antimicrobial natural products. Their structures were established by liquid
chromatography-mass spectrometry (LC-MS), NMR, and HPLC-ESI-MS analyses. Compound 34 inhibited the growth of test pathogens with an MIC of
4.3–31.0 μM, 35 showed an MIC of 4.4–29.3 μM, and 36 was active with an MIC
of 2.1–5.6 μM, whereas 37 exhibited an MIC of 0.7–3.3 μM. Among these isolates,
compound 38 was not active. Compounds 39 and 40 also showed nonsignificant
activity. Compound 37 was the most active, in particular against the
9.3
Antimicrobial Natural Products Bearing an Acetylene Function
227
dermatophytes T. mentagrophytes and Trichophyton rubrum and the opportunistic pathogens C. albicans and A. fumigatus with MICs comparable to several control drugs [40]. In vitro toxicity testing against mammalian cell lines indicated that
none of the isolates was toxic at concentrations up to 32 μM. Taking into account
the low in vitro and in vivo toxicities and significant antifungal potencies, these 6acetylenic acids (34–40) may be excellent leads for further preclinical studies [40].
O
O
OH
34
OH
35
O
O
OH
36
OH
37
O
O
OH
38
H2
C
H2
C
40
O
H2
C
OH
39
OH
Phomallenic acids A–C (41–43) with acetylenic functions, isolated from a
leaf litter fungus Phoma sp., exhibited MIC of 7.8 and 3.9 μg ml−1 , respectively,
against wild-type S. aureus. Phomallenic acid C (42), an analog with the longest
chain, exhibited the best overall activity, and was superior to cerulenin (44) and
thiolactomycin (45), the two most studied and commonly used FabF inhibitors
[41]. In vitro antifungal testing demonstrated that the antifungal activity of the
acetylenic acids is associated with their chain lengths and position of triple bonds
[40]. Another non-acidic natural acetylenic compound (46) is obtained from a
sterile dark ectotrophic fungus isolated from the roots of an Australian native
grass, Neurachne alopecuroidea. Compound 46 totally inhibited the growth of
Phytophthora cinnamomi, Rhizoctonia solani, Pythium irregulare, and Alternaria
alternata at 0.98, 7.81, and 15.63 μg ml−1 , respectively [42].
H
H
C
H
OH
H
OH
O
41
O
42
OH
O
H
NH2
O
H
C
O
OH
43
44
CH3
HO
S
45
O
O
H
O
OH
46
O
228
9 Natural Products as Antimicrobial Agents – an Update
9.4
Antimicrobial Carbohydrates
Protection of foods from microbial spoilage using salt (salting by usually sodium
chloride) or sugar (corning by usually sucrose) has ancient roots. Some modern
synthetic preservatives have become controversial because of their allergic effects.
Every manufacturer adds preservative to the food during processing. The purpose
of all of these chemicals is either to kill microbes or to act as antioxidants or both
[43]. Natural products have the advantage over other chemicals because of their
low side effects. Among them, sugars are known for their antimicrobial properties and one potential example is natural honey. Honey has been reported to have
wound healing and antimicrobial properties, which is partially attributed to the
carbohydrate contents in honey [44].
Sugars have also been reported to be used for wound treatment because of
their antimicrobial properties. Many natural sources including plants contain
carbohydrates as their active constituents and bear antimicrobial properties. For
example, partially acetylated oligorhamnoside derivatives 47–49 were isolated
by capillary-scale NMR probe and LR-/high resolution electron spray ionizationmass spectrometry (HRESIMS) spectroscopic methods from Cleistopholis
patens. These compounds exhibit significant in vitro antibacterial activity against
the Gram-positive bacteria MRSA ATCC 33591 and S. aureus 78-13607A with
MICs of ≤16 μg ml−1 [45].
O CH2(CH2)10CH3
O
O
AcO
O
HO
O
OH
O
O
AcO
OH
O
O
R1O
O CH2(CH2)10CH3
O
O
R4O
R3O
OR2
OH
OH
O
AcO
OAc
O
HO
HO
OAc
48
47 R1 = R3 =Ac, R2 = R4 = H
49 R1 = R4 = Ac, R2 = R3 = H
9.5
Antimicrobial Natural Chromenes
Chromene (benzopyran) is one of the privileged medicinal pharmacophores,
which appears as an important structural component in natural compounds and
has generated great attention because of its interesting biological activity. It is
a heterocyclic ring system consisting of a benzene ring fused to a pyran ring.
Chromene constitutes the basic backbone of various types of polyphenols and
9.6
Antimicrobial Natural Coumarins
is widely found in natural alkaloids, tocopherols, flavonoids, and anthocyanins
[46]. It is known that certain natural and synthetic chromene derivatives possess
important biological activities including antimicrobial action [47–50]. Despite
the fact that this class of natural products is very important, only a few reports
have been published in recent literature concerning their roles as antimicrobial
agents. However, a review the literature over the last two decades revealed
that many of these compounds from natural sources are reported to possess
antimicrobial activity [51].
4H-chromene, uvafzlelin (50) isolated from the stem part of Uvaria ufielii,
shows broad spectrum antimicrobial activity against Gram-positive and acid-fast
bacteria [51d]. 5-Hydroxy-8-(30-methyl-20-butenyl)-2,2,7-trimethyl-2H-1chromene-6-carboxylic acid (51) and methyl-5-hydroxy-2,2,7-trimethyl-2H-1chromene-6-carboxylate (52) were isolated from the leaves of Peperomia serpens.
In a bioautographic microbial assay against Cladosporium cladosporioides and
Cladosporium sphaerospermum, 51 showed a detection limit of 10.0 μg, whereas
52 exhibited a limit of 20.0 μg. These results showed the antifungal potential of
these compounds [52].
O
O HO
O
O
O
OH
MeO
HO
O
O
O
50
OH
O
O
51
52
9.6
Antimicrobial Natural Coumarins
Coumarins (2H-1-benzopyran-2-one) constitute a large class of phenolic substances which occur in plants and are made of fused benzene and α-pyrone
rings [53]. More than 1300 coumarins have been isolated from plants, bacterial,
and fungal sources [54]. The biological activity of coumarins varies according
to their substitution patterns on the main nucleus. For example, substituted4-(1-piperazinyl) coumarins showed anti-platelet aggregation activity [55]
and 8-substituted-7-geranyloxycoumarin derivatives (specially the 8-methoxy
and 8-acetoxy derivative) exhibited anti-inflammatory activity [56], whereas
8-substituted-7-methoxycoumarins were found to be anticancer agents [57].
Over the last few years, many antimicrobial coumarins have also been reported
from various natural sources.
9.6.1
Antimicrobial Coumarins from Plants
Chakthong et al. [58] isolated several coumarins from the green fruits of Aegle
marmelos and evaluated their antimicrobial potential. Only compounds 53–55
229
230
9 Natural Products as Antimicrobial Agents – an Update
inhibited the growth of R. solani NPRC750, Rigidoporus microporus PR720, and
Sclerotium oryzae NPRC760 with an MIC/MFC (minimum fungicidal concentration) in the range of 7.8/7.8 to 31.5/31.5 μg ml−1 . Compound 56 was active only
against E. faecalis TISTR459 with an MIC of 18.75 μg ml−1 . The structure activity relationships of other coumarins included in this reference [58] are generally
poorly understood.
O
O
MeO
O
O
HO
OHC
MeO
O
O
53
O
HO
O
O
O
HO
O
55
54
O
56
The fruits of Angelica lucida produced imperatorin (57), isoimperatorin
(58), heraclenol (59), oxypeucedanin hydrate (60), and heraclenin (61). These
coumarins were studied for their antimicrobial activity against six Gram-positive
and -negative bacteria, two oral pathogens, and three human pathogenic fungi
[59]. The tested compounds were inactive against the three assayed Candida
species, whereas they generally inhibited the growth of all assayed bacterial
strains with MIC values ranging between 0.012 and 0.85 mg ml−1 . Compounds
57 and 58 showed the best activity with MIC values of 0.012 and 0.70 mg ml−1 .
Schinkovitz et al. [60] reported that the presence of an isoprenyl unit attached
to the carbocyclic ring favors their antimicrobial activities. Other similar results
revealed that the antibacterial potential can be attributed to the coumarin
ring. Since the natural products are known to exert their effects by inhibition
of bacterial nucleic acid synthesis [61] and the activity of furanocoumarins is
increased owing to the addition of a prenyl group to the main skeleton, the
antibacterial potential could be attributed to the increased lipophilicity of the
molecule, facilitating its passage though the thick bacterial membrane to its
target [62].
HO
O
O
O
O
O
57
O
O
58
O
O
O
O
O
O
O
O
OH
HO
OH
59
O
O
60
O
O
O
O
61
Aegelinol (62) and agasyllin (63), isolated from the roots of Ferulago campestris,
showed antibacterial activity against nine ATCC clinically isolated Gram-positive
and Gram-negative bacterial strains [63]. At concentrations ranging between 16
and 125 μg ml−1 both coumarins showed a significant antibacterial effect against
S. aureus, Salmonella typhi, Enterobacter cloacae, and Enterobacter aerogenes
9.6
Antimicrobial Natural Coumarins
(MIC = 16 and 32 μg ml−1 , respectively. Antibacterial activity was also found
against Helicobacter pylori: a dose-dependent inhibition was observed between
5 and 25 μg ml−1 . The growth inhibition by these two compounds is interesting
because the Gram-negative H. pylori is a microaerophilic-shaped bacterium that
is free living in the mucous layer of the human stomach. This bacterium spreads
worldwide, with a frequency ranging from 25% in the developed to 90% in the
developing areas. Infection from this bacterium leads to different clinical disorders
including chronic gastritis, peptic ulcer, and gastric adenocarcinoma [64].
HO
O
O
O
62
O
O
O
O
63
Hydroxycoumarin scopoletin (64) isolated from the seed kernels of Melia
azedarach L. showed a strong antifungal synergistic effect against Fusarium verticillioides. Although compound 64 exhibited weak activity (MIC of 1.5 mg ml−1 )
in the microbroth dilution method, when it was combined with vanillin (65),
4-hydroxy-3-methoxycinnamaldehyde (66), and (±) pinoresinol (67), it boosted
the antifungal effect of these compounds, even showing a complete inhibition in
the growth of the pathogen when 64 was added at a concentration of up to 5% of
its MIC value [65].
Two dihydroxylated coumarins, 7,8-dihydroxy-6-methoxycoumarin (68) and
7,8-dihydroxycoumarin (69), the secondary metabolites of Mexican tarragon
(Tagetes lucida Cv.), showed anti-E. coli and anti-Vibrio cholerae activity. Compound 68 is quantitatively more active than 69, and the MICs of compounds
68 and 69 against V. cholerae (Tor: strain CDC-V12) were 25 and 200 μg ml−1 ,
respectively, and against E. coli, they were 250 and 450 μg ml−1 , respectively [66].
This data indicated that the additional methoxyl group in 68 might be playing
an important role in antibacterial activity. Compounds 70 and 71 inhibited
the mycelial growth of Fusarium solani with IC50 values of 27.7 and 125 μM,
respectively, whereas 71 and 72 showed high activity against B. cinerea at 300 μM.
At a concentration of 30 μM, 70 and 71 inhibited P. obscurans growth by 80%
after 120 h of exposure [17].
Khatune et al. [67] have reported psoralidin (73), bakuchicin (74), psoralin (75),
and angelicin (76), isolated from the seeds of Psoralea corylifolia, which were
found to exhibit significant antibacterial activities against a number of Grampositive and Gram-negative bacteria. At concentrations of 200 and 400 μg/disc,
all substances exhibited 9–27 mm zone of inhibition against Gram-positive
pathogens, whereas against Gram-negative bacteria, at the same concentration,
they showed 9–23 mm zone of inhibition.
231
232
9 Natural Products as Antimicrobial Agents – an Update
O
H
O
MeO
H
HO
O
OMe
MeO
MeO
MeO
O
OH
OH
65
64
OH
O
O
HO
67
66
MeO
OMe
HO
O
O HO
OH
O
OH
68
O
O
O
O
O
69
70
O
O
O
MeO
O
OMe
71
72
OH
O
O
O
HO
O
O
O
O
O
O
O
O
O
73
74
75
76
9.6.1.1 Antimicrobial Coumarins from Bacteria
Two antifungal coumarins, 5,7-dimethoxy-4-p-methoxylphenylcoumarin (77)
and 5,7-dimethoxy-4-phenylcoumarin (78), were isolated from the culture extract
of Streptomyces aureofaciens CMUAc130. Both the compounds were purified by
silica gel-column chromatography and were identified by NMR and mass-spectral
data. Compounds 77 and 78 inhibited the growth of phytopathogenic fungi with
MICs values 120 and 150 μg ml−1 , respectively [68].
R
77: R = OCH3
OCH3
H3CO
O
78: R = H
O
9.7
Antimicrobial Flavonoids
Flavonoids are amongst the biggest classes of known secondary metabolites and
are distributed in various plant families. They protect the plants from UV radiation and other environmental stress; thus, prominent antioxidant properties are
9.7
Antimicrobial Flavonoids
233
associated with them. Flavonoids have been found in vitro to be effective antimicrobial agents against a wide array of microorganisms [69]. Common flavonoids
or flavonoid glycosides such as luteolin, quercetin, apigenin, kaempferol, isorhamnetin, acacetin, tamarixetin, chrysin, and galangin [70], quercetin-3-O-rutinoside,
kaempferol-3-O-rutinoside, and isorhamnetin-3-O-rutinoside [69] are known to
possess antimicrobial properties. Not only natural but synthetic flavonoids are
also reported to exhibit antimicrobial activities [71].
9.7.1
Antimicrobial Flavonoids from Plants
Two flavonoids, 5,7-dihydroxy-4′ ,6,8-trimethoxyflavone (79) and 5,6-dihydroxy4′ ,7,8-trimethoxyflavone (80), isolated from Limnophila heterophylla Benth.
and Limnophila indica (Linn.) Druce, respectively, exhibited moderate but
broad-spectrum antimicrobial activity against B. subtilis, S. aureus, E. coli,
Salmonella typhimurium, Alternaria solani, and C. albicans [72]. It is stated that
the compounds probably acted either on the cell wall integrity or on the process
of energy metabolism to halt the microbial growth. This report indicates that
these compounds can be used as food preservatives and as possible therapeutics
against microbial infections.
Anticancer prenylated flavonoids, propolin C (81), propolin D (82), propolin G
(83), and propolin H (84), were identified as the constituents of propolis, a natural
product from the beehive traditionally used in folk medicine for its antimicrobial
properties. These compounds exhibited anti-MRSA activity against S. aureus with
MIC values ranging between 8 and 32 mg l−1 [73].
OH
OH
OMe
OMe
HO
OMe
HO
MeO
OH O
79
80
81
OH
OH
OH
OH
O
OH O
82
O
OH O
OH O
HO
HO
O
MeO
O
OMe
HO
O
OH O
83
OH
HO
O
OH O
84
234
9 Natural Products as Antimicrobial Agents – an Update
Dihydrokaempferol (85) and 8-C-glucopyranosylapigenin (86) isolated from
the extract of Trilepisium madagascariense exhibited antimicrobial activity
against Providencia smartii, P. aeruginosa, Klebsiella pneumoniae, S. aureus, S.
typhi, E. coli, and C. albicans. Varying MIC values are reported between 8 and
32 μg ml−1 . However, many other flavonoids were also evaluated for antimicrobial
potential and possessed moderate to low activity [74].
Luteolin (87), a common phytochemical, was isolated from the leaf of Struchium
sparganophora (Linn) Ktze. This compound inhibited the growth of the Grampositive organism S. aureus and the Gram-negative organism Klebsiella aerogenes
with MIC values ranging from 100 to 6.25 μg ml−1 [75].
OH
O
OH HO
HO
OH
OH
O
HO
O
HO
OH
HO
O
OH
OH
OH O
OH O
85
OH O
87
86
J. Y. Lee [76] studied the antimicrobial potential of 3,6-dihydroxyflavone
(88) and the chalcone-derived compound 3-(4-hydroxyphenyl)-1-(2,4,6-trihydroxyphenyl)propan-1-one (89). Owing to a high binding affinity the two
compounds are reported to be potent inhibitors of β-ketoacyl-acyl carrier
protein synthase III (KAS III). In particular, these compounds displayed excellent
antimicrobial effect against S. aureus and MRSA in the range of 16–32 μM.
HO
O
OH
OH
OH
HO
O
88
OH O
89
Two flavonols, lavandulylated flavanones (2R 3R)-8-lavandulyl-2′ -methoxy5,7,4′ -trihydroxy-flavanonol (90) and 8-lavandulyl-5,7,4′ -trihydroxyflavonol (91),
were identified as the constituents of Sophora flavescens. Both the compounds
exhibited significant antibacterial activities against S. aureus and B. subtilis
that equaled that of chloramphenicol. Each compound was tested at a dose
of 0.2 ml of 100 μg ml−1 and exhibited a zone of inhibition in the rage of
16–20 mm [77]. Amentoflavone (92), another flavonoid, derived from Selaginella
tamariscina, exhibited potent inhibitory activity against the fungal pathogens
C. albicans (MIC = 5.0 μg ml−1 ), S. cerevisiae (MIC 5.0 = μg ml−1 ), and T. beigelii
(MIC = 5.0–10.0 μg ml−1 ), indicating its broad-spectrum antibiotic potential [78].
9.7
Antimicrobial Flavonoids
235
OH
OH
CH2
HO
O
OH
O
HO
O
O
OH
HO
OH O
OH
CH2
HO
O
OH
OH
OH O
OH O
91
92
90
The oxidized products of quercetin 2-(3,4-dihydroxyphenyl)-4,6-dihydroxy-2methoxybenzofuran-3-one (93) and 3-(quercetin-8-yl)-2,3-epoxyflavanone (94)
were isolated from the water extract of onion (Allium cepa) skin. Compound
93 selectively inhibited (100 μg ml−1 , zone inhibition 12–15 mm) the growth of
Gram-negative bacteria, primarily H. pylori strains and was inactive against other
tested Gram-positive pathogens. This information suggested that compound 93
has some specific mechanism of action against H. pylori. Compound 94 showed
comparatively the highest activity (100 μg ml−1 , zone inhibition 15–18 mm)
against MRSA strains and also high activity against H. pylori strains. In addition,
94 increased the antibacterial effect against both strains of MRSA in the presence
of oxacillin, suggesting that this compound has a synergistic growth inhibitory
effect with the β-lactam against the antibiotic-resistant bacteria [79].
OH
OH
OH
HO
OH
HO
O
OH
OH O
OMe
O
94
93
HO
O
OH O
OH
O
OH OH
95
O
O
O
OH O
OH O
OH
O
OH
OH
HO
HO
O
O
OH
O
HO
O
O
HO
O
OH
OCH3
O
HO
O
H
OH
OH O
96
OH
OH
HO
HO
97
98
Morin-3-O-α-L-lyxopyranoside (95) and morin-3-O-α-L-arabopyranoside
(96) were identified as metabolites of the leaves of guava (Psidium guajava L.).
The structures of these compounds were established on the basis of chemical
and spectroscopic evidence. The MIC of 95 and 96 was 200 μg ml−1 for each
236
9 Natural Products as Antimicrobial Agents – an Update
against Salmonella enteritidis, and 250 and 300 μg ml−1 against Bacillus cereus,
respectively [80].
Donnell et al. [81] identified homoisoflavanone (97) as 3-(4′ -methoxybenzyl)7,8-methylenedioxy-chroman-4-one, which was found to be an active agent
against a fast-growing mycobacteria. It exhibited MIC values ranging from 16
to 256 μg ml−1 . Kaemferol (98), another very common phytochemical, has been
reported as an antimicrobial. In an antibacterial test, it inhibited the growth of
two Gram-negative and four Gram-positive pathogens with an MIC of 2.4 μg ml−1
and showed activity against Candida glabrata (MIC = 4.8–9.7 μg ml−1 ) [82].
Apigenin (99) and luteolin (100) from Scutellaria barbata D. Don are also
common phytochemicals. Apigenin (99) potentially inhibited (MIC = 3.9–
15.6 μg ml−1 ) the growth of 20 strains of MRSA, whereas luteolin (100) exhibited
a moderate activity (MIC = 62.5–125 μg ml−1 ) against the tested pathogens [83].
Flavone glycoside (101) purified from the leaves of Vitex negundo is also known
as an antimicrobial agent. Compound 101 showed promising activity against T.
mentagrophytes and C. neoformans with MIC values of 6.25 μg ml−1 as compared
to the standard antifungal drug fluconazole having MIC of 2.0 μg ml−1 [84].
Di-benzyloxy flavone (102) from Helichrysum gymnocomum was characterized
by NMR and MS. It afforded high activity against C. neoformans ATCC 90112
with an MIC value of 7.8 μg ml−1 [85].
Laburnetin (103) isolated from the methanolic extract of Ficus chlamydocarpa
exhibited potent inhibitory activity against Mycobacterium smegmatis and M.
tuberculosis with MIC values of 0.61 and 4.88 μg ml−1 , respectively [86]. The
efficacy of flavonoids against various pathogens can be attributed to cell-wall
permeability and the porins in the outer membrane present in microorganisms.
It appears that the compounds may block the charges of amino acids in porins
[87]. Further, a keen comparison of the above-discussed properties revealed that
the flavonoids having free hydroxyl groups at C-5 and C-7 in ring A are more
active and are also supported by their synthetic analogs [88].
O
OH
OH
OH
HO
HO
HO
OMe
O
O
OH
HO
O
O
HO
O
OH O
OH O
99
OH O
101
100
HO
O
O
O
O
102
O
OH
OH O
103
OH
OH
9.8
Antimicrobial Iridoids
237
9.8
Antimicrobial Iridoids
Iridoids are widely distributed in dicotyledonous plant families such as Apocynaceae, Scrophulariaceae, Diervillaceae, Lamiaceae, Loganiaceae, and Rubiaceae.
These natural products are associated with a wide range of health benefits. They
provide support against a broad range of physical, chemical, or biological stressors
without exhibiting any toxicity. These compounds possess anticancer [89], antiinflammatory [90], and anti-aging [91] activities. They are also known for their
other potential biological activities, and recent developments revealed that they
also possess antimicrobial properties [92].
9.8.1
Antimicrobial Iridoids from Plants
Two iridoids (104 and 105) isolated from the roots of Patrinia rupestris (Pall.)
were identified as potential antibacterial drug candidates. They were evaluated to
exhibit strong antibacterial activities (inhibition zone ranging from 13 to 20 mm
at a concentration of 100 μg ml−1 ) against E. coli and S. aureus, respectively [93]. A
study of the chemistry and antibacterial activity of Scrophularia deserti also led to
the isolation of ajugoside (106) and scropolioside B (107) as antibacterial agents.
These compounds moderately inhibited the growth of multidrug-resistant Staphylococcus aureus (MDRSA) and MRSA and a panel of rapidly growing mycobacteria
(Mycobacterium fortuitum, Mycobacterium phlei, Mycobacterium aurum, and M.
smegmatis) with MIC values ranging from 32 to 128 μg ml−1 [94].
OR
HO
O
HO
ClH2C
O
O
OR
104 R = COCH2(CH3)2
OR
O
O
O
HO
O
OR
105
O
107
106
O
O
OH
OH
HO
OH
OH
HO
O
O
HO
ClH2C
O
O
OH
OH
RO
O
O
O
O
O
O
OH
Two iridoid glycosides (108 and 109) purified from alcoholic extract of leaves
of V. negundo were characterized by 1D, 2D NMR and fast atomic bombardmentmass spectrometry (FABMS) techniques. Compound 108 inhibited the growth of
C. neoformans and T. mentagrophytes exhibiting an MIC of 12.5 μg ml−1 , whereas
109 displayed promising activity against T. mentagrophytes and C. neoformans
with an MIC of 6.25 μg ml−1 [84].
238
9 Natural Products as Antimicrobial Agents – an Update
H
HO
COOH
H
O
O
O
H
HO
O
O
O
OH
O
HO
HO
H
HO
O
O
OH
OH
108
OH
OH
O
109
HO
9.9
Antimicrobial Lignans
Lignans are a class of phytochemicals produced by biological oxidative dimerization of two phenylpropanoid units [95]. In addition to their normal biological
roles, lignans also possess significant pharmacological activities, including antitumor, anti-inflammatory, immunosuppressive, cardiovascular, antioxidant, and
antiviral actions [96].
9.9.1
Antimicrobial Lignans from Plants
Heyneanol A (110) was obtained from the root extract of Vitis sp. (grape vines)
and was identified as an antibacterial agent against Gram-positive pathogens.
Using the disc diffusion method, the compound 110 exhibited an MIC value
of 2.0 μg ml−1 toward MRSA and a value of 2.0–4.0 μg ml−1 for Enterococcus
faecium, S. aureus, Streptococcus agalactiae, and Streptococcus pyogenes [97].
OH
O
OH
HO
HO
HO
O
OH
OH
HO
110
O
OH
Another lignin (+)-lyoniresinol-3α-O-β-D-glucopyranoside (111) was identified as the constituent of the root bark of Lycium chinense Miller. This compound
exhibited potent antimicrobial activity against antibiotic-resistant bacterial
strains, such as MRSA (MIC 2.5–5.0 μg ml−1 ), and human pathogenic fungi
such as C. albicans (MIC 5.0 μg ml−1 ), S. cerevisiae (MIC = 5.0 μg ml−1 ), and
9.9
Antimicrobial Lignans
T. beigelii (MIC = 10.0 μg ml−1 ), without having any hemolytic effect on human
erythrocytes. In particular, compound 111 induced the accumulation of intracellular trehalose on C. albicans as stress response to the drug, and disrupted the
dimorphic transition that forms the pseudohyphae caused by pathogenesis. This
indicates that (+)-lyoniresinol-3-α-O-β-D-glucopyranoside (111) has an excellent
potential as a lead compound for the development of antibiotic agents [98].
A resveratrol trimer with an ortho-quinone nucleus, hopeanolin (112), was
isolated from the stem bark of Hopea exalata. Compound 112 demonstrated
antifungal activity in the MIC range of 0.1–22.5 μg ml−1 against Alternaria
altenata (22.5 μg ml−1 ), A. solani (1.56 μg ml−1 ), Colletotrichum lagenarium
(10.6 μg ml−1 ), F. oxysporum f. sp. Vasinfectum (6.22 μg ml−1 ), Pyricularia
oryzae (0.10 μg ml−1 ), and Valsa mali (1.55 μg ml−1 ). These results revealed the
broad-spectrum antifungal properties of the compound and made it a potential
antimicrobial drug candidate [99]. The hexahydroxydiphenoyl ester vescalagin
(113) was isolated from Lythrum salicaria as the active principal of the antibacterial activity. Minimal inhibitory and bactericidal concentration (MIC and MBC;
0.062–0.125 and 0.50–1.0 mg ml−1 , respectively) were observed in the isolated
compounds [100].
HO
MeO
O
OH HO
O
HO
OH
OH
O
OMe
OH
H
HO
H
H
HO
MeO
H
O
O
OMe
O
OH
O
111
HO
112
HO
OH
HO
OH
OH
HO
O
O
O
O
H
O
O
O O
HO
OH
OH
O
OH
O
OH
HO
OH
OH
HO
113
OH
H
OH
239
240
9 Natural Products as Antimicrobial Agents – an Update
9.10
Antimicrobial Phenolics Other Than Flavonoids and Lignans
Antimicrobial activity of plant phenolics has been intensively studied, and, in
addition to controlling invasion and growth of plant pathogens, their activity
against human pathogens has been investigated to characterize and develop new
healthy food ingredients, medical compounds, and pharmaceuticals [101–103].
9.10.1
Antimicrobial Phenolics from Plants
The leaves of Piper regnellii produced eupomatenoid-3 (114), eupomatenoid-5
(115), eupomatenoid-6 (116), and conocarpan (117). The compounds 115 and
116 showed good activity against S. aureus with an MIC of 3.12 and 1.56 μg ml−1 ,
respectively. Both compounds demonstrated an MIC of 3.12 μg ml−1 against B.
subtilis, while 117 was potentially active against S. aureus and B. subtilis with
an MIC of 6.25 μg ml−1 [104]. Eupomatenoid-5 (115) showed strong activity on
T. rubrum with an MIC value of 6.2 μg ml−1 . The results showed that the plant
could be explored for possible antifungal agents and provide preliminary scientific
validation for its traditional medicinal use [105]. The low or absence of activity of 118–120 and structural comparison of 114–120 and their activity levels revealed that phenolic functions (free hydroxyl group on benzene ring) are
important for the compound to be antimicrobial as was observed in the case of
flavonoids.
O
R1
O
R
O
O
114
O
R
115: R OH, R1 = OMe
116: R = OH, R1 = H
118: R = OMe, R1 = H
119: R = OMe, R1 = OMe
117: R = OH
120: R = OMe
Other antimicrobial phenolics include pterocarpans – erybraedin B (121), erybraedin A (122), phaseollin (123), erythrabyssin II (124), erystagallin A (125),
erythrabissin-1 (126), and erycristagallin (127) – isolated from the stems of Erythrina subumbrans (Leguminosae). Their structures were identified by means of
spectroscopy. Compounds 122 and 124 exhibited the highest degree of activity
against Streptococcus strains with an MIC range of 0.78–1.56 μg ml−1 , whereas
compound 127 was found to be potent against Staphylococcus strains, including
drug-resistant strains (MRSA and vancomycin-resistant Staphylococcus aureus
(VRSA)), with an MIC range of 0.39–1.56 μg ml−1 . Compounds 122, 124, 125,
9.10
Antimicrobial Phenolics Other Than Flavonoids and Lignans
241
and 127 potentially inhibited the growth of several strains of Streptococcus and
Staphylococcus and were found to be more active than the standard antibiotics
vancomycin and oxacillin [106]. In addition, 127 showed the highest level of activity against all VRSA strains tested, with an MIC range of 0.39–1.56 μg ml−1 . These
compounds may prove to be potent phytochemical agents for antibacterial activity, especially against the MRSA and VRSA strains [106]. Structure–activity relationship analysis of these compounds revealed that the dimethylallyl units and
higher phenolic characteristic play an important role in activity, and this could be
the reason that compounds 121 and 123 are nearly inactive.
HO
O
HO
H
H
O
H
O
125
O
OH
124
O
HO
O
HO
OH
O
123
122
O
H
O
OH
O
121
HO
H
O
H
H
H
H
O
HO
O
HO
O
OH
OMe
H
O
126
O
OMe
127
Kanzonols are well-known phytophenolics possessing antimicrobial activities
[107]; for example, kanzonol C (128) is known for its high biological activities. Compound 128 and its analogs, isobavachalcone (129), stipulin (130),
4-hydroxylonchocarpin (131) isolated from the extract of the twigs of Dorstenia
barteri and characterized spectroscopically showed broad-spectrum inhibitory
activities against Gram-positive and Gram-negative bacteria and fungi. Compound 129 was highly active against E. cloacae, Streptococcus faecalis, S. aureus,
Bacillus stearothermophilus, C. albicans and C. glabrata (MIC = 0.3 μg ml−1 ),
Enterobacter aerogenes, Morganella morganii, Shigella flexneri, B. cereus, Bacillus
megaterium and B. subtilis (MIC = 0.6 μg ml−1 ), Proteus mirabilis, P. vulgaris,
Microsporum audorium, and T. rubrum (MIC = 1.2 μg ml−1 ), whereas compound
131 inhibited growth of E. cloacae, M. morganii, B. megaterium and B. stearothermophilus (MIC = 1.2 μg ml−1 ), E. aerogenes, S. flexneri, S. faecalis, S. aureus, B.
cereus, B. subtilis, C. albicans, C. glabrata, and T. rubrum (MIC = 4.9 μg ml−1 )
and M. audorium (MIC = 9.8 μg ml−1 ). Phenolic 128 showed comparatively
moderate activity against E. aerogenes, E. cloacae, M. morganii, S. flexneri, S.
faecalis, B. megaterium, B. stearothermophilus, C. albicans and C. glabrata (MIC,
OH
242
9 Natural Products as Antimicrobial Agents – an Update
4.9 μg ml−1 ), P. mirabilis, P. vulgaris, B. cereus, B. subtilis, and M. audorium
(MIC = 9.8 μg ml−1 ). Although the structure–activity relationship has not been
discussed for these compounds, another compound 130 from the same source
was found to be inactive. The only structural difference between these compounds
is the position of the isoprene unit, which is the same in 128 and 129 but different
in 130. This difference in structure and activities revealed that the position of
the isoprene unit on the phenyl nucleus is important for such a compound to be
antimicrobial [107].
Curcumin (132), a similar type of compound, has been reported to induce filamentation in B. subtilis, suggesting that it inhibits bacterial cytokinesis. Further,
132 strongly inhibited the formation of the cytokinetic Z-ring in B [108]. On the
basis of these facts, the site of action of kanzonols was predicted but further study
on these antimicrobials is needed to confirm findings.
O
HO
HO
O
OH HO
HO
128
OH HO
HO
HO
O
O
O
OH
HO
OMe
OMe
131
OH
130
129
O
HO
O
132
Structural similarities and antimicrobial activity levels were observed in
acylphloroglucinol (133) and its analog, 2-methyl-1-[2,4,6-trihydroxy-3-(2hydroxy-3-methyl-3-butenyl)phenyl]-1-propanone (134) isolated from H.
gymnocomum, and characterized by NMR and mass spectroscopic means. Compound 133 displayed high activity against C. neoformans ATCC 90112 with an
MIC value of 7.8 μg ml−1 , while 134 inhibited the growth of S. aureus ATCC 12600
(MIC = 6.8 μg ml−1 ), E. faecalis ATCC 29212, S. aureus (methicillin and gentamycin resistant) ATCC 33592 and B. cereus ATCC 11778 (MIC = 7.8 μg ml−1 ),
and Staphylococcus epidermidis ATCC 2223 (MIC = 9.8 μg ml−1 ) [85]. Other
phytochemicals, grandinol (135) and jensenone (136) and their synthetic analog
(137), also showed high antimicrobial activities. The comparative data revealed
that phloroglucinols possess various bioactivities and their acyl-derivatives
possess antimicrobial activities; therefore, they can be good candidates for
further clinical studies to develop new antimicrobial drugs [109].
9.10
OH
Antimicrobial Phenolics Other Than Flavonoids and Lignans
OH
OH
O
H
OH
HO
OH
HO
HO
O
OH
O
133
OH O
134
O
HO
135
O
H
OH
OHC
OH O
H
OH
HO
OH
OH O
136
137
The stem bark of Irvingia gabonensis contains 3,3′ ,4′ -tri-O-methylellagic acid
(138) and 3,4-di-O-methylellagic acid (139). The lowest MIC value (9.76 μg ml−1 )
of 138 and 139 was observed against E. coli, P. vulgaris, and B. subtilis, while
139 exhibited high activity against P. aeruginosa (MIC = 4.8 μg ml−1 ) [110]. Again,
compounds with more phenolic character such as 139 are more active.
O
OMe
O
138: R = CH3
RO
OMe
O
HO
139: R = H
O
Psoracorylifols A–E (140–144) were isolated from a well-known traditional
Chinese medicine, the seeds of P. corylifolia. The structures of compounds
140–144, including their absolute configurations, were established on the basis
of spectral methods and biogenetic deduction. The structure of 140 was confirmed by single-crystal X-ray diffraction. Psoracorylifols D and E (143 and 144)
represent a novel carbon skeleton. These tested compounds showed significant
inhibitory activity against two strains of H. pylori (SS1 and ATCC 43504) at
the level of MICs of 12.5–25 μg ml−1 . It is remarkable that psoracorylifols A–E
(140–144) are 5–10 times stronger than the standard drug metronidazole,
which is a critical ingredient for combination therapies of H. pylori infection, and
particularly, against H. pylori ATCC 43504. Therefore, such compounds or their
derivatives can be potential substitutes for the present antibacterial drugs [111].
243
244
9 Natural Products as Antimicrobial Agents – an Update
The phenylpropyl benzoates (145–147) isolated from Croton hutchinsonianus
were found to exhibit antifungal activity against C. albicans (IC50 = 11.41, 7.05,
5.36, respectively) [112]. Methyl gallate (148) and protocatechuic acid (149) from
Sebastiania brasiliensis were identified as the growth inhibitors of B. subtilis,
M. luteus, S. aureus, P. aeruginosa, E. coli, C. albicans, Mucor sp., and A. niger,
both with an MIC of 128 μg ml−1 [113].
H
OH
H
H
O
HO
H
H
O
HO
O
HO
HO
O
O
140
H H
H
141
H
HO
O
O
H
O
144
143
142
O
H
H
O
R1
HO
OH
HO
R2
O
OMe HO
OH
HO
OH
OH
148
149
145: R1= R2 = OMe
146: R1 = OMe, R2 = OH
147: R1 = R2 = H
9.10.2
Antimicrobial Phenolics from Microbial Sources
2-Hydroxy-5-(3-methylbut-2-enyl)benzaldehyde (150) and 2-hepta-1,5-dienyl3,6-dihydroxy-5-(3-methylbut-2-enyl) benzaldehyde (151) are constituents of
a marine-derived Streptomyces atrovirens isolated from the rhizosphere of the
marine seaweed Undaria pinnatifida. Both the compounds inhibited the growth
of Gram-positive (Lactococcus garvieae, Streptococcus iniae, and Streptococcus
parauberis) and Gram-negative (Edwardsiella tarda, Vibrio anguillarum, and
V. harveyi) pathogens with MIC values of 15–128 μg ml−1 [114].
2,3-Dihydroxy-5-(hydroxymethyl) benzaldehyde (152), 4-(4-hydroxyphenoxy)
butan-2-one (153), and acetic acid-2-hydroxy-6-(3-oxo-butyl)-phenyl ester (154)
are known as the secondary metabolites of Streptomyces sp. TK-VL_333. In an
antimicrobial assay, it was discovered that the pathogenic Gram-positive bacteria
tested, S. aureus and Staphylococcus epidermidis, were highly sensitive to these
metabolites. However, against a large panel of bacteria, yeast, and filamentous
fungi, all the metabolites exhibited inhibitory activity with MIC values ranging
between 15 and 50 μg ml−1 . It has been reported that most of the benzaldehyde
derivatives having active hydroxyl groups show antimicrobial activity [114]; however, their mode of action has not yet been reported.
9.10
Antimicrobial Phenolics Other Than Flavonoids and Lignans
HO
HO
OH
HO
HO
CHO
CHO
150
CHO
152
151
O
O
O
O
OH
O
154
153
Petit et al. [115] have isolated methyl-6-acetyl-4-methoxy-5,7,8-trihydroxynaphthalene-2-carboxylate (155) from the culture extract of a Penicillium sp.
In an antimicrobial assay, compound 155 exhibited promising activity against
C. albicans with an MIC value of 32 μg ml−1 , and against Listeria monocytogenes and B. cereus with an MIC value of 64 μg ml−1 . Sophisticated NMR
techniques helped characterize the isolates of mushroom Merulius incarnatus
as 5-alkylresorcinols (156–161). Compound 157 is the first 5-alkylresorcinol
derivative that contains a trans–cis conjugated double bond system. Compounds
156–161 were found to inhibit (MRSA) with IC50 values of 2.5, 15, 9.5, 8.0, 5.0,
and 6.5 μg ml−1 , respectively [116].
HO
156
OH
HO
157
OH
OMe OH O
MeO
OH
O
HO
158
OH
OH
155
HO
159
HO
OH
160
OH
HO
161
OH
245
246
9 Natural Products as Antimicrobial Agents – an Update
The culture extract of an endophytic fungus yielded 1-(2-hydroxy-6methoxyphenyl)butan-1-one (162) and 1-(2,6-dihydroxyphenyl)butan-1-one
(163) as the growth inhibitor of the Gram-positive bacterium B. megaterium,
the fungi Microbotryum violaceum, and Septoria tritici at a concentration of
0.25 mg/paper disc [117]. Sorbicillin analogs (164 and 165) are the constituents of
a fungicolous isolate of the genus Phaeoacremonium (NRRL 32148). Compound
164 displayed potent activity in disc diffusion assays against Aspergillus flavus
(40-mm diameter clear zone at 48 h) and moderate activity against F. verticillioides (18-mm zone of partial clearing at 48 h; 200 μg/disk in each case), while
compound 165 showed similar activity against F. verticillioides (18-mm zone of
partial clearing at 48 h), but slightly weaker activity against A. flavus (18-mm
zone of reduced sporulation at 48 h) [118].
CHO
OH
OMe
OH O
OH O
162
CH2OH
OMe
O
163
OMe
OH
O
164
OH
165
9.10.3
Antimicrobial Phenolics from Marine Source
Plaza et al. [119] identified chrysophaentins A (116), D–F (167–169), and H
(170) from the marine chrysophyte alga Chrysophaeum taylori as antimicrobial
agents. These compounds were active against the drug-susceptible bacteria
Staphylococcus aureus (SA) and E. faecium, and drug-resistant strains MRSA and
vancomycin-resistant Enterococcus faecium (VREF). Disc diffusion assays showed
that these compounds inhibited the growth of all tested strains at loads ranging
from 2 to 25 μg/disk. Chrysophaentin A (166) was found to be the most potent
HO
OH
HO
R2
OH
HO
OH
O
OH
Cl
Cl
Cl
Cl
O
OH
OH
Cl
HO
Cl
Cl
HO
HO
HO
R
O
O
R1
Cl
HO
HO
HO
OH
166: R1 = R2 = Cl
167: R1 = R2 = Br
O
Cl
Cl
OH
OH
168
169: R = H
170: R = Br
9.11
Antimicrobial Polypeptides
antibiotic with MIC50 values of 1.8 ± 0.6, 1.5 ± 0.7, and 1.3 ± 0.4 μg ml−1 against
SA, MRSA, and MDRSA, respectively; and 3.8 ± 1.9 and 2.9 ± 0.8 μg ml−1 toward
E. faecium and VREF. Chrysophaentins F (169) and H (170) were the next most
potent compounds with MIC50 values of 4–6 μg ml−1 toward S. aureus and MRSA,
and ∼9.5 μg ml−1 against VREF. Structure–activity analysis revealed that hydroxyl
group and bromine atom are important for antimicrobial action.
9.11
Antimicrobial Polypeptides
Polypeptides are well known for their pharmaceutical uses including antimicrobial
properties. They are well known to be used as antimicrobial agents, for example, in
pharmaceutical applications, as well as for preservation, cleaning, and disinfection
of various surfaces, objects, and substances. The polypeptides may in particular be
used to treat textiles or laundry, for example, in detergents, for reducing microbes
on textile or laundry, and for odor reduction by reducing microbial growth [120].
Mostly, the polypeptides are of microbial origin and several polypeptides are discovered in recent years as new antimicrobial agents.
The culture of Beauveria sp. FKI-1366 yielded beauvericins A (171), D (172), E
(173), and F (174), which were evaluated for their antifungal potential. Although
these compounds were inactive against the tested pathogens, interestingly, they
potentiated miconazole activity against wild C. albicans and fluconazole-resistant
C. albicans. Beauvericins D (172) and E (173) decreased the IC50 value of miconazole against fluconazole-resistant C. albicans from 1.3 μM to 0.25 and 0.31 μM,
respectively [121].
R3
R2
R1
O
R1
R2
R3
171:
CH(CH3)CH2CH3
CH3
CH2C6H5
172:
CH(CH3)2
H
CH2C6H5
173:
CH(CH3)2
H
CH2CH(CH3)2
174:
CH2CH(CH3)2
CH3
CH2C6H5
O
N
O
O
N
O
O
O
O
N
O
GE 23077 factors A1 (175), A2 (176), B1 (177), and B2 (178) are novel
antibiotics isolated from fermentation broths of an Actinomadura sp. strain.
GE23077 antibiotics are cyclic peptides, which as a complex, inhibited E. coli
and B. subtilis RNA polymerase at concentrations of 0.02 μg ml−1 . An E. coli
rifampicin-resistant RNA polymerase was also inhibited at similar concentrations
(IC50 = 0.04 μg ml−1 ). The individual GE23077 factors (175–178) inhibited E. coli
RNA polymerase, showing IC50 values of 0.15, 0.035, 0.1, and 0.02 μg ml−1 ,
respectively, whereas E. coli DNA polymerase was not inhibited even at high
antibiotic concentrations (IC50 > 1000 μg ml−1 ) [122].
247
248
9 Natural Products as Antimicrobial Agents – an Update
OH
H
N
O
OH
H2N
HN
O
HO
HN
O
OH
O
O
O
175, 176: R =
HN
O
H
N
N
H
HO
OH
HN
O
O
177, 178: R =
O
O HN
R
Tripropeptins (TPPs) A–E and Z (179–183 and 184) were purified from
culture of Lysobacter sp; All the isolates exhibited potent antibacterial activities
(MIC of 0.39–12.5 μg ml−1 ) against a wide range of pathogens. Particularly, TPP
C (181) and TPP D (182) showed excellent activities against Gram-positive
bacteria including both MRSA and vancomycin-resistant Enterococcus (VRE)
[123]. The antimicrobial activities of these TPPs is correlated with the length of
fatty acyl side chains, indicating that the longer the side chain the more active
the TPP. The antimicrobial activities of TPPC (181), TPPD (182), and TPPE
(183) against clinically isolated S. aureus strains, both methicillin-susceptible
Staphylococcus aureus (MSSA) and MRSA, were favorable compared with the
currently available antimicrobial agents such as vancomycin, meropenem, and
so on. Although antimicrobial activity of TPPs (179–184) is correlated with
the length of the fatty acyl side chain, it was proved that the acyl side chain
plays an important role in antimicrobial activities not only of TPPs (179–184)
but also of other lipopeptides, such as echinocandin micafungin group [124],
daptomycin group (LY146032 and A21978Cs) [125], polymyxins, and octapeptins
[126]. Comparison of antimicrobial activities between TPPs (179–184) and the
TPP-like lipopeptides [127, 128] suggested that the C-13 fatty acyl chain seems
to be the best one for antimicrobial activity [129].
OH
O
NH
H2N
N
H
N
NH H
O
O
H
180: R = (CH2)8CH(CH3)CH3
NH
O
R
O
NH
O
HO
HO2C
179: R = (CH2)7CH(CH3)CH3
N
H
N
O
HO
O
O
H
N
O
182: R = (CH2)10CH(CH3)CH3
OH
N
H
OH
181: R = (CH2)9CH(CH3)CH3
CO2H
183: R = (CH2)11CH(CH3)CH3
184: R = (CH2)6CH(CH3)CH3
Hassallidin B (185), a member of a broad-spectrum antifungal class of compounds, was isolated from cyanobacterium Hassallia sp. [130]. Hassallidin A
(186) [131] is the first identified member and compound 185 is the second
member of this class. Antifungal tests against 10 species of Candida showed that
the MICs of 185 were similar to those of 186, ranging from 8.0 to 16.0 μg ml−1 ,
and the given data in the literature suggested that 185 does not result in a different
9.12
Antimicrobial Polyketides
mode of action compared to hassallidin A (186). The additional carbohydrate
moiety does not play a decisive role in the antifungal mode of action. However,
owing to the additional hydrophilic unit, hassallidin B (185) showed better water
solubility. Within the drug design feature, the biosynthetic modification could
be important for improved bioavailability through enhanced water solubility,
keeping the broad spectrum of antifungal activity. So, hassallidin B (185) can be
regarded as an excellent source for the development of new antifungal drugs.
OH
HO
HO
H
N
HN
O
O
O
N
H
O
O
H2N
O
HO
H
N
N
O
OH
O
HO
O
N
H
O
O
O
NH
OH
O
NH2
HN
O
O
OH
O
HO HO O
HO HO
HO
H
N
185
H
N
O
O
HO
H
N
O
N
H
O
O
O
O
N
H
O
NH
N
H
O
N
OO
CH2OH
O
OH
NH2
HN
O
H2N
OH
O
O
186
OH HO
HO
9.12
Antimicrobial Polyketides
Polyketides are a group of secondary metabolites, exhibiting remarkable diversity
in terms of both their structure and function. Polyketide natural products are
known to possess a wealth of pharmacologically important activities, including
antimicrobial properties. A wide range of biological activities associated with
polyketides makes them economically, clinically, and industrially the most sought
after molecules. The importance of this class of natural products is also reflected
by the presence of several molecules of this class in current pharmaceutical
market. For example, erythromycin A, a broad-spectrum antibiotic, rapamycin,
rifamycin, lovastatin, and many oxytetracyclines are a few more of the thousands
of polyketides discovered so far [132].
249
250
9 Natural Products as Antimicrobial Agents – an Update
OH
O
O
OMe
NMe2
O
OH HO
HO
O
OH
O
O
O
O
O
N
O
HO
O
OH
OMe
Erythromycin A (antibacterial)
O
O
O
MeO
O
OH OH
O
OMe
O
OH OH
OH
O
NH
O
O
Rapamycin (immunosuppressant)
O
O
O
Rifamycin B (antituberculosis)
O
HO
O
O
O
H
Lovastatin (anticholesterol)
Owing to the biosynthetic modes in different organisms, polyketides are divided
into several subclasses, of which the most common are macrolides, quninones,
and tetracylines.
9.12.1
Antimicrobial Polyketides as Macrolides
Since the discovery of erythromycin, macrolides have been an important class
of antibiotics. The term macrolide is used to describe drugs with a macrocyclic
lactone ring of 12 or more carbons [133]. This class of natural products includes
a variety of bioactive agents, including antibiotics, antifungal drugs, prokinetics,
and immunosuppressants. Owing to their excellent tissue penetration ability,
14-, 15-, and 16-membered macrolides are a widely used family of antibiotics.
This class of natural antibiotics has proved effective against Gram-positive cocci
and atypical pathogens [134].
Macrolides are best known as the bacterial constituents as a number of these
compounds have been identified from bacterial sources and evaluated for their
activities [135]. Two 10-membered macrolides, phomolides A (187) and B
(188), were purified from the culture of Phomopsis sp. hzla01-1. Their structures were identified with the help of 1D and 2D NMR spectroscopic analyses.
Compounds 187 and 188 showed significant antimicrobial activities against
E. coli CMCC44103, C. albicans AS2.538, and S. cerevisiae ATCC9763 with
MICs of 5–10 μg ml−1 . On the other hand, these compounds showed significant
cytotoxicity against the HeLa cell line at 10 μg ml−1 [136]. A relatively larger
macrolide, the 24-membered ring lactone macrolactin N (189), was isolated
9.12
Antimicrobial Polyketides
251
from a culture broth of B. subtilis. Macrolactin N (189) inhibited S. aureus
peptide deformylase with an IC50 value of 7.5 μM [137]. Another cytotoxic
tetraene macrolide CE-108 (190), produced by Streptomyces diastaticus 108,
was tested for its antimicrobial potential. The macrolide tetraene CE-108 (190)
exhibited potent antimicrobial properties against C. albicans (MIC = 32 μg ml−1 ),
Candida krusei (MIC = 32 μg ml−1 ), C. neoformans (MIC = 16 μg ml−1 ), A. niger
(MIC = 8 μg ml−1 ), Aspergillus candidus (MIC = 8 μg ml−1 ), F. oxysporum
(MIC = 8 μg ml−1 ), Microsporum canis (MIC = 16 μg ml−1 ), Microsporum gypseum (MIC = 8 μg ml−1 ), T. mentagrophytes (MIC = 8 μg ml−1 ), T. rubrum
(MIC = 16 μg ml−1 ), and Trichophyton tonsurans (MIC = 16 μg ml−1 ). This compound can be a good candidate to develop a broad-spectrum antifungal drug
[138].
OH
O
O
HO
O
O
O
O
OH
HO
O
HO
O
O
187
188
O
O
OH
O
O
O
OH
NH2
OH
189
Novel tartrolons 191 and 192 produced by symbiotic cellulose-degrading
bacteria in shipworm gills inhibited B. subtilis in disc diffusion assays. The novel
boronated tartrolons 192 exhibited significant antibacterial activity against
P. aeruginosa (MIC = 0.31 μg ml−1 ) in addition to methicillin-sensitive and MRSA
(MIC = 0.08–1.0 μg ml−1 ). Compound 192 also killed the marine organism V.
anguillarum. The MIC value of compound 192 against B. subtilis was determined
to be 1.0 μg ml−1 [139].
OH O
O
O
OH
O
OH
HO
HO
O
O
OH O
O
O
O
O
191
O
O
B
O
Na O
+
O
O
O
OH
O
OH
192
O
OH O
190
252
9 Natural Products as Antimicrobial Agents – an Update
9.12.2
Antimicrobial Polyketides as Quinones and Xanthones
The natural sources containing quinones as their chemical constituents afford
many biological activities including antimicrobial properties. A survey of the
literature revealed that natural quinones and xanthones are well known for
their bactericidal properties [140, 141] and are discovered from plant as well
as microbial sources. Xanthones are a unique class of phytochemicals and are
related to phytoalexins bearing MRSA inhibitory properties. Their biological
evaluation revealed that these compounds have less side effects and low tendency
to acquire resistance compared with conventional antibiotics and synthetic
antibacterial agents [142].
9.12.2.1 Antimicrobial Quinones and Xanthones from Plants
Chromatographic purification of a methanolic extract from the root bark of
Newbouldia laevis yielded chrysoeriol (193), which offered broad-spectrum
in vitro antimicrobial activity against 4 Gram-positive and 10 Gram-negative
bacterial species as well as 3 Candida species, and the observed MIC values varied from 1.2 to 9.76 μg ml−1 [143]. The same plant also contains
2-acetylfuro-1,4-naphthoquinone (194), 2-hydroxy-3-methoxy-9,10-dioxo-9,10dihydroanthracene-1-carbaldehyde (195), and lapachol (196). Compounds
194–196 also exhibited broad-spectrum in vitro antimicrobial activity against
6 Gram-positive and 12 Gram-negative bacterial species as well as Candida
strains. The MIC values reported varied from 0.076 to 9.76 μg ml−1 [143]. All
these compounds were purified by chromatographic techniques and were
characterized by NMR methods and after further studies can be good leads for
antimicrobial drug development.
O
O
HO
O
OMe
O
O
O
O
O
O
193
194
OH
OH
O
195
CHO
O
196
Kuete et al. [82] identified antimicrobial quinones (197–201) from the extract
of Vismia laurentii De Wild. Quinones 197–201 are active against Gramnegative bacteria and two Candida species with different efficacy. Compound
197 inhibits the growth of S. faecalis and M. morganii (MIC = 4.8 μg ml−1 ),
and P. aeruginosa and S. flexneri (MIC of 2.4 = μg ml−1 ); 198 showed an MIC
of 4.8 μg ml−1 against M. morganii and S. faecalis, and 2.4 μg ml−1 against P.
aeruginosa, S. flexneri, B. subtilis, C. albicans, and C. glabrata; 199 was only
active against Gram-positive and fungal pathogens with an MIC of 2.4 μg ml−1
9.12
Antimicrobial Polyketides
(B. megaterium, B. subtilis, and C. albicans, and C. glabrata). Compound 200
exhibited an MIC of 4.8 μg ml−1 against P. aeruginosa, S. flexneri, and B. subtilis
and 9.7 μg ml−1 against M. morganii, whereas compound 201 was highly active
against B. subtilis with an MIC value of 1.2 μg ml−1 . It further showed a significant
growth inhibition (MIC = 4.8 μg ml−1 ) of Shigella dysenteriae, S. faecalis, and B.
stearothermophilus, while exhibiting an MIC of 2.4 μg ml−1 against B. cereus.
The same compound extended moderate activity (MIC = 19.5 μg ml−1 ) against
M. morganii, P. vulgaris, and S. aureus.
OH O
OH
OH O
OMe
OMe
O
O
198
197
OH O
OH O
OH
OH
O
O
199
OH
OH O
OH
OMe
OMe
OH O
200
OMe
O
OH
201
Omicron-naphthoquinone, 9-methyl-3-(4-methyl-3-pentenyl)-2,3-dihydronaphtho[1,8-bc]pyran-7,8-dione (202) is a constituent of Australian plant
Eremophila serrulata and was isolated by silica gel column chromatography and
characterized by 1 and 2D NMR techniques coupled with MS. Compound 202
has been reported as an antibacterial natural product that exhibited antimicrobial
activity against S. aureus ATCC 25923 and Streptococcus pneumoniae ATCC
49619, and S. pyogenes ATCC 10389 with an MIC of 7.8 μg ml−1 [144].
O
O
O
H
202
Diospyros lycioides, a plant commonly known as muthala, yielded bioactive
quinone glycosides, diospyrosides A (203), B (204), C (205), and D (206) along
with juglone (207) and 7-methyljuglone (208) and were evaluated as growth
inhibitors of oral cariogenic bacteria (Streptococcus mutans and Streptococcus
253
254
9 Natural Products as Antimicrobial Agents – an Update
sanguis) and periodontal pathogens (Porphyromonas gingivalis and Prevotella
intermedia) at an MIC ranging from 0.019 to 1.25 mg ml−1 . Juglone (207)
exhibited the strongest inhibitory activity (0.019–0.078 mg ml−1 ) among these
compounds. This is in accordance with the fact that tannin-like compounds
afford antibacterial activities [145].
O
OH
OH
O
O
OH
OH
OH
O
β
Glc
β
6
Xyl
β
4
Xyl
203
O
O
β
Glc
β
6
Xyl
204
O
OH
OH
O
β
Glc
β
6
Xyl
205
O
O
O
HO
Xyl
β
6
Glc
β
O
OH
O
OH
O
OH
206
207
208
3,4-Dihydroxy-1-methoxy-anthraquinone-2-carboxaldehyde (209) and damnacanthal (210) were identified from the aerial parts of Saprosma fragrans
as antifungal compounds with MIC values of 12.5 μg ml−1 against T. mentagrophytes and 25 μg ml−1 against Sporothrix schenckii, 1.56 μg ml−1 against
T. mentagrophytes and 6.25 μg ml−1 against S. schenckii, respectively [146].
The study of the chemical constituents of the roots of N. laevis (Bignoniaceae)
resulted in the isolation and characterization of a naphthoquinone–anthraquinone
coupled pigment named newbouldiaquinone A (211). Its antimicrobial activity
against a wide range of microorganisms was 13- and 24-fold more potent than the
standard antibiotics nystatin and gentamycin against C. glabrata and E. aerogenes.
It exhibited very pronounced activities against Gram-negative bacteria with the
MIC ranging between 0.31 and 9.76 μg ml−1 . Its antifungal activity was also
important, with C. glabrata being the most sensitive yeast [147].
The above-discussed results revealed that both kinds of quinones exhibit
antimicrobial properties with variable efficacy. The activity difference may lie in
9.12
Antimicrobial Polyketides
the substituent present at various positions on the benzene ring and may also lie
in the solubility difference.
O
O
O
OMe O
OMe O
O
OH
OH
O
O
H
H
O
O
OH
209
HO
211
210
O
O
O
HO
O
212
7-Epiclusianone (212), identified from the antimicrobial extract of Rheedia
brasiliensis fruit, showed high antibacterial activity at low concentrations (MIC,
1.25–2.5 μg ml−1 ) against S. mutans. This potential may lead to the use of 212 as
a new agent to control S. mutans biofilms; however, more studies are needed to
further elucidate the mechanisms of action [148].
Tetraoxygenated xanthones, mangostanin (213) and α-mangostin (214), were
isolated from the fruits of Garcinia cowa. Compound 213 inhibited the growth of
the penicillin-sensitive strain ATCC 25923 and methicillin-resistant strain MRSA
SK1 S. aureus with MIC values of 4.0 μg ml−1 . Compound 214 was only moderately active against both strains of S. aureus with an MIC value of 8.0 μg ml−1 [149].
O
O
OH
HO
O
213
OH
MeO
MeO
O
HO
O
214
OH
255
256
9 Natural Products as Antimicrobial Agents – an Update
O
O
OH
OH
O
HO
O
O
OH
OH
OH
215
216
O
O
OH
OH
217: R =
O
HO
HO
O
O
OH
OH
OH
218: R =
R
219
Boonsri et al. [150] isolated formoxanthone A (215), formoxanthone C (216),
macluraxanthone (217), xanthone V1 (218), and gerontoxanthone I (219)
as potential antibiotics from the roots of Cratoxylum formosum. Compound
215 inhibited the growth of B. substilis (MIC = 18.7 μg ml−1 ) and S. aureus
(MIC = 37.5 μg ml−1 ), whereas other compounds were more active against B. substilis (MIC = 4.6 μg ml−1 ), S. aureus (MIC = 2.3 and 4.6 μg ml−1 , respectively),
S. faecalis (MIC = 18.7 and 2.3 μg ml−1 , respectively), and S. typhi (MIC = 4.6
and 9.3 μg ml−1 , respectively). However, all these compounds were also found
to be cytotoxic; therefore, further studies are needed before considering these
metabolites as drug candidates.
Xanthones 220 and 221 have been reported from the extract of V. laurentii De
Wild. Compound 220 showed potential inhibition of B. subtilis and C. glabrata
with an MIC value of 1.2 μg ml−1 , whereas 221 was highly active against B.
stearothermophilus and B. subtilis (MIC of 4.8 μg ml−1 ) [82].
O
O
OH
O
OH
O
O
O
OH
OH
HO
OH
220
221
9.12.2.2 Antimicrobial Quinones from Bacteria
Screening of microbial extracts using antisense-sensitized rpsD S. aureus strain
led to the isolation of pleosporone (222), with modest antibacterial activities
exhibiting an MIC ranging from 1.0 to 64.0 μg ml−1 . This compound showed
the highest sensitivity for S. pneumoniae and Haemophilus influenzae, and
exhibited MICs of 4.0 and 1.0 μg ml−1 , respectively. Pleosporone (222) showed
greater inhibition of S. aureus RNA synthesis (IC50 = 1.3 μg ml−1 ) over DNA
(IC50 = 8.4 μg ml−1 ) and protein synthesis (IC50 = 15.4 μg ml−1 ). These results
9.12
Antimicrobial Polyketides
257
suggest that this compound has another unknown mode of action in addition to
the weak interaction with RPSD protein [151].
The antifungal substances produced by Streptomyces strain AcH 505 were
identified as the antibiotics WS-5995 B (223) and C (224). WS-5995 B (223)
completely blocked mycelial growth at a concentration of 60 μM and caused a cell
stress-related gene expression response in Amanita muscaria. Characterization
of these compounds provides the foundation for molecular analysis of the
fungus–bacterium interaction in the ectomycorrhizal symbiosis between fly
agaric and spruce [152].
The bacterial metabolites bisanthraquinone (225 and 225), isolated from
streptomycete sp. associated with the tropical tunicate Ecteinascidia turbinata,
potently inhibited the growth of MRSA (IC50 = 0.15 μM), but were >10-fold less
effective against VREF (VRE, IC50 = 2.0 μM) [153] The structures of these compounds were elucidated by the combination of various spectroscopic techniques.
Anthracycline antibiotics mutactimycin PR (226) and C (227) were isolated from
a soil isolate Saccharothrix sp. These compounds showed an antibiotic activity
against certain Gram-positive bacteria in vitro [154].
OH O
OH
O
OH
OH
OH
H3CO
O
OH O
O O
OH
OH
OH
R
CO2H
HO
OR O
OH O
O
O
OH
HO
OMe
OH O
O
O
222
R
222: R = H
224: R = H
223: R = OH
225: R = OH
O
226: R =
HO
OH
OH
227: R = CH3
9.12.2.3 Antimicrobial Quinones and Xanthones from Fungi
Polyketides 228–233 were isolated from the culture extract of a marine-derived
fungus Nigrospora sp. MA75. These compounds were evaluated for their antibacterial activity and compound 233 exhibited the growth of MRSA, E. coli, P.
aeruginosa, Pseudomonas fluorescens, and S. epidermidis with MIC values of
8.0, 4.0, 4.0, 0.5, and 0.5 μg ml−1 , respectively. The compound was rather more
active than the standard antibiotic ampicillin. Compound 232 is also reported as
a broad-spectrum antibiotic that strongly inhibited MRSA, E. coli, P. aeruginosa,
and P. fluorescens with MIC values of 0.5, 2, 0.5, and 0.5 μg ml−1 , respectively.
Compound 229 showed significant inhibition of MRSA and E. coli (MIC = 2.0
and 0.5 μg ml−1 , respectively), while its analog, 230, only showed activity against
E. coli with an MIC value of 4.0 μg ml−1 . Structure–activity relationship analysis
revealed that the OH group at C(4) might be important for the activity against
258
9 Natural Products as Antimicrobial Agents – an Update
MRSA [155]. Compound 228 showed moderate activity against V. mali and
Stemphylium solani, both with MIC values of 16 μg ml−1 . Griseofulvin (228) is an
antifungal antibiotic and is used for the treatment of mycotic diseases of human,
OH
OMe O OMe
MeO
OH
H
OH O
OH
O
O
MeO
OH O
Cl
H
OH
229: R = OH
230: R = H
228
231
HO
OH
O
OH
O
HO
H
MeO
OH OMe
OH
O
HO
R
O
H
O
OH
O
232
233
veterinary, and plant systems [156]. However, other members of the griseofulvin
family have been reported to exhibit much weaker or no antifungal activities [157]
suggesting that the activity of 228 is related to both its planar structure and spatial
configuration.
The endophytic fungus Microdiplodia sp. produced several xanthone derivatives (234–239). All these compounds possess weak antibacterial activity against
Legionella pneumophila Corby, E. coli K12, and B. megaterium. In addition, compounds 235, 237, and 238 also exhibited antifungal activity against the fungus
M. violaceum [158].
OH O
R
OH
OH
O
OH
234: R = Me
235: R = CH2OH
OH
O
OH
H
HO
O
O
CO2Me
OMe
237
OH O
OH
O
OH
238
OH
OH
O
O
236
OH O
O
O
239
O
9.12
Antimicrobial Polyketides
An unidentified endophytic fungus of the Pleosporales group isolated from
Anthillis vulneraria L. (Fabaceae) produced pleosporone (240) and phaeosphenone (241). The compounds were isolated by silica gel and Sephadex LH-20
chromatography followed by reversed-phase HPLC, and the structure was
determined by NMR and MS means. In a two-plate whole-cell differential sensitivity screening assay using an antisense-sensitized S. aureus strain [159, 160],
240 exhibited significant sensitivity for S. pneumoniae and H. influenzae, and
exhibited MICs of 4.0 and 1.0 μg ml−1 , respectively, and 8.0 and 16 μg ml−1 for
B. subtilis and E. coli [151]. Compound 241 showed broad-spectrum antibacterial
activity against Gram-positive bacteria, exhibiting MIC values ranging from 8 to
64 μg ml−1 . It showed the highest sensitivity for S. pneumoniae (MIC = 8 μg ml−1 )
and inhibited the growth of C. albicans with an MIC of 8.0 μg ml−1 , whereas 241
showed a modest selectivity for the inhibition of RNA synthesis over DNA and
protein synthesis in S. aureus [161]. Ribosomal protein S4 (RPSD), a part of the
ribosomal small subunit, is one of the proteins that is a part of the ribosomal
machinery and is a potential new target for the discovery of antibacterial agents.
Therefore, these compounds can also be a good lead to future antibacterial drug
development.
Another endophytic fungus Chloridium sp. produces a highly functionalized
naphthoquinone, javanicin (242). This compound exhibits strong antibacterial
activity against P. fluorescens, P. aeruginosa (MIC = 2.0 μg ml−1 ), R. solani,
and Verticillium dahliae (MIC = 10 μg ml−1 ), and antifungal activity against
Cercospora arachidicola (MIC = 5.0 μg ml−1 ) [162]. Li et al. [163] isolated a
polyketide, asperflavin ribofuranoside (243), from the marine-derived fungus
Microsporum. In an antibacterial evaluation assay, the isolate showed moderate
activity against the MRSA and MDRSA with an MIC value of 50 μg ml−1 .
Decaspirones A–E (244–248), related to the palmarumycins class of compounds, were isolated from cultures of the freshwater aquatic fungal species
Decaisnella thyridioides. In a standard antibacterial disc assay, all the compounds
244–248 showed significant activity against B. subtilis (ATCC 6051) when
tested at 50 μg/disc, showing inhibition zones of 39, 19, 34, 30, and 30 mm,
respectively. Compounds 244, 246, and 247 were active against S. aureus
(ATCC 29213), causing zones of inhibition of 41, 28, and 30 mm, respectively,
at 100 μg/disc. Compounds 244, 247, and 248 also showed activity against
C. albicans (ATCC 14053) at 100 μg/disk, affording inhibition zones of 30, 13, and
17 mm, respectively, while compounds 245 and 246 were inactive in this assay.
Compounds 244 and 246 were evaluated for activity against A. flavus (NRRL
6541) and F. verticillioides (NRRL 25457). Assays at 200 μg/disk indicated that
both of them have significant activity against A. flavus (inhibition zones of 20
and 15 mm, respectively) and F. verticillioides (inhibition zones of 20 and 26 mm,
respectively). Upon further evaluation, compound 244 displayed MIC values of
approximately 10 and 5 μg ml−1 against A. flavus and F. verticillioides, respectively
[164].
259
260
9 Natural Products as Antimicrobial Agents – an Update
OH O
OH O
OH
H
OH
OH
HO
O
OH
Me
O
O
241
O
OMe OH O
H
OH
243: R = -D=ribofuranose
O
O
O
OH
O
242
HO
OR1
H
RO
OH
MeO
OH
O
240
O
OH
O
OR2
244: R1 = H, R2 = H
245: R1 = H, R2 = OAc
246: R1 = OAc, R2 = H
OH OAc
H
O
H
H
O
247
OH
O
H
O
OH
OH
248
Chrysoqueen (249) and chrysolandol (250) were identified as metabolites from
the culture broth of Chrysosporium queenslandicum. Both compounds are new
members of the naphthoquinone-type altersolanol family of antibiotics, which are
also known as products of fungi. Compounds 249 and 250 showed selective activity against Gram-positive bacteria such as M. luteus IFM 2066 (MIC = 33 μg ml−1 )
and B. subtilis bacterial strain passport (PCI) 219 (MIC = 33 μg ml−1 ) [165].
HO
OH
HO O
OMe
HO
O
O
OH
O
O
249
HO
OH
OH
HO
OH
O
OH
OMe
250
Protein synthesis is one of the best antibacterial targets that has led to the
development of a number of highly successful clinical drugs. In addition to
9.12
Antimicrobial Polyketides
providing a source of stable free radicals, quinones are known to complex
irreversibly with nucleophilic amino acids in proteins [166], often leading to
inactivation of the protein and loss of function. For that reason, the potential
range of quinone antimicrobial effects is great. Probable targets in the microbial
cell are surface-exposed adhesins, cell wall polypeptides, and membrane-bound
enzymes. Quinones may also render substrates unavailable to the microorganism. The above-discussed results revealed that quinones exhibit antimicrobial
properties with variable efficacy. The difference may lie in the substituent
present at various positions on the benzene ring and also in the solubility
differences.
9.12.3
Antimicrobial Fatty Acids and Other polyketides
Fatty acids are biosynthetically derived through the polyketide biosynthetic
pathway and several are known to possess antimicrobial activity. A survey of the
literature reveals that unsaturated and hydroxy fatty acids show better antibacterial activities [167].
Marine Bacillus sp., culture yielded unsaturated fatty acids or esters:
ieodomycins A–D (251–254), which exhibited antimicrobial activities against
B. subtilis (KCTC 1021), E. coli (KCTC 1923), and S. cerevisiae (KCTC 7913)
OH
OH OH O
OMe
O
251
252
OH
OH
O
OH
O
OH
OH
253
O
254
with MIC values of 32–64 μg ml−1 , except against the yeast S. cerevisiae,
with an MIC of 256 μg ml−1 [168]. Despite extensive study on the mode of
action of these compounds, the precise mechanism for the antimicrobial activity remains unclear; however, it is suggested that the antimicrobial activity
of unsaturated fatty acids is related to the inhibition of bacterial fatty acid
synthesis [169].
261
262
9 Natural Products as Antimicrobial Agents – an Update
Bicyclic lactones, glabramycins A–C (255–257), were discovered from
a Neosartorya glabra strain using the antisense technique. Compound 257
inhibited the growth of S. aureus with an MIC value of 16 μg ml−1 and exhibited a similar inhibitory activity against B. subtilis, but showed low activity
against E. faecalis (MIC = 432 μg ml−1 ). This compound exhibited potent activity
against S. pneumoniae with an MIC value of 2.0 μg ml−1 . Glabramycins A (255)
and B (256) showed weak activity against these tested organisms. All these
compounds were identified as selective inhibitors as they were inactive against
Gram-negative bacteria or fungi C. albicans. A study on the mode of action of
these compounds revealed that these compounds may inhibit the synthesis of
RNA [170].
O
HO
OH
O
O
O
O
H
O
O
O
256
OH
H
O
O
255
OH
H
O
OH
O
O
H
257
Using the antisense technique, Singh et al. [171] discovered novel polyketidederived molecules lucensimycins A–G (258–263) from the culture extracts
of Streptomyces lucensis. In an rpsD antisense two-plate assay, compound 260
inhibited the growth of S. aureus exhibiting a zone of clearance of 12 mm
at a concentration of 0.5 μg ml−1 . Compound 258 and 261 also showed
similar activity. All of the compounds were also tested for antibacterial
activity in standard National Laboratory Standard Institute (NLSI) protocol. Lucensimycin E (261) potently inhibited the growth of S. aureus Smith
strain with an MIC value of 32 μg ml−1 . Further, it showed better sensitivity for S. pneumoniae CL 2883 and killed the pathogen with an MIC value
of 8.0 μg ml−1 . Lucensimycin D (260) did not inhibit S. aureus growth at
250 μg ml−1 in this assay. In addition, the investigators tested the activity
of these compounds at higher concentrations on agar plates impregnated
with S. aureus Ep167. In this assay, compounds 258, 262, and 263 showed
better activities than 260 and 261, whereas compound 259 was the least
active.
9.13
Antimicrobial Steroids
263
O
HO
O
O
O
O
H
AcO
O
H
O
O
O
H
OH
H
AcO
O
OH
H
258
259
H
OH
H
OH
HN
O
O
HO
O
O
O
O
O
O
H
AcO
O
O
O
H
O
OH
H
AcO
260
OH
H
OH
OH
HO
HO
OH
O
N
H
AcO
O
H
N
HN
OH
H
O
O
O
H
O
O
OH
H
AcO
H
O
O
H
S
O
OH
OH
H
N
O
OH
S
OH
O
O
OH
OH O
O
O
261
HO
HO
O
H
OH
H
OH
HO
HO
S
H
H
S
N
H
O
O
O
OH
H
263
H
OH
262
H
OH
9.13
Antimicrobial Steroids
Steroids are an important class of secondary metabolites, exhibiting diverse
biological activities. In addition to their role as hormones, they also showed
other pharmaceutical functions including antimicrobial properties. For example,
diethylstilbestrol exhibited a degree of bactericidal action on Gram-positive bacteria. Reiss [172, 173] reported the inhibitory effects of methyl testosterone and
deoxycorticosterone on the growth of Trichophyton purpureum and Trichophyton
264
9 Natural Products as Antimicrobial Agents – an Update
gypseum in cultures and a certain curative effect on experimental infection in
castrated rabbits.
9.13.1
Antimicrobial Steroids from Plants
The antibacterial 5′ ,7′ -dimethyl-6′ -hydroxy-3′ -phenyl-3α-amineβ-yne sitosterol
(264) has been reported as a constituent of leaves of Datura metel Linn. Compound 264 weakly inhibited the growth of S. aureus, P. aeruginosa, P. mirabilis, S.
typhi, B. subtilis, and K. pneumonia but could not inhibit E. coli with MIC values
ranging between 12.5 and 25 mg ml−1 [174].
H
N
HO
H
O
264
Aginoside (265) and (25R)-5α-spirostan-3β,6β-diol-3-O-{β-D-glucopyranosyl(1 → 2)-O-[β-D-xylopyranosyl-(1 → 3)]-O-β-D-glucopyranosyl-(1 → 4)- β-D-galactopyranoside} (266) isolated from the flowers of Allium leucanthum exhibited
antifungal activity against seven Candida strains with an MFC ranging from
6.25 to 12.5 μg ml−1 [175]. However, further studies are required to identify
this compound as a potent anti-Candida drug. Another steroidal saponin
CAY-1 (267) from Capsicum frutescens showed antifungal activity and is
reported to be active against 16 different fungal strains, including Candida
spp. and A. fumigatus with the MIC ranging from 4.0 to 16 μg ml−1 , and was
especially active against C. neoformans exhibiting 90% inhibition at 1.0 μg ml−1 .
No significant cytotoxicity was demonstrated when 267 was tested against 55
mammalian cell lines at up to 100 μg ml−1 . Importantly, CAY-1 (267) appears
to act by disrupting the membrane integrity of fungal cells [176]. This nontoxic
compound can be an important source for the development of antimicrobial
agents.
A common phytochemical β-sitosterol-3-O-β-D-glucopyranoside (268) affords
broad-spectrum in vitro antimicrobial activity against three Gram-positive and
nine Gram-negative bacterial species as well as three Candida species. The MIC
values observed varied from 0.61 to 9.76 μg ml−1 [143]. The mechanism of action
of steroids is not much clear; however, it is suggested that their antimicrobial properties might be due to their ability to disrupt membrane integrity.
9.13
OH
O
O
OH
O
R2O
HO
O
O
O
HO
H
R1
H
O
H
OH
OH
H
R1
H
265: OH
β-D-Xyl
266: H
β-D-Xyl
H
OH
HO
HO
H
OH
OH
HO
O
HO
HO
OH
O
H
H
OH
O
HO
O
OH
OH
OH
OH
O
H
O
O
O
O
O
O
H
HO
O
R2
O
HO
HO
265
OH
HO
HO
Antimicrobial Steroids
OH
267
HO
HO
O
OH
O
268
Similarly to other steroidal saponins, tigogenin-3-O-β-D-xylopyranosyl (1 → 2)[β-D-xylopyranosyl (1 → 3)]-β-D-glucopyranosyl (1 → 4)-[α-L-rhamnopyranosyl
(1 → 2)]-β-D-galactopyranoside (TTS-12, 269) and tigogenin-3-O-β-D-glucopyranosyl (1 → 2)-[β-D-xylopyranosyl (1 → 3)]-β-D-glucopyranosyl (1 → 4)-β-Dgalactopyranoside (TTS-15, 270) of Tribulus terrestris L. also showed in vivo
activity against C. albicans with MIC80 of 10 and 2.3 μg ml−1 respectively. Both
the compounds also inhibited C. neoformans with an MIC80 of 1.7 and 6.7 μg ml−1
respectively. Phase contrast microscopy and transmission electron microscopy
revealed that these compounds kill the fungi by weakening the virulence of C.
albicans and by destroying the cell membrane [177].
Other steroidal saponins atropurosides B (271) and F (272), and dioscin (273)
from the rhizomes of Smilacina atropurpurea were found to be antimicrobial with
fungicidal activity against C. albicans, C. glabrata, C. neoformans, and A. fumigatus with MFCs ≤20 μg ml−1 , while dioscin (273) was selectively active against C.
albicans and C. glabrata (MFC ≤ 5.0 μg ml−1 ). Unlike other saponins, these compounds were also found to be cytotoxic, and it appears that the antifungal activity
of these steroidal saponins correlates with their cytotoxicity against mammalian
cells [178].
Some low polar sterols stigmast-5-ene-3-β-7-α-diol (274) and stigmast-5-ene3-β-7α, 20-𝛜-triol (275), which are also common steroids and are isolated from
various plant sources, exhibited moderate antibacterial activities against E. coli
(zone of inhibition 14.1 and 14.0 mm, respectively), S. aureus (zone of inhibition
266
9 Natural Products as Antimicrobial Agents – an Update
13.20 and 13.24 mm, respectively), and P. aeruginosa (zone of inhibition 13.4 and
14.0 mm, respectively) [179].
O
O
O
OR
O
H
HO
H
R1
H
HO
Glu
HO O
271 = R Xyl(2-1)Rha, R1 = OH
RO
^
2
^
269: R = A D gal 4 A D glu
A D xyl
Gal
O
3
2
^
A L rha
^
A D xyl
O
H
HO
H
H
H
HO
^
4
270: R = A D gal
^
2
A D glu
A D glu
272: R = Gal, R1 = H
3
^
A D xyl
R2
O
H
O
H
H
H
H
HO
HO
274: R1 =
H
R1
OH
R2 = H
OH
273:R Glc[(2-1)Rha](4-1)Rha
275: R1 =
R2 = OH
9.13.2
Steroids from Fungi
Similarly to plants, a variety of fungi also produce steroids, but very few have been
reported to possess antimicrobial activities. Ergokonin A (276) is a metabolite
of an endophytic fungus from Trichoderma longibrachiatum, which has been
reported to possess broad-spectrum antifungal activity against many Candida
sp., S. cerevisiae and most of the filamentous fungi and the activity was comparable with that of amphotericin B [180]. It is notable that the above discussed
steroidal saponins and compound 276 are active against Candida species,
which may offer a wide area of research to the pharmacologists to study these
compounds.
9.14
Antimicrobial Terpenoids
HO2C
HO
O
O
S
O
O
H
OH
O
NH2
O
276
9.14
Antimicrobial Terpenoids
Terpenoids are a class of secondary metabolites made of isoprene units, and essential oils are composed of a mixture of mono and sesquiterpenoids, in addition to
other low molecular weight compounds. More than 60% of essential oil derivatives
have been reported as inhibitors of fungi and bacteria [181]. These facts reveal the
antimicrobial potential of this important class of compounds.
9.14.1
Antimicrobial Terpenoids from Plants
Paeonia suffruticosa root bark furnished three monoterpene glycosides, 6-Oanillyoxypaeoniflorin (277), mudanpioside-H (278), and galloyl-oxypaeoniflorin
(279), isolated from the root bark of P. suffruticosa showed moderate antibacterial
activity with MIC values in the range of 100–500 μg ml−1 against 18 human and
animal pathogens [182].
A sesquiterpene named xanthorrhizol (280), isolated from the Curcuma xanthorrhiza Roxb, has been identified as an antibacterial agent. The MIC values
against B. cereus, Clostridium perfringens, Listeria monocytogenes, S. aureus, S.
Typhimurium, and Vibrio parahaemolyticus are reported to be 8.0–16.0 μg ml−1 .
Xanthorrhizol (280) maintained its antibacterial activity after thermal treatments
(121 ∘ C, 15 min) under various pH ranges (pH 3.0, 7.0, and 11.0). These results
strongly suggest that 280 confers strong antibacterial activity with thermal and
pH stability, and can be effectively used as a natural preservative to prevent the
growth of food-borne pathogens [183].
Three linear sesquiterpene lactones, anthecotulide (281), hydroxyanthecotulide
(282), acetoxyanthecotulide (283), and taraxast-20(30)en-3β-ol (284), are known
as the secondary metabolites of Anthemis auriculata. The in vitro activity of compounds 281–284 against E. coli, P. mirabilis, Agrobacterium tumefaciens, P. aeruginosa, Pseudomonas tolaasii, S. enteritidis, S. aureus, M. luteus, Sarcina lutea, and
B. cereus with MIC values ranging between 29 and 202 μM has been reported
[184].
267
268
9 Natural Products as Antimicrobial Agents – an Update
O
HO
R1
O
HO
O
O
O
OH
O
O
O
OH
O
OH
R2
OH
O
O
OH
281
280
277: R1 = OMe, R2 = H
278: R1 = H, R2 = H
279: R1 = OH, R2 = OH
O
H
OH
H
H
O
O
H
O
O
O
282
H
OAc
HO
H
283
284
Eudesmanes carissone (285), dehydrocarissone (286), and carindone (287),
obtained from the wood of Carissa lanceolata R.Br. (Apocynaceae), showed weak
antibacterial activity against S. aureus, E. coli, and P. aeruginosa with an MIC of
<0.5 mg ml−1 [185].
HO
O
O
O
OH
OH
285
O
286
OH
O O
287
The leaves and bark of Pleodendron costaricense produced cinnamodial (288)
and cinnamosmolide (289), mukaadial (290), and a drimane-type sesquiterpene,
parritadial (291). Antifungal evaluation tests revealed that 288 exhibited high
activity against A. alternata (MIC = 3.9 μg ml−1 ), C. albicans azole-resistant strain
D10, and Wangiella dermatitides (MIC = 15.6 μg ml−1 ). Compound 289 showed
less potent antifungal activities than 288 but was more effective against the C.
albicans azole-resistant strain CN1A (MIC = 23.4 μg ml−1 ) and Pseudallescheria
boydii (MIC = 78.1 μg ml−1 ). A combination of the dialdehyde sesquiterpenes with
dillapiol showed a synergistic antifungal effect with 288 and an additive effect with
290 and 291 [186].
Conventional chromatographic techniques and bioassay-guided fractionation
of the extract of Vernonia colorata (Compositae) resulted in the isolation of
sesquiterpene lactones – vernolide (292), 11β,13-dihydrovernolide (293), and
9.14
Antimicrobial Terpenoids
269
vernodalin (294). The isolates afforded moderate antibacterial activity with an
MIC of 0.1 to 0.5 μg ml−1 against Gram-positive bacteria [187].
A group of cariogenic bacteria are responsible for dental caries, one of the main
oral human diseases [188]. The organisms responsible such as Streptococcus and
Lactobacillus species produce a biofilm on the tooth surface called dental plaque.
S. mutans is considered one of the main cariogenic microorganisms, since it is
responsible for the beginning of the caries process. Other aerobic bacteria such as
E. faecalis, Lactobacillus casei, Streptococcus mitis, S. sanguinis, S. sobrinus, and
Streptococcus salivarius are also important in the later formation of the dental
biofilm [189]. To kill these pathogens, natural products have also been identified
as potential candidates [189, 190]. For example, the activity of bafoudiosbulbins
A (295) and B (296), isolated from the tubers of Dioscorea bulbifera L. var sativa,
was compared with the potential of chloramphenicol and amoxicillin. Both the
compounds moderately inhibit the growth of P. aeruginosa, S. typhi, Salmonella
paratyphi A, and S. paratyphi [191].
HO
H
R1 O
O
R3
O
O
O
O
H
290: R1 = OH, R2 = OH, R3 = H
OAc
O
289
292
291: R1 = OH, R2 = OAc, R3 = OAc(β)
O
H
O
H
O
294
O
O
HH O
O CH3
295
H
HO
O
OH
O
293
H
O
H
O
H
O
O
OH
O
O
O
O
O
O
O
OH
O
R2
288: R1 = OH, R2 = OAc, R3 = H
O
O
O
H
H
O
O
O
O
H
O
296
The kaurane diterpenes, ent-kaur-16(17)-en-19-oic acid (297) isolated from
Aspilia foliacea showed potent activities with MIC values of 10.0 μg ml−1 against
oral pathogens S. sobrinus, S. mutans, S. mitis, S. sanguinis, and L. casei [190a, 192,
193]; therefore, compound 297 can be a lead to the development of anti-plaque
drugs. Glycoside of 297, 18-β-L-3′ ,5′ -diacetoxyarabinofuranosyl-ent-kaur-16ene (298) isolated from Sagittaria pygmaea also inhibited the growth of oral
pathogens, S. mutans ATCC 25 175 and Actinomyces viscosus ATCC 27 044, with
an MIC value of 15.6 μg ml−1 [194].
(−)-Copalic acid (299) isolated from the resin of Copaifera langsddorffii was
also reported to exhibit potent activity (MIC ranging between 2.0 and 6.0 μg ml−1 )
against the main microorganisms responsible for dental caries, S. salivarius, S.
270
9 Natural Products as Antimicrobial Agents – an Update
sobrinus, S. mutans, S. mitis, S. sanguinis, and L. casei. However, other analogs
of this compound were either weakly active or inactive [195]. The mechanism of
action of these compounds has not yet been reported; however, Urzúa et al. [196]
have suggested that these metabolites promote bacterial lysis and disruption of
the cell membrane [196].
CO2H
H
CO2H
297
O
298
H
O
R
R = 3′,5′-diacetoxyarabinofuranoyl
299
Other terpenoids of pimarane type-skeleton have also been reported to possess
anti-plaque activity. For example, ent-pimara diterpenoids (300–302, 304) isolated from Viguiera arenaria Baker and the sodium salt of compound 300 (305)
were evaluated in vitro against S. salivarius, S. sobrinus, S. mutans, S. mitis, S. sanguinis, and L. casei. All the compounds were potentially active, exhibiting MIC
values ranging from 2 to 8 μg ml−1 [190b].
A labdane diterpenoid, 6α-malonyloxymanoyl oxide (306), isolated from the
aerial parts of Stemodia foliosa, inhibited the growth of S. aureus, B. cereus, B.
subtilis, M. smegmatis, and M. phlei with an MIC of 7.0–15.0 μg ml−1 [197]. The
diterpenoid hardwickiic acid (307) obtained from the stem bark of I. gabonensis exhibited potent inhibition against five Gram-negative and four Gram-positive
bacteria with the MIC values of 1.22–4.8 μg ml−1 , with the lowest reported against
Neisseria gonorrhoeae [110].
O
O
H
OH
R2
H
R1
300: R1 = CO2H, R2 = H
301: R1 = CH3, R2 = OH,
302: R1 = CH3, R2 = OCOCH3
303: R1 = CH2OH, R2 = OH
305: R1 = CO2Na, R2 = H
H
CH2OH
H
O
O
304
O
OH
306
CO2H
307
3,4-Epoxyclerodan-13E-en-15-oic acid (308), 5R,8R-(2-oxokolavenic acid)
(309) and 3,4-dihydroxyclerodan-13Z-en-15-oic acid (310) were identified as
the constituents of fruit pulp of Detarium microcarpum. Compound 309 slightly
inhibited the growth of plant pathogenic fungus Cladosporium cucumerinum at
100 μg, while compounds 308 and 310 were moderate inhibitors at 10 μg [198].
The ent-clerodane-type diterpenoids, 311 and 312, were extracted from the
9.14
Antimicrobial Terpenoids
271
aerial parts of Pulicaria wightiana. Both the compounds moderately inhibit the
growth of Gram-positive organisms B. subtilis, Bacillus sphaericus, and S. aureus.
However, they were found to be less active against the Gram-negative organisms
K. aerogenes and Chromobacterium violaceum, revealing their selectivity [199].
The ent-rosane diterpenoids, sagittines A–E (313–317), isolated from Sagittaria sagittifolia, were evaluated against S. mutans ATCC 25175 and Actinomyces
naeslundii ATCC 12104 and were found active with MIC values between 62.5 and
125 μg ml−1 . Compound 317 was active against only A. naeslundii ATCC 12104,
with an MIC value of 62.5 μg ml−1 [200].
O
O
OH
H
H
OH
O
OH
H
O
O
308
309
H
HO
HO
310
O
X
H
H
H
ROH2C
HOOC
MeO
O
H
H
OH
311: X = O
312: X = H2
313
314: R = arabinofuranosyl
315: R = 5′-acetoxyarabinofuranosyl
316: R = 2′-acetoxyarabinofurnaosyl
317: R = 2′,5′-diacetoxyarabinofurnaosyl
Wellsow et al. [201] investigated Plectranthus saccatus for its antimicrobial secondary metabolites and isolated coleon A (318) from Plectranthus puberulentus,
and coleon U (319) and coleon U quinine (320) from lactone Plectranthus forsteri.
These compounds were assayed for their antimicrobial activities and the results
showed that compound 318 was moderately active against B. subtilis and Pseudomonas syringae. Compound 319 was found to be more potent against these two
tested organisms with MICs of 3.13 and 6.25 μg ml−1 , respectively. Compound 320
exhibited moderate activity against the Gram-positive pathogen B. subtilis but it
was more potent against the Gram-negative bacterium P. syringae.
Ferruginol (321), 12-methoxy-6α,11-dihydroxyabieta-8,11,13-triene (322),
cryptojaponol (323), 6α-hydroxysugiol (324), Sugiol (325), 7-hydroxy-11,14dioxo-8,12-abietadiene (326), isopimarol (327), and isopimaric acid (328) were
obtained from the bark of Cryptomeria japonica. The antifungal activities of
these diterpenes were evaluated against the phytopathogenic fungi A. alternata,
P. oryzae, R. solani, and F. oxysporum. Compounds 321–328 showed moderate
antifungal activity (4–85% inhibition at 100 μg/culture) [202]. Chiisanogenin
(329) was isolated from the leaves of Acanthopanax senticosus as a broad
272
9 Natural Products as Antimicrobial Agents – an Update
spectrum antibacterial drug. It showed moderate antibacterial activity against
Gram-positive and Gram-negative bacteria, the MIC being in the range of
50–100 μg ml−1 [203].
HO
O
O
OH
O
OH
O
O
O
OH O
318
R3
R2
OH
OH
H
O
OH
OH
319
320
R1
321: R1 = H, R2 = H, R3 = OH
322: R1 = OH, R2 = OH, R3 = OMe
R1
R1
O
O
H
H
O
O
R1
H
323: R1 = H, R2 = OH, R3 = OMe
324: R1 = OH, R2 = H, R3 = OH
325: R1 = H, R2 = H, R3 = OH
H
O
R
H
HO
COOH
H
OH
327: R = CH2OH
328: R = COOH
326
329
15-β-Isovaleryloxy-ent-kaur-16(17)-en-19-oic acid (330), along with 297,
isolated from A. foliacea, showed activity against oral pathogens S. sobrinus, S.
mutans, S. mitis, S. sanguinis, and L. casei with an MIC value of 100 μg ml−1
[190a].
The antibacterial activity of casbane diterpene (331), a macromonocyclic
isolated from Croton nepetaefolius, showed antibacterial activity against Streptococcus oralis with an MIC value of 62.5 μg ml−1 , whereas this compound weakly
inhibited the growth of S. mutans, S. salivarius, S. sobrinus, S. mitis, and S.
sanguinis with MIC ranging between 125 and 500 μg ml−1 . These results revealed
that compound 331 can also be a candidate for the development of a natural
treatment against oral pathogens for dental biofilms [204].
H
H
OH
O
CO2H
O
H
O
H OH
330
331
Antimicrobial activities have also been described for pentacyclic triterpenes,
such as oleananes, ursanes [205], friedelanes [206], and lupanes [207]. It is speculated that the mechanism of action of triterpenes is due to a disruption in the
microorganism’s cellular membrane [205, 207].
9.14
Antimicrobial Terpenoids
273
9.14.2
Antimicrobial Terpenoids from Microbial Sources
Anti-plaque agents have not only been isolated from plants, but microbial sources
are also utilized to get important anticariogenic. natural products. For example,
ent-8(14),15-pimaradien-19-ol (332) was produced by fungal biotransformation.
This compound displayed very promising antibiotic activity with MIC values
(ranging from 1.5 to 4.0 μg ml−1 ) against S. salivarius, S. sobrinus, S. mutans, S.
mitis, S. sanguinis, and L. casei [208]. However, other biotransformed products
(333–335) were relatively less active, indicating the importance of the endocyclic
double bond and alcoholic group in such skeleton.
Other microbial-transformed products of the antibacterial diterpene sclareol
(336), (6α)-6,18-dihydroxysclareol (337), (1β)-1-hydroxysclareol (338), and
(12S)-12-hydroxysclareol (339) were found to exhibit increased antibacterial
activities when compared with the precursor 336. Compound 338 was found to
be more active against B. subtilis than sclareol proper (14- vs 12-mm inhibition
zone, respectively). Compounds 337 and 339 were found to be approximately as
active against B. subtilis as sclareol (336), with inhibition zones of 12 and 10 mm,
respectively [209].
A fusidane triterpene, 16-deacetoxy-7-β-hydroxy-fusidic acid (340), was
isolated from the culture of mitosporic fungus Acremonium crotocinigenum.
This compound was tested against a panel of MDR and MRSA strains and
showed MIC values of 16 μg ml−1 , which, although occasionally more active than
erythromycin and norfloxacin, was significantly less potent than the fusidic acid
comparator [210].
H
HOH2C
H
CO2H
H
CO2H
332
333
O
H
CO2H
334
335
HO
R2
HO
HO
R1
H
HO
H
H
H
OH
338: R1 = OH, R2 = H
336
337
H
HO
OH
OH
OH
H
H
339: R1 = H, R2 = OH
HO
H
OH
340
CO2H
274
9 Natural Products as Antimicrobial Agents – an Update
Fridelin (341) is known as the constituent of V. laurentii De Wild and is reported
to possess a broad-spectrum antimicrobial character against six Gram-negative,
four Gram-positive, and two fungal strains with an MIC of 2.4–9.7 μg ml−1 [82].
Besides the above terpenoids, sordarins are another group of natural products that
are distributed mostly across the filamentous fungi, and are also well known for
their antimicrobial properties [211].
OH
O
O
H
OMe
OH
H
OHC
O
341
CO2H
Sordarin
9.14.3
Antimicrobial Terpenoids from Marine Sources
Laurencia majuscula Harvey collected from two locations in the waters of
Sabah, Malaysia, produced two major halogenated compounds: elatol (342) and
iso-obtusol (343). At a concentration of 30 μg/paper disk 342 inhibited six species
of bacteria with significant antibacterial activities (zone of inhibition ranging
from 7 to 30 mm) against S. epidermidis, K. pneumonia, and Salmonella sp., while
343 exhibited antibacterial activity against four bacterial species with significant
activity (zone of inhibition ranging from 7 to 24 mm) against K. pneumonia and
Salmonella sp. [212].
Compounds laurinterol acetate (344), laurinterol (345), debromolaurinterol
acetate (346), and debromolaurinterol (347) isolated from the sea hare, Aplysia
kurodai, exhibited antibacterial activity against S. aureus with the zone of
inhibition between 7 and 16 mm at a concentration of 12.5–50 μg/paper disc
[213]. Another diterpene, xeniolides I (348), isolated from the Kenyan soft coral
Xenia novaebrittanniae, possesses antibacterial activity at a concentration of
1.25 μg ml−1 [214].
O
HO
HO
Br
Br
R1
OR2
Br
Cl
Cl
342
343
344: R1 = Br: R2 = Ac
345: R1 = Br: R2 = H
346:R1 = H: R2 = Ac
347: R1 = R2 = H
H
O
OH
OH
O
O
H
348
9.15
Miscellaneous Antimicrobial Compounds
9.15
Miscellaneous Antimicrobial Compounds
9.15.1
Miscellaneous Antimicrobial Natural Products from Plants
Beilschmiedic acid derivatives (349–358) have been reported from the leaves
of a Beilschmiedia gabonese. Compounds 350–357 have been reported to
possess antibacterial activity against MRSA with MIC values in the range of
10–13 μg ml−1 . Despite their important antibacterial activities, these compounds
may not be good antibiotic candidates for being cytotoxic in nature [215].
349: R1 = α−OH, R2 =
350: R1 = β−OH, R2 =
351: R1 = H, R2 =
OH
H
R1
352: R1 = α−OH, R2 =
O
H
HO
353: R1 = α−OH, R2 =
H
H
H
H
354: R1 = α−OH, R2 =
R2
355: R1 = α−OH, R2 =
356: R1 = α−OH, R2 =
O
O
O
357: R1 = β−OH, R2 =
358: R1 = α−OH, R2 =
Three antifungal compounds pseudodillapiol (359), eupomatenoid-6 (360), and
conocarpan (361) were isolated from the aerial parts of Piper abutiloides. These
compounds exhibited different potencies against the fungal species C. albicans,
Candida parapsilosis, C. krusei, C. glabrata, C. tropicalis, C. neoformans, and
S. schenckii with the strongest effect (0.3 μg/spot) being observed for compound
360 against C. glabrata. These compounds were further declared as non-cytotoxic
in nature.
275
276
9 Natural Products as Antimicrobial Agents – an Update
Icariside B6 (362) was purified from the leaves of Acer truncatum as an antiB. cereus and anti-M. luteus agent. At a concentration of 24 μg/disk in standard
Petri-plate assays, the compound exhibited inhibitory zone sizes of 14 and 17 mm,
respectively [216]. Sedanolide (363) obtained from Apium a graveolens seed was
identified as the inhibitor of C. albicans and C. parapsilosis at a concentration of
100 μg ml−1 [217]. Botcinins B (364), D (365), and E (366), and 3-O-deacetyl-2epi-botcinin A (367) and botcinin E (368) isolated from B. cinerea were evaluated
for their antifungal activities. Compounds 366, 367, and 368 showed weak antifungal activity, all with MICs of 100 μM, against Magnaporthe grisea. Botcinins B
(364) and D (365) are reported to exhibit better activity (both with MIC 12.5 μM)
[218].
O
O
O
OMe
O
HO
OMe
HO
359
360
361
O
O
H
R1
O
HO
O
OH OH
362
R4
O
R2
O
HO
O
O
O
OR3
364: R1 = CH3, R2 = H, R3 = Ac, R4 = (CH2)2CH3
366: R1 = CH3, R2 = H, R3 = H, R4 = CH3
367: R1 = H, R2 = CH3, R3 = H, R4 = CH3
O
363
OH
O
O
HO
OH
H
O
O
O
O
HO
O
365
O
OH
O
368
The cytotoxic compound aculeatin D (369) is reported as a minor constituent
from the rhizomes of Amomum aculeatum, which showed remarkable activity
against two Plasmodium falciparum (MIC of 0.42 μg ml−1 ) strains, as well as
against Trypanosoma brucei rhodesiense (MIC of 0.20 μg ml−1 ) and Trypanosoma
cruzi (MIC of 0.49 μg ml−1 ). In an antibacterial assay against B. cereus (MIC of
16 μg ml−1 ), E. coli (MIC of 16 μg ml−1 ), and S. epidermidis (MIC of 8 μg ml−1 ) the
compound showed moderate to strong activity [219]. Carneic acids A (370) and B
(371) were isolated as major constituents of the stromata of Hypoxylon carneum.
Their chemical structures were elucidated by a combination of spectroscopic
methods and by the preparation of derivatives. An X-ray crystal structure of the
9.15
Miscellaneous Antimicrobial Compounds
277
dinitrobenzoate of carneic acid B methyl ester (372) helped in the determination
of its absolute structure. Carneic acids (370 and 371) showed activity against
filamentous fungi Mucor hiemalis, Penicillium griseofulvum, Stachybotrys chartarum, and Trichoderma harzianum (MIC = 12.0–50 μg ml−1 ), and the yeast
Yarrowia lipolytica (MIC = 25 μg ml−1 ). B. subtilis was about equally affected
(MIC = 50 μg ml−1 ) by rubiginosin A (373), 1,1′ -binaphthalene-4,4′ -5,5′ -terol
(BNT) (374) [220].
O
COOR1
O
370: R1 = H; R2 = H;
H
OH
O
R2
*
372: R1 = Me; R2 = R3 = p-NO2-C6H4-COO-
H
OR3
10
369
R3 = H
371: R1 = H; R2 = OH; R3 = H
HO
HO
OH
HO
OH
O
OH
H
OAc
O
HO
O
O
374
373
Two more cytotoxic compounds, 2,4-dihydroxy-6-[(10E,30E)-penta-10,30dienyl]-benzaldehyde (375) and periconicin B (376), were isolated from the
leaves of Xylopia aromatica, a native plant of the Brazilian Cerrado. Both the
compounds were evaluated against C. sphaerospermum and C. cladosporioides
using direct bioautography and 375 exhibited strong antifungal activity against
both fungi, showing a detection limit of 1.0 μg, comparable to nystatin (used
as positive control). Compound 376 showed a relatively weak detection limit
of 25.0 μg [221]. The phytochemical analysis of the aerial parts of Salsola
tetrandra gave two moderate antimicrobial agents, taxiphyllin (377) and S-(−)trans-N-feruloyloctopamine (378), which displayed mild antibacterial activity
(MIC = 200 μg ml−1 ) against S. aureus, whereas compound 377 showed the
highest activity (MIC = 0.96 μM) in the Artemia salina bioassay [222].
O
H
O H
HO
H
OH O
375
OH
Glu O
OMe
CN
OH
H
HO
OH
376
377
HO
OH
H
N
O
378
278
9 Natural Products as Antimicrobial Agents – an Update
The ethanolic extract of Trichodesma indicum (Linn.) comprises several fatty
acids and their derivatives including n-decanyl laurate (379), n-tetradecanyl
laurate (380), n-nonacosanyl palmitate (381), stigmast-5-en-3β-ol-21(24)-olide
(382), n-pentacos-9-one (383), n-dotriacont-9-one-13-ene (384), stigmast-5-en3β-ol-23-one (385), and lanast-5-en-3β-D-glucopyranosyl-21-(24)-olide (386).
Most of the fatty acids displayed inhibitory activity against S. aureus, B. subtilis,
and C. albicans with varying MIC values (2.4–20.6 μg ml−1 ) [223]. Fatty acids
have been identified to afford antimicrobial properties and a few are already in
the market.
O
O
O
O
8
6
8
10
379
O
O
25
381
380
O
12
12
4
O
O
383
HO
4
14
382
O
384
O
O
O
HO
385
HO
386
9.15.2
Miscellaneous Antimicrobials from Bacteria
A non-proteinogenic amino acid L-furanomycin [2S,2′ R,5′ S)-2-amino-2(5′ methyl-2′ ,5′ -dihydrofuran-2′ -yl)acetic acid (387)] has been identified as the
constituent of culture extract of P. fluorescens SBW25 [224]. Compound 387
has also been reported from the cultures of Streptomyces threomyceticus ATCC
15795 [225]. This compound potentially inhibited the growth of several bacterial
strains, including a number of plant pathogenic microbes.
9.15
Miscellaneous Antimicrobial Compounds
279
H O
O
OH
NH2
387
OMe
O
MeO
OMe
HO
O
O
H
H
O
O
OMe R
O
O
HO
O
O
O
O
HO
HN
O
MeO
O
H
HO
H
O
HO
O
H
H
O
H
O
O
HN
O
O
O
H
OMe NH2
H
O
HO
O
H
O
HO
388: R = NH2
389: R = NO2
O
O
390
Culture extracts of Actinomycetes strain, designated as MS100061 yielded
lobophorins A (388), B (389), and G (390). These compounds inhibited the
growth of Mycobacterium bovis bacillus calmette-guérin (BCG) with MIC values
of 1.56, 1.56, and 0.78 μg ml−1 , respectively, and M. tuberculosis H37Rv with MIC
values of 32, 32, and 16 μg ml−1 , respectively. Further, these compounds
exhibited MIC values against B. subtilis as 3.12, 12.5, and 1.56 μg ml−1 ,
respectively. These compounds can be potential candidates for antibiotic
development [226].
Synoxazolidinone A (391) and B (392), the constituents of Synoicum pulmonaria, displayed MIC values against the Gram-positive bacteria S. aureus
and MRSA at a concentration of 10 μg ml−1 . Synoxazolidinone A (391) inhibited
the growth of the Gram-positive bacterium Corynebacterium glutamicum with
an MIC value of 12.5 μg ml−1 , whereas the growth of fungus S. cerevisiae was
inhibited by an MIC value of 6.25 μg ml−1 . Compound 392 displayed relatively
low potential (MIC of 30 μg ml−1 ) against MRSA. This lowered activity may
suggest that the chlorine atom is important for biological activity [227].
280
9 Natural Products as Antimicrobial Agents – an Update
HN
HN
NH2
Cl H
H
O
NH
Br
MeO
Br
O
NH2
NH
NH
Br
H
O
MeO
NH
Br
O
392
391
9.15.3
Miscellaneous Antimicrobials from Fungi
A derivative of benzofuran (393) was purified from the culture of Phomopsis
sp. hzla01-1 and was characterized by 1D and 2D NMR spectroscopic analyses.
Compounds 393 showed significant antimicrobial activities against E. coli
CMCC44103, C. albicans AS2.538, and S. cerevisiae ATCC9763 with MICs of
5–10 μg ml−1 and showed the strongest activity against B. subtilis CMCC63501
with an MIC value of 1.25 μg ml−1 . None of the three compounds showed
significant cytotoxicity against the HeLa cell line at 10 μg ml−1 [136].
O
OH
O
393
The compounds rotiorinols A (394), C (395), (−)-rotiorin (396), and rubrorotiorin (397) were isolated from the fungus Chaetomium cupreum CC3003. The
absolute configuration of 394 was determined by the modified Mosher’s method
along with an X-ray analysis of its acetate derivative, as well as by chemical transformation. Compounds 394–397 exhibited antifungal activity against C. albicans
with IC50 values of 10.5, 16.7, 24.3, and 0.6 μg ml−1 , respectively [228].
O
O
O
O
O
OH
OH
O
O
O
OH
395
394
Cl
O
O
O
O
O
O
O
O
396
O
O
397
9.15
Miscellaneous Antimicrobial Compounds
Two cyclopentenones VM 4798-1a (398) and VM 4798-1b (399) were obtained
as a 3 : 1 inseparable mixture from fermentations of Dasyscyphus sp. A47-98. The
structures were elucidated by NMR techniques. The cytotoxic mixture of these
two isomers showed weak antibacterial and antifungal properties with MIC of
50–100 μg ml−1 [229].
Cultures of the freshwater aquatic fungus Helicodendron giganteum afforded
heliconols A (400), which contains an unusual reduced furanocyclopentane unit.
The structure of this metabolite was assigned by analysis of 1D and 2D NMR data.
The absolute configuration of heliconol A (400) was assigned by single-crystal Xray crystallographic analysis of its dibromobenzoate derivative. This compound
inhibits the growth of A. flavus and F. verticillioides at 200 μg/disk affording a 15mm diameter zone of partial clearing after 48 h. Heliconol A (400) also exhibited
activity against C. albicans, S. aureus, and B. subtilis with clear inhibition zones
measuring 18, 23, and 35 mm diameter, respectively, after 48 h at 100 μg/paper
disc [230].
Ascopyrone P (401), the secondary metabolite of the fungi Anthracobia
melaloma, Plicaria anthracina, Plicaria Leiocarpa, and Peziza petersi, was
evaluated for its antibacterial properties. At a level of 2000 mg l−1 , 401 demonstrated growth inhibitory activity against a broad range of Gram-positive and
Gram-negative bacteria. This study suggests its possible application in food
preservation, to control the growth of Gram-negative and Gram-positive bacteria
in raw and cooked foods [231].
Phenylacetic acid (402) and sodium phenylacetate (403) were isolated from
Streptomyces humidus strain S5-55. Both the compounds completely inhibited
the growth of Pythium ultimum, Phytophthora capsici, R. solani, S. cerevisiae,
and P. syringae pv. syringae at concentrations ranging between 10 and 50 μg ml−1 .
The two compounds were as effective as the commercial fungicide metalaxyl in
inhibiting spore germination and hyphal growth of P. capsici [232]. Lavermicocca
et al. [233] also obtained compound 402 from Sourdough Lactobacillus plantarum Strain 21B and found it to be active against A. niger, A. flavus, Eurotium
rubrum, Eurotium repens, Endomyces fibuliger, Penicillium corylophilum, Penicillium roqueforti, and Monilia sitophila at a concentration of 50 mg ml−1 .
O
O
CO2Me
OH
Cl
HO HO
Cl
Cl
398
Cl
CO2Me
OH
OH
O
OH
400
399
O
ONa
OH
HO
O
401
O
O
OH
402
403
281
282
9 Natural Products as Antimicrobial Agents – an Update
9.16
Platensimycin Family as Antibacterial Natural Products
Platensimycin (404) is a novel broad-spectrum Gram-positive antibacterial
natural product isolated from the cultures of Streptomyces platensis [234].
This unique substance was discovered by target-based whole-cell screening
strategy using antisense differential sensitivity assay and was purified using
Amberchrome, Sephadex LH-20, and reversed-phase HPLC. The structure
was elucidated by 2D NMR methods and confirmed by X-ray crystallographic
analysis of a bromo derivative. It inhibits the growth of bacterial pathogens S.
aureus and E. coli. It is the molecule of interest because of its modern mode
of action. It inhibits bacterial growth by selectively inhibiting the condensing
enzyme FabF of the fatty acid synthesis pathway, with IC50 values of 48 and 160
nM. This compound inhibits phospholipid synthesis (IC50 = 0.1 μg ml−1 ) and
exhibits MIC values of 0.1–1 μg ml−1 against MRSA. Mechanistically, a novel
mode of action that involves specific binding with the acyl enzyme intermediate
of the key condensing enzyme FabF is involved in its activity. Therefore, because
of its unique mode of action, it can be a potential candidate for antibacterial drug
development [235]. After the discovery of platencimycin (404), a variety of its
analogs have been isolated or synthesized to develop new antibiotics [234].
Platencin (405), isoplatensimycin (406), platensimycin A1 (407) and its methyl
ester (408), and platensimycin B1 –B3 (409–411) are discovered from bacterial
sources possessing antibiotic properties. These antibiotics attract much attention
because of their novel mode of action against MDR bacterial strains. Not only
the naturally occurring molecules but also their synthetic analogs, such as
carbaplatensimycin (412), adamantaplatensimycins (413), oxazinidinyl platensimycin (414), platencin A1 (415), (−)-nor-platencin (416), homoplatensimide
(417), and platensimide, are under study [234].
It is concluded that hundreds of natural products isolated from various
natural sources belonging to several classes of secondary metabolites exhibit
antimicrobial properties. A yearwise review of the literature revealed that little
work has been done on antimicrobial natural products; however, it is increasing
day by day. Although substantial work has been published during this decade,
it still seems insufficient because of the rapid appearance of antibiotic-resistant
strains. In the past, most of the antimicrobials were derived from microorganisms; still microorganisms together with plants can be good sources of new
antimicrobials [236].
Compounds of different classes have been observed to exhibit variable antimicrobial activities such as the polypeptides that have been identified as potential
antimicrobial agents. Quinones are the other group of secondary metabolites
affording potential activities. The reason may lie in the fact that besides providing
a source of stable free radicals, quinones are known to complex irreversibly with
nucleophilic amino acids in proteins, resulting in loss of protein function [166].
For that reason, the potential range of quinone antimicrobial effects is great.
9.16
OH
O
HO2C
OH
Platensimycin Family as Antibacterial Natural Products
OH
O
O
N
H
HO2C
OH
O
N
H
O
404
RO
405
OH
O
O
H2N
O
OH
O
OH
OH
O
HO
OH
O
O
O
N
H
O
HN
O
409
OH
O
O
HO
N
OH H
O
OH
410
OH
O
O
H2N
N
H
O
N
H
O
O
407: = R = H
408: = R = Me
CO2H
406
OH
O
N
H
OH
H
N
O
O O
OH
O
283
O
OH
N
N
H
OH
O
O
O HO2C
OH
HO
N
H
O
O
411
412
H
N
413
414
OH
CO2H
O
O
O
HO2C
H2N
O O
HO
HO2C
O
N
H
N
H
O
N
H
O
O
415
416
417
O
O
284
9 Natural Products as Antimicrobial Agents – an Update
On the other hand, the possible toxic effects of quinones obtained from natural
sources must be thoroughly examined before bringing the new isolates into
practice. Terpenoids is another group of natural products that is identified as a
candidate for antimicrobial drug development.
Besides the above three groups, some diverse and unusual structures isolated
from natural sources may also be the future potential candidates. For example,
platensimycin (404) was identified from S. platensis as a novel broad-spectrum
Gram-positive antibiotic. These facts and figures demand that maximum natural
sources be explored using advanced and sophisticated techniques for purification
and characterization. The use of modern techniques in future may result in the
isolation of more potential antimicrobial drug candidates.
References
1. Soejarto, D.D. and Farnsworth, N.R.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
(1989) Perspect. Biol. Med., 32,
244–256.
Li, J.W.H. and Vederas, J.C. (2009)
Science, 325, 161–165.
Koehn, F.E. and Carter, G.T. (2005)
Nat. Rev. Drug Discovery, 4, 206–220.
Mishra, B.B. and Tiwari, V.K. (2011)
Eur. J. Med. Chem., 46, 4769–4807.
Hamilton-Miller, J.M.T. (2008) Int. J.
Antimicrob. Agents, 31, 189–192.
Butler, M.S. (2008) Nat. Prod. Rep., 25,
475–516.
(a) Kobayashi, R., Konomi, M.,
Hasegawa, K., Morozumi, M.,
Sunakawa, K., and Ubukata, K. (2005)
Antimicrob. Agents Chemother., 49,
889–894; (b) Thomson, K.S. and
Moland, E.S. (2004) J. Antimicrob.
Chemother., 54, 557–562.
Khosla, C. (2010) J. Biol. Chem., 285,
27499–27531.
Behal, V. (2001) Folia Microbiol., 46,
363–370.
Omulokoli, E.B.K. and Chhabra, S.C.
(1997) J. Ethnopharmacol., 56,
133–137.
Al-Rehaily, A.J., Ahmad, M.S.,
Mustafa, J., Al-Oqail, M.M.,
Hassan, W.H., Khan, S.I., and
Khan, I.A. (2013) J. Saudi Chem. Soc.,
17, 67–76.
Martin, F., Grkovic, T., Sykes, M.L.,
Shelper, T., Avery, V.M., Camp, D.,
Quinn, R.J., and Davis, R.A. (2011) J.
Nat. Prod., 74, 2425–2430.
13. Zuo, G.Y., Meng, F.Y.,
14.
15.
16.
17.
18.
19.
20.
21.
22.
Hao, X.Y., Zhang, Y.L.,
Wang, G.C., and Xu, G.L. (2008)
J. Pharm. Pharm. Sci., 11, 90–94.
Ileana, I.R., Abimael, D.R.,
Yuehong, W., and Scott, F.G. (2006)
Tetrahedron Lett., 47, 3229–3232.
Pitayya, T., Photchana, P., Yupa, P., and
Walter, C.T. (2006) Chem. Pharm. Bull.,
54, 149–151.
Asolkar, R.N., Schroder, D.,
Heckmann, R., Lang, S.,
Wagner-Dobler, I., and Laatsch, H.
(2004) J. Antibiot., 57, 17–23.
Oliva, O.A., Kumudini, M.M.,
David, E.W., Dewayne, H., Amber, L.H.,
Giovanni, A., and Stephen, O.D. (2003)
J. Agric. Food Chem., 51, 890–896.
O’Donnell, G. and Gibbons, S. (2007)
Phytother. Res., 21, 653–657.
Céline, T., Jean-Charles, G.,
Philippe, D., Christophe, F.,
Reynald, H., Maria-Elena, F.,
Antonieta, R.A., and Alain, F. (2003)
Phytother. Res., 17, 678–680.
Anjum, A., Ekramul Haque, M.,
Mukhlesur Rahman, M., and
Sarker, S.D. (2002) Fitoterapia, 73,
526–528.
Preetha, R., Jose, S., Prathapan, S.,
Vijayan, K.K., Jayaprakash, N.S.,
Philip, R., and Singh, I.S.B. (2010)
Aquacult. Res., 41, 1452–1461.
Sato, A., Takahashi, S., Ogita, T.,
Sugano, M., and Kodama, K. (1995)
Annu. Rep. Sankyo Res. Lab., 47, 1–58.
References
23. Turner, J.M. and Messenger, A.J. (1986)
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
Adv. Microb. Physiol., 27, 211–275.
Kerr, J.R. (2000) Infect. Dis. Rev., 2,
184–194.
(a) Maul, C., Sattler, I., Zerlin, M.,
Hinze, C., Koch, C., Maier, A.,
Grabley, S., and Thierick, R. (1999)
J. Antibiot., 52, 1124–1134; (b)
Yagishita, K. (1960) J. Antibiot., Ser.
A, 13, 83.
Giddens, S.R., Feng, Y., and
Mahanty, H.K. (2002) Mol. Microbiol.,
45, 769–783.
Giddens, S.R. and Bean, D.C. (2007)
Int. J. Antimicrob. Agents, 29, 93–97.
Shibazaki, M., Taniguchi, M., Yokoi, T.,
Nagai, K., Watanabe, M., Suzuki, K.,
and Yamamoto, T. (2004) J. Antibiot.,
57, 379–382.
Kunze, B., Jansen, R., Hofle, G., and
Reichenbach, H.A. (2004) J. Antibiot.,
57, 151–155.
Zheng, C., Kim, C.-J., Bae, K.S.,
Kim, Y.-H., and Kim, W.-G. (2006)
J. Nat. Prod., 69, 1816–1819.
Takahashi, C., Numata, A., Ito, Y.,
Matsumura, E., Araki, H., Iwaki, H.,
and Kushida, K. (1994) J. Chem. Soc.,
Perkin Trans. 1, 1859–1864.
Dong, J.-Y., He, H.-P., Shen, Y.-M., and
Zhang, K.-Q. (2005) J. Nat. Prod., 68,
1510–1513.
Joshi, B.K., Gloer, J.B., and
Wicklow, D.T. (1999) J. Nat. Prod.,
62, 730–733.
Yamada, T., Iwamoto, C., Yamagaki, N.,
Yamanouchi, T., Minoura, K.,
Yamori, T., Uehara, Y., Andoh, T.,
Umemura, K., and Numata, A. (2002)
Tetrahedron, 58, 479–487.
Proksch, P., Edrada, R.A., and Ebel, R.
(2002) Appl. Microbiol. Biotechnol., 59
(2-3), 125–134.
Kossuga, M.H., Lira, S.P., McHugh, S.,
Torres, Y.R., Lima, B.A., Gonçalves, R.,
Veloso, K., Ferreira, A.G., Rocha, R.M.,
and Berlinck, R.G.S. (2009) J. Braz.
Chem. Soc., 20, 704–711.
Al-Zereini, W., Yao, C.B.F.F.,
Laatsch, H., and Anke, H. (2010) J.
Antibiot., 63, 297–301.
Gershon, H. and Shanks, L. (1978) Can.
J. Microbiol., 24, 593–597.
39. Jurenka, S.J. (2002) Altern. Med. Rev., 7,
68–70.
40. Li, X.C., Jacob, M.R., Khan, S.I.,
41.
42.
43.
44.
45.
46.
47.
48.
49.
50.
51.
Ashfaq, M.K., Babu, K.S.,
Agarwal, A.K., Elsohly, H.N.,
Manly, S.P., and Clark, A.M. (2008)
Antimicrob. Agents Chemother., 52,
2442–2448.
Ondeyka, J.G., Zink, D.L., Young, K.,
Painter, R., Kodali, S., Galgoci, A.,
Collado, J., Tormo, J.R., Basilio, A.,
Vicente, F., Wang, J., and Singh, S.B.
(2006) J. Nat. Prod., 69, 377–380.
Kim, H.-J., Vinale, F., Ghisalberti, E.L.,
Worth, C.M., Sivasithamparam, K.,
Skelton, B.W., and White, A.H. (2006)
Phytochemistry, 67, 2277–2280.
Demizu, S., Kajiyama, K., Takahashi, K.,
Hiraga, Y., Yamamoto, S., Tamura, Y.,
Okada, K., and Kinoshita, T. (1988)
Chem. Pharm. Bull., 36, 3474–3479.
Basson, N.J. and Grobler, S.R. (2008)
BMC Complement. Altern. Med., 8, 41.
Hu, J.-F., Garo, E., Hough, G.W.,
Goering, M.G., O’Neil-Johnson, M.,
and Eldridge, G.R. (2006) J. Nat. Prod.,
69, 585–590.
Ren, Q., Siau, W.-Y., Du, Z., Zhang, K.,
and Wang, J. (2011) Chem. Eur. J., 17,
7781–7785.
Chetan, B.S., Nimesh, M.S.,
Manish, P.P., and Ranjan, G.P. (2012) J.
Serb. Chem. Soc., 77, 1–17.
Suresh, T., Arunima, V., Atin, K.,
Sandeep, G., Prarthana, V.R., and
Ganesh, R.K. (2010) Acta Pol. Pharm.,
67, 423–427.
Khafagy, M.M., El-Wahas, A.H.F.A.,
Eid, F.A., and El-Agrody, A.M. (2002)
Farmaco, 57, 715–722.
Nicolaou, K.C., Pfefferkorn, J.A.,
Roecker, A.J., Cao, G.Q., Barluenga, S.,
and Mitchell, H.J. (2000) J. Am. Chem.
Soc., 122, 9939–9953.
(a) Thomas, N. and Zachariah, S.M.
(2013) Asian J. Pharm. Clin. Res., 6,
11–15; (b) Willem, A.L., Lindani, N.E.,
Samuel, K., Garreth, L.M., Simon, S.M.,
and Charles, B.K. (2005) Tetrahedron, 61, 9996–10006; (c) Hester, L.V.,
Wei, Z., Tore, H., Floris, P.J.T., and
Karl, A.J. (2003) Org. Biomol. Chem.,
1, 1953–1958; (d) Charles, D.H.,
Babajide, O.O., Donna, V.E., David, M.,
285
286
9 Natural Products as Antimicrobial Agents – an Update
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
and Jon, C. (1980) J. Am. Chem. Soc.,
102, 7365–7367; (e) Victorine, F.,
Augustin, E.N., Tanee, F., Beibam, L.S.,
and Bernard, B. (1998) J. Nat.
Prod., 61, 380–383; (f ) Jean, W.,
Tanee, F.Z., François, T., Francine, L.,
and Michel, K. (1995) J. Nat. Prod.,
58, 105–108; (g) Turner, C.E. and
Elsohly, M.A. (1981) J. Clin. Pharmacol., 21, 283–291.
Kitamura, R.O.S., Romo, P.,
Young, M.C.M., Kato, M.J., and
Lago, J.H.G. (2006) Phytochemistry,
67, 2398–2402.
Aoyama, Y., Katayama, T.,
Yamamoto, M., Tanaka, H., and Kon, K.
(1992) J. Antibiot., 45, 875–878.
Iranshahi, M., Askari, M., Sahebkar, A.,
and Hadjipavlou-Litina, D. (2009)
DARU J. Pharm. Sci., 17, 99–103.
Braccio, M.D., Grossi, G., Roma, G.,
Signorello, M.G., and Leoncini, G.
(2004) Eur. J. Med. Chem., 39,
397–409.
Curini, M., Epifano, F., Maltese, F.,
Marcotullio, M.C., Tubaro, A.,
Altinier, G., Gonzales, S.P., and
Rodriguez, J. (2004) Bioorg. Med. Chem.
Lett., 14, 2241–2243.
Ito, C., Itoigawa, M., Furukuda, H.,
Tokuda, H., Okuda, Y., Mukainaka, T.,
Okuda, M., and Nishino, H. (1999)
Cancer Lett., 138, 87–92.
Chakthong, S., Weaaryee, P.,
Puangphet, P., Mahabusarakam, W.,
Plodpai, P., Voravuthikunchai, S.P., and
Kanjana-Opas, A. (2012) Phytochemistry, 75, 108–113.
Widelski, J., Popova, M., Graikou, K.,
Glowniak, K., and Chinou, I. (2009)
Molecules, 14, 2729–2734.
Schinkovitz, A., Gibbons, S., Stavri, M.,
Cocksedge, M.J., Bucar, F., and
Ostruthin, F. (2003) Planta Med.,
69, 369–371.
Rosselli, S., Maggio, A., Bellone, G.,
Formisano, C., Basile, A., Cicala, C.,
Alfieri, A., Mascolo, N., and Bruno, M.
(2007) Planta Med., 72, 116–120.
Stavri, M. and Gibbons, S. (2005)
Phytother. Res., 19, 938–941.
Basile, A., Sorbo, S., Spadaro, V.,
Bruno, M., Maggio, A., Faraone, N.,
64.
65.
66.
67.
68.
69.
70.
71.
72.
73.
74.
75.
76.
and Rosselli, S. (2009) Molecules, 14,
939–952.
Zarrilli, R., Ricci, V., and Romano, M.
(1999) Cell Microbiol., 1, 93–99.
Carpinella, M.C., Ferrayoli, C.G., and
Palacios, S.M. (2005) J. Agric. Food
Chem., 53, 2922–2927.
Spedes, C.E., Avila, J.G., Martýńez, A.S.,
Serrato, B., Mugica, J.C.C.N., and
Garciglia, R.S. (2006) J. Agric. Food
Chem., 54, 3521–3527.
Khatune, N.A., Islam, M.E.,
Haque, M.E., Khondkar, P., and
Rahman, M.M. (2004) Fitoterapia,
75, 228–230.
Taechowisan, T., Lu, C., Shen, Y., and
Lumyong, S. (2005) Microbiology, 151,
1691–1695.
(a) Cushnie, T. and Lamb, A. (2005)
Int. J. Antimicrob. Agents, 26, 343–356;
(b) Zhou, M., Luo, H., Li, Z., Wu, F.,
Huang, C., Ding, Z., and Li, R. (2012)
Comb. Chem. High Throughput Screening, 15, 306–315; (c) Nenaah, G. (2013)
World J. Microbiol. Biotechnol., 29,
1255–1262.
Candiracci, M., Citterio, B.,
Diamantini, G., Blasa, M., Accorsi, A.,
and Piatti, E. (2011) Int. J. Food Prop.,
14, 799–808.
Fowler, Z.L., Shah, K., Panepinto, J.C.,
Jacobs, A., and Koffas, M.A.G. (2011)
PLoS One, 6, e25681.
Brahmachari, G., Mandal, N.C.,
Jash, S.K., Roy, R., Mandal, L.C.,
Mukhopadhyay, A., Behera, B.,
Majhi, S., Mondal, A., and
Gangopadhyay, A. (2011) Chem. Biodivers., 8, 1139–1151.
Raghukumar, R., Vali, L., Watson, D.,
Fearnley, J., and Seidel, V. (2010) Phytother. Res., 24, 1181–1187.
Ango, P.Y., Kapche, D.W.F.G., Kuete, V.,
Ngadjui, B.T., Bezabih, M., and
Abegaz, B.M. (2012) Phytochem. Lett.,
5, 524–528.
Kasim, L.S., Ferro, V.A., Odukoya, O.A.,
Drummond, A., Ukpo, G.E., Seidel, V.,
Gray, A.I., and Waigh, R. (2011)
J. Microbiol. Antimicrob., 3, 13–17.
Lee, J.–.Y., Jeong, K.-W., Shin, S.,
Lee, J.-U., and Kim, Y. (2009) Bioorg.
Med. Chem., 17, 5408–5413.
References
77. Cao, M.A., Sun, X.B., Zhaou, P.H., and
78.
79.
80.
81.
82.
83.
84.
85.
86.
87.
88.
89.
Yuan, C.S. (2006) Chin. Chem. Lett., 17,
1048–1050.
Jung, H.J., Sung, W.S., Yeo, S.-H.,
Kim, H.S., Lee, I.-S., Woo, E.-R., and
Lee, D.G. (2006) Arch. Pharm. Res., 29,
746–751.
Ramos, F.A., Takaishi, Y., Shirotori, M.,
Kawaguchi, Y., Tsuchiya, K.,
Shibata, H., Higuti, T., Tadokoro, T.,
and Takeuchi, M. (2006) J. Agric. Food
Chem., 54, 3551–3557.
Hidetoshi, A. and Gen-ichi, D. (2002)
Biosci. Biotechnol., Biochem., 66,
1727–1730.
Donnell, G.O., Bucar, F., and
Gibbons, S. (2006) Phytochemistry,
67, 178–182.
Kuete, V., Nguemeving, J.R., Beng, V.P.,
Azebaze, A.G., Etoa, F.X., Meyer, M.,
Bodo, B., and Nkengfack, A.E. (2007)
J. Ethnopharmacol., 109, 372–379.
Sato, Y., Suzaki, S., Nishikawa, T.,
Kihara, M., Shibata, H., and Higuti, T.
(2000) J. Ethnopharmacol., 72,
483–488.
Sathiamoorthy, B., Gupta, P.,
Kumar, M., Chaturvedi, A.K.,
Shukla, P.K., and Maurya, R. (2007)
Bioorg. Med. Chem. Lett., 17, 239–242.
Drewes, S.E. and van Vuuren, S.F.
(2008) Phytochemistry, 69, 1745–1749.
Kuete, V., Ngameni, B., Simo, C.C.,
Tankeu, P.K., Ngadjui, B.T., Meyer, J.I.,
Lall, N., and Kuiate, J.R. (2008)
J. Ethnopharmacol., 120, 17–24.
Alvarez, M.A., Debattista, N.B., and
Pappano, N.B. (2008) Folia Microbiol.,
53, 23–28.
Sherif, B., Abdel, G., Louise, W.,
Zidan, H.Z., Hussein, M.A.,
Keevil, C.W., and Richard, C.D.B.
(2008) Bioorg. Med. Chem. Lett., 18,
518–522.
(a) Masuda, M., Itoh, K., Murata, K.,
Naruto, S., Uwaya, A., Isami, F., and
Matsuda, H. (2012) Biol. Pharm.
Bull., 35, 78–83; (b) Akihisa, T.,
Seino, K., Kaneko, E., Watanabe, K.,
Tochizawa, S., Fukatsu, M., Banno, N.,
Metori, K., and Kimura, Y. (2010) J.
Oleo Sci., 59, 49–57; (c) Masuda, M.,
Murata, K., Naruto, S., Uwaya, A.,
Isami, F., and Matsuda, H. (2012)
90.
91.
92.
93.
94.
95.
96.
97.
98.
Biol. Pharm. Bull., 35, 210–215;
(d) Saracoglu, I. and Harput, U.S.
(2012) Phytother. Res., 26, 148–152;
(e) Saracoglu, I., Oztunca, F.H.,
Nagatsu, A., and Harput, U.S. (2011)
Pharm. Biol., 49, 1150–1157.
Háznagy-Radnai, E., Balogh, A.,
Czigle, S., Máthé, I., Hohmann, J.,
and Blazsó, G. (2012) Phytother. Res.,
26, 505–509.
Masuda, M., Murata, K., Fukuhama, A.,
Naruto, S., Fujita, T., Uwaya, A.,
Isami, F., and Matsuda, H. (2009) J.
Nat. Med., 63, 267–273.
Tundis, R., Loizzo, M.R., Menichini, F.,
Statti, G.A., and Menichini, F. (2008)
Mini Rev. Med. Chem., 8, 399–420.
Yang, X.P., Yuan, C.S., and Jia, Z.J.
(2006) Chin. Chem. Lett., 17, 337–340.
Stavri, M., Mathew, K.T., and
Gibbons, S. (2006) Phytochemistry,
67, 1530–1533.
Saleem, M., Kim, H.J., Ali, M.S., and
Lee, Y.S. (2005) Nat. Prod. Rep., 22,
696–716.
(a) Hirano, T., Gotoh, M., and
Oka, K. (1994) Life Sci., 55, 1061;
(b) Thompson, L.U., Seidl, M.M.,
Rickard, S.E., Orcheson, L.J., and
Fong, H.H.S. (1996) Nutr. Cancer, 26,
159–165; (c) Kangas, L., Saarinen, N.,
Mutanen, M., Ahotupa, M.,
Hirsinummi, R., Unkila, M., Perala, M.,
Soininen, P., Laatikainen, R., Korte, H.,
and Santti, R. (2002) Eur. J. Cancer
Prev., 11, S48; (d) Lu, H. and Liu, G.T.
(1992) Planta Med., 58, 311–313;
(e) Ghisalberti, E.L. (1997) Phytomedicine, 4, 151–166. (f ) Kitts, D.D.,
Yuan, Y.V., Wijewickreme, A.N.,
and Thompson, L.U. (1999) Mol.
Cell. Biochem., 202, 91–100; (g)
Yamauchi, S., Ina, T., Kirikihira, T.,
and Masuda, T. (2004) Biosci. Biotechnol., Biochem., 68, 183–192; (h)
Charlton, J.L. (1998) J. Nat. Prod.,
61, 1447–1451.
Peng, S.C., Cheng, C.Y., Sheu, F., and
Su, C.H. (2008) J. Appl. Microbiol., 105,
485–491.
Lee, D.G., Jung, H.J., and Woo, E.R. (2005) Arch. Pharm. Res., 28,
1031–1036.
287
288
9 Natural Products as Antimicrobial Agents – an Update
99. Ge, H.M., Huang, B., Tan, S.H.,
100.
101.
102.
103.
104.
105.
106.
107.
108.
109.
110.
111.
112.
Shi, D.H., Song, Y.C., and Tan, R.X.
(2006) J. Nat. Prod., 69, 1800–1802.
Becker, H., Scher, J.M., Speakman, J.-B.,
and Zapp, J. (2005) Fitoterapia, 76,
580–584.
Puupponen-Pimiä, R.,
Nohynek, L., Alakomi, H.L., and
Oksman-Caldentey, K.M. (2005) Appl.
Microbiol. Biotechnol., 67, 8–18.
Cavanagh, H.M.A., Hipwell, M., and
Wilkinson, J.M. (2003) J. Med. Food, 6,
57–61.
Chung, K.M., Wong, T.Y., Wei, C.I.,
Huang, Y.W., and Lin, Y. (1998) Crit.
Rev. Food Sci. Nutr., 38, 421–464.
Pessini, G.L., Dias Filho, B.P.,
Nakamura, C.V., and Cortez, D.A.G.
(2003) Mem. Inst. Oswaldo Cruz, Rio de
Janeiro, 98, 1115–1120.
Koroishi, A.M., Foss, S.R., Cortez, D.A.,
Ueda-Nakamura, T., Nakamura, C.V.,
and Dias Filho, B.P. (2008)
J. Ethnopharmacol., 117, 270–277.
Rukachaisirikul, T., Innok, P.,
Aroonrerk, N., Boonamnuaylap, W.,
Limrangsun, S., Boonyon, C.,
Woonjina, U., and Suksamrarn, A.
(2007) J. Ethnopharmacol., 110,
171–175.
Mbaveng, A.T., Ngameni, B., Kuete, V.,
Simo, I.K., Ambassa, P., Roy, R.,
Bezabih, M., Etoa, F.X., Ngadjui, B.T.,
Abegaz, B.M., Meyer, J.I., Lall, N., and
Beng, V.P. (2008) J. Ethnopharmacol.,
116, 483–489.
Rai, D., Singh, J.K., Roy, N., and
Panda, D. (2008) Biochem. J., 410,
147–155.
Bharate, S.B., Khan, S.I., Yunus, N.A.,
Chauthe, S.K., Jacob, M.R.,
Tekwani, B.L., Khan, I.A., and
Singh, I.P. (2007) Bioorg. Med. Chem.,
15, 87–96.
Kuete, V., Wabo, G.F., Ngameni, B.,
Mbaveng, A.T., Metuno, R., Etoa, F.X.,
Ngadjui, B.T., Beng, V.P., Meyer, J.I.,
and Lall, N. (2007) J. Ethnopharmacol.,
114, 54–60.
Yin, S., Fan, C.-Q., Dong, L., and
Yue, J.-M. (2006) Tetrahedron, 62,
2569–2575.
Thikomkulchai, S.A., Prawat, H.,
Thasana, N., Ruangrungsi, N., and
113.
114.
115.
116.
117.
118.
119.
120.
121.
122.
123.
124.
125.
126.
Ruchirawat, S. (2006) Chem. Pharm.
Bull., 54, 262–264.
Penna, C., Marino, S., Vivot, E.,
Cruan∼es, M.C., de Mun∼oz, J.D.,
Cruan∼es, J., Ferraro, G., Gutkind, G.,
and Martino, V. (2001) J. Ethnopharmacol., 77, 37–40.
Cho, J.Y. and Kim, M.S. (2012) Fish Sci.,
78, 1065–1073.
Petit, P., Lucas, E.M.F., Abreu, L.M.,
Pfenning, L.H., and Takahashi, J.A.
(2009) Electron. J. Biotechnol., 12, SP
1–9.
Jin, W. and Zjawiony, J.K. (2006) J. Nat.
Prod., 69, 704–706.
Dai, J., Krohn, K., Flörke, U.,
Draeger, S., Schulz, B., Kiss-Szikszai, A.,
Antus, S., Kurtán, T., and Ree, T.V.
(2006) Eur. J. Org. Chem., 2006,
3498–3506.
Reátegui, R.F., Wicklow, D.T., and
Gloer, J.B. (2006) J. Nat. Prod., 69,
113–117.
Plaza, A., Keffer, J.L., Bifulco, G.,
Lloyd, J.R., and Bewley, C.A. (2010)
J. Am. Chem. Soc., 132, 9069–9077.
Debbie, Y. and Erik, G. (2007) Use
of polypeptides having antimicrobial activity. US Patent IPC8 Class
AA61K3816FI, US USPC Class: 514 12,
New York.
Fukuda, T., Arai, M., Yamaguchi, Y.,
Masuma, R., Tomoda, H., and
Omura, S. (2004) J. Antibiot., 57,
110–116.
Ciciliato, I., Corti, E., Sarubbi, E.,
Stefanelli, S., Gastaldo, L.,
Montanini, N., Kurz, M., Losi, D.,
Marinelli, F., and Selva, E. (2004) J.
Antibiot., 57, 210–217.
Hashizume, H., Igarashi, M., Hattori, S.,
Hori, M., Hamada, M., and Takeuchi, T.
(2001) J. Antibiot., 54, 1054–1059.
Hino, M., Fujie, A., Iwamoto, T.,
Hori, Y., Hashimoto, M., Tsurumi, Y.,
Sakamoto, K., Takase, S., and
Hashimoto, S. (2001) J. Ind. Microbiol.
Biotechnol., 27, 157–162.
Jeremy, H.L. and Lea, E.J.A. (1986)
Biochim. Biophys. Acta, 859, 219–226.
Daniel, R.S., Rosenthal, K.S., and
Swanson, P.E. (1977) Annu. Rev.
Biochem., 46, 723–763.
References
127. Shoji, J., Hinoo, H., Katayama, T.,
128.
129.
130.
131.
132.
133.
134.
135.
136.
137.
138.
139.
140.
Matsumoto, K., Tanimoto, T.,
Hattori, T., Higashiyama, I., Miwa, H.,
Motokawa, K., and Yoshida, T. (1992) J.
Antibiot., 45, 817–823.
Shoji, J., Hinoo, H., Katayama, T.,
Nakagawa, Y., Ikenishi, Y., Iwatani, K.,
and Yoshida, T. (1992) J. Antibiot., 45,
824–831.
Hashizume, H., Hattori, S., Igarashi, M.,
and Akamatsu, Y. (2004) J. Antibiot.,
57, 394–399.
Neuhof, T., Schmieder, P., Seibold, M.,
Preussel, K., and von Dohren, H.
(2006) Bioorg. Med. Chem. Lett., 16,
4220–4222.
Neuhof, T., Schmieder, P., Preussel, K.,
Dieckmann, R., Pham, H., Bartl, F., and
von Dohren, H. (2005) J. Nat. Prod., 68,
695–700.
(a) Khosla, C., Gokhale, R.S.,
Jacobsen, J.R., and Cane, D.E. (1999)
Annu. Rev. Biochem., 68, 219–253; (b)
Cox, R.J. (2000) Annu. Rep. Prog. Chem.
Sect. B, 96, 231–258.
Mazzei, T., Mini, E., Novelli, A.,
and Perti, P. (1993) J. Antimicrob.
Chemother., 31, 1–9.
Bearden, D.T. and Rodvold, K.A. (1999)
Infect. Med., 16, 480A–484A.
(a) Jaruchoktaweechai, C. (2000) J.
Nat. Prod., 63, 984–986; (b) Nagao, T.
(2001) J. Antibiot., 54, 333–339.
Du, X., Lu, C., Li, Y., Zheng, Z., Su, W.,
and Shen, Y. (2008) J. Antibiot., 61,
250–253.
Yoo, J.-S., Zheng, C.-J., Lee, S.,
Kwak, J.-H., and Kim, W.-G. (2006)
Bioorg. Med. Chem. Lett., 16,
4889–4892.
Perez-zuniga, F.J., Seco, E.M.,
Cuesta, T., Degenhardt, F., Rohr, J.,
Vallin, C., Iznaga, Y., Perez, E.M.,
Gonzalez, L., and Malpartida, F. (2004)
J. Antibiot., 57, 197–204.
Elshahawi, S.I., Trindade-Silva, A.E.,
Hanora, A., Han, A.W., Flores, M.S.,
Vizzoni, V., Schrago, C.G., Soares, C.A.,
Concepcion, G.P., Distel, D.L.,
Schmidt, E.W., and Haygood, M.G.
(2013) Proc. Natl. Acad. Sci. U.S.A.
Plus, 110, E295–E304.
Malkina, N.D. et al. (1994) J. Antibiot.,
47, 342–348.
141. Omura, S., Iwai, Y., Hinotozawa, K.,
142.
143.
144.
145.
146.
147.
148.
149.
150.
151.
152.
153.
Tanaka, H., Takahashi, Y., and
Nakagawa, A. (1982) J. Antibiot., 35,
1425–1429.
Townsend, D.E., Ashdown, N.,
Bolton, S., Bradley, J., Duckworth, G.,
Moorhouse, E., and Grubb, W.B. (1987)
J. Hosp. Infect., 9, 60–71.
Kuete, V., Eyong, K.O., Folefoc, G.N.,
Beng, V.P., Hussain, H., Krohn, K., and
Nkengfack, A.E. (2007) Pharmazie, 62,
552–556.
Ndi, C.P., Semple, S.J., Griesser, H.J.,
Pyke, S.M., and Barton, M.D. (2007)
Phytochemistry, 68, 2684–2690.
Cai, L., Wei, G.-X., Van, B.P., and der
Wu, C.D. (2000) J. Agric. Food Chem.,
48, 909–914.
Singh, D.N., Verma, N.,
Raghuwanshi, S., Shukla, P.K., and
Kulshreshtha, D.K. (2006) Bioorg. Med.
Chem. Lett., 16, 4512–4514.
Eyong, K.O., Folefoc, G.N., Kuete, V.,
Beng, V.P., Krohn, K., Hussain, H.,
Nkengfack, A.E., Saeftel, M.,
Sarite, S.R., and Hoerauf, A. (2006)
Phytochemistry, 67, 605–609.
Almeida, L.S., Murata, R.M.,
Yatsuda, R., Dos Santos, M.H.,
Nagem, T.J., Alencar, S.M., Koo, H.,
and Rosalen, P.L. (2008) Phytomedicine,
15, 886–891.
Panthong, K., Pongcharoen, W.,
Phongpaichit, S., and Walter, C.T.
(2006) Phytochemistry, 67, 999–1004.
Boonsri, S., Karalai, C.,
Ponglimanont, C., Kanjana-opas, A.,
and Chantrapromma, K. (2006) Phytochemistry, 67, 723–727.
Zhang, C., Ondeyka, J.G., Zink, D.L.,
Basilio, A., Vicente, F., Collado, J.,
Platas, G., Huber, J., Dorso, K.,
Motyl, M., Byrne, K., and Singh, S.B.
(2009) Bioorg. Med. Chem., 17,
2162–2166.
Riedlinger, J., Schrey, S.D.,
Tarkka, M.T., Hampp, R., Kapur, M.,
and Fiedler, H.-P. (2006) Appl. Environ.
Microbiol., 72, 3550–3557.
Socha, A.M., Garcia, D., Sheffer, R., and
Rowley, D.C. (2006) J. Nat. Prod., 69,
1070–1073.
289
290
9 Natural Products as Antimicrobial Agents – an Update
154. Zitouni, A., Boudjella, H., Mathieu, F.,
155.
156.
157.
158.
159.
160.
161.
162.
163.
164.
165.
Sabaou, N., and Lebrihi, A. (2004)
J. Antibiot., 57, 367–372.
Shang, Z., Li, X.-M., Li, C.-S., and
Wang, B.-G. (2012) Chem. Biodivers., 9,
1338–1348.
Venkata Dasu, V. and Panda, T. (1999)
Bioprocess Biosyst. Eng., 21, 489.
(a) Mustafa, A. (1974) Benzofurans,
Chemistry of Heterocyclic Compounds,
1st edn, Vol. 29, John Wiley & Sons,
Inc, New York, p. 420; (b) Rahman, A.
(2005) Studies in Natural Products
Chemistry: Bioactive Natural Products (Part L), 1st edn, Vol. 32, Gulf
Professional Publishing, p. 478.
Siddiqui, I.N., Zahoor, A., Hussain, H.,
Ahmed, I., Ahmad, V.U., Padula, D.,
Draeger, S., Schulz, B., Meier, K.,
Steinert, M., Kurt, T., Florke, U.,
Pescitelli, G., and Krohn, K. (2011)
J. Nat. Prod., 74, 365–373.
Young, K., Jayasuriya, H.,
Ondeyka, J.G., Herath, K., Zhang, C.,
Kodali, S., Galgoci, A., Painter, R.,
Brown-Driver, V., Yamamoto, R.,
Silver, L.L., Zheng, Y., Ventura, J.I.,
Sigmund, J., Ha, S., Basilio, A.,
Vicente, F., Tormo, J.R., Pelaez, F.,
Youngman, P., Cully, D., Barrett, J.F.,
Schmatz, D., Singh, S.B., and Wang, J.
(2006) Antimicrob. Agents Chemother.,
50, 519–526.
Singh, S.B., Phillips, J.W., and Wang, J.
(2007) Curr. Opin. Drug Discovery Dev.,
10, 160–166.
Zhang, C., Ondeyka, J.G., Zink, D.L.,
Basilio, A., Vicente, F., Collado, J.,
Platas, G., Huber, J., Dorso, K.,
Motyl, M., Byrne, K., and Singh, S.B.
(2008) J. Nat. Prod., 71, 1304–1317.
Kharwar, R.N., Verma, V.C., Kumar, A.,
Gond, S.K., Harper, J.K., Hess, W.M.,
Lobkovosky, E., Ma, C., Ren, Y., and
Strobel, G.A. (2009) Curr. Microbiol.,
58, 233–238.
Li, Y., Li, X., Lee, U., Kang, J.S.,
Choi, H.D., and Son, B.W. (2006) Chem.
Pharm. Bull., 54, 882–883.
Jiao, P., Swenson, D.C., Gloer, J.B.,
Campbell, J., and Shearer, C.A. (2006)
J. Nat. Prod., 69, 1667–1671.
Ivanova, V.B., Hoshino, Y., Yazawa, K.,
Ando, A., Mikami, Y., Zaki, S.M.,
166.
167.
168.
169.
170.
171.
172.
173.
174.
175.
176.
177.
178.
and Udo, G. (2002) J. Antibiot., 55,
914–918.
Stern, J.L., Hagerman, A.E.,
Steinberg, P.D., and Mason, P.K. (1996)
J. Chem. Ecol., 22, 1887–1899.
(a) Shin, S.Y., Bajpai, V.K., Kim, H.R.,
and Kang, S.C. (2007) Int. J. Food
Microbiol., 11, 3233–3236; (b)
Mundt, S., Kreitlow, S., and Jansen, R.
(2003) J. Appl. Phycol., 15, 263–267; (c)
Ouattara, B., Simard, R.E., Holley, R.A.,
Piette, G.J.P., and Begin, A. (1997) Int.
J. Food Microbiol., 37, 155–162.
Mondol, M.A.M., Kim, J.H., Lee, M.A.,
Tareq, F.S., Lee, H.-S., Lee, Y.-J., and
Shin, H.J. (2011) J. Nat. Prod., 74,
1606–1612.
Zheng, C.J., Yoo, J.-S., Lee, T.-G.,
Cho, H.-Y., Kim, Y.-H., and Kim, W.G.
(2005) FEBS Lett., 579, 5157–5162.
Jayasuriy, H., Zink, D.L., Basilio, A.,
Vicente, F., Collado, J., Bills, G.,
Goldman, M.L., Motyl, M., Huber, J.,
Dezeny, G., Byrne, K., and Singh, S.B.
(2009) J. Antibiot., 62, 265–269.
Singh, S.B., Zink, D.L., Dorso, K.,
Motyl, M., Salazar, O., Basilio, A.,
Vicente, F., Byrne, K.M., Ha, S., and
Genilloud, O. (2009) J. Nat. Prod., 72,
345–352.
Reiss, F. (1947) J. Invest. Dermatol., 8,
245–250.
Reiss, F. (1947) J. Invest. Dermatol., 8,
251–253.
Okwu, D.E. and Igara, E.C. (2009) Afr.
J. Pharm. Pharmacol., 3, 277–281.
Mskhiladze, L., Kutchukhidze, J.,
Chincharadze, D., Delmas, F., Elias, R.,
and Favel, A. (2008) Georgian Med.
News, 154, 39–43.
Renault, S., De Lucca, A.J., Boue, S.,
Bland, J.M., Vigo, C.B., and
Selitrennikoff, C.P. (2003) Med. Mycol.,
41, 75–81.
Zhang, J.-D., Xu, Z., Cao, Y.-B.,
Chen, H.-S., Yan, L., An, M.-M.,
Gao, P.-H., Wang, Y., Jia, X.-M., and
Jiang, Y.-Y. (2006) J. Ethnopharmacol.,
103, 76–84.
Zhang, Y., Li, H.-Z., Zhan, Y.-J.,
Jacob, M.R., Khan, S.I., Li, X.-C.,
and Yang, C.-R. (2006) Steroids, 71,
712–719.
References
179. Zhao, C.-C., Shao, J.-H., Li, X., Xu, J.,
180.
181.
182.
183.
184.
185.
186.
187.
188.
189.
190.
and Zhang, P. (2005) Arch. Pharm. Res.,
28, 1147–1151.
Vicente, M.F., Cabello, A., Platas, G.,
Basilio, A., DõÂez, M.T., Dreikorn, S.,
Giacobbe, R.A., Onishi, J.C.,
Meinz, M., Kurtz, M.B., Rosenbach, M.,
Thompson, J., Abruzzo, G., Flattery, A.,
Kong, L., Tsipouras, A., Wilson, K.E.,
and PelaÂez, F. (2001) J. Appl. Microbiol., 91, 806–813.
Chaurasia, S.C. and Vyas, K.K. (1977)
J. Res. Indian Med., Yoga Homeopathy,
1977, 24–26.
An, R.-B., Kim, H.-C., Lee, S.-H.,
Jeong, G.-S., Sohn, D.-H., Park, H.,
Kwon, D.-Y., Lee, J.H., and Kim, Y.-C.
(2006) Arch. Pharm. Res., 29, 815–820.
Lee, L.Y., Shim, J.S., Rukayadi, Y., and
Hwang, J.K. (2008) J. Food Prot., 71,
1926–1930.
Theodori, R., Karioti, A., Rancić, A.,
and Skaltsa, H. (2006) J. Nat. Prod., 69,
662–664.
Berry, Y., Jamie, J.F., and Bremner, J.B.
(2000) Phytochemistry, 55, 403–406.
Amiguet, V.T., Petit, P., Ta, C.A.,
Nun∼ez, R., Sańchez-Vindas, P.,
Alvarez, L.P., Smith, M.L., Arnason, J.T.,
and Durst, T. (2006) J. Nat. Prod., 69,
1005–1009.
Rabe, T., Mullholland, D., and van
Staden, J. (2002) J. Ethnopharmacol.,
80, 91–94.
More, G., Tshikalange, T.E., Lall, N.,
Botha, F., and Meyer, J.J.M. (2008) J.
Ethnopharmacol., 119, 473–477.
Chung, J.Y., Choo, J.H., Lee, M.H., and
Hwang, J.K. (2006) Phytomedicine, 13,
261–266.
(a) Ambrosio, S.R., Furtado, N.A.J.C.,
De Oliveira, D.C.R., Da Costa, F.B.,
Martins, C.H.G., De Carvalho, T.C.,
Porto, T.S., and Veneziani, R.C.S.
(2008) Z. Naturforsch., C, 63,
326–330; (b) Porto, T.S., Rangel, R.,
Furtado, N.A.J.C., De Carvalho, T.C.,
Martins, C.H.G., Veneziani, R.C.S.,
Da Costa, F.B., Vinholis, A.H.C.,
Cunha, W.R., Heleno, V.C.G., and
Ambrosio, S.R. (2009) Molecules, 14,
191–199; (c) Koo, H., Rosalen, P.L.,
Cury, J.A., Park, Y.K., and Bowen, W.H.
(2002) Antimicrob. Agents Chemother.,
191.
192.
193.
194.
195.
196.
197.
198.
199.
200.
201.
202.
46, 1302–1309; (d) Tomczyk, M.,
Pleszczynska, M., and Wiater, A. (2010)
Molecules, 15, 4639–4651.
Teponno, R.B., Tapondjou, A.L.,
Gatsing, D., Djoukeng, J.D.,
Abou-Mansour, E., Tabacchi, R.,
Tane, P., Stoekli-Evans, H., and
Lontsi, D. (2006) Phytochemistry, 67,
1957–1963.
Ghisalberti, E.L. (1997) Fitoterapia, 68,
303–325.
Slimestad, R., Marston, A., Mavi, S.,
and Hostettmann, K. (1995) Planta
Med., 61, 562–563.
Shi, Y., Yu, B., Williams, I.D.,
Sung, H.H., Zhang, Q., Liang, J.Y.,
Ip, N.Y., and Min, Z.D. (2007) Planta
Med., 73, 84–90.
Souza, A.B., Martins, C.H.G.,
Souza, M.G.M., Furtado, N.A.J.C.,
Heleno, V.C.G., de Sousa, J.P.B.,
Rocha, E.M.P., Bastos, J.K.,
Cunha, W.R., Veneziani, R.C.S., and
Ambrósio, S.R. (2011) Phytother. Res.,
25, 215–220.
Urzúa, A., Rezende, M.C.,
Mascayano, C., and Vásques, L. (2008)
Molecules, 13, 882–891.
da Silva, L.L., Nascimento, M.S.,
Cavalheiro, A.J., Silva, D.H.,
Castro-Gamboa, I., Furlan, M., and
Bolzani, V.S. (2008) J. Nat. Prod., 71,
1291–1293.
Cavin, A.-L., Hay, A.-E., Marston, A.,
Stoeckli-Evans, H., Scopelliti, R.,
Diallo, D., and Hostettmann, K. (2006)
J. Nat. Prod., 69, 768–773.
Das, B., Ramu, R., Venkateswarlu, K.,
Rao, Y.K., Reddy, M.R.,
Ramakrishna, K.V.S., Harakishore, K.,
and Murty, U.S. (2006) Chem. Biodivers., 175–179.
Liu, X.-T., Pan, Q., Shi, Y.,
Williams, I.D., Sung, H.H.-Y., Zhang, Q.,
Liang, J.-Y., Ip, N.Y., and Min, Z.-D.
(2006) J. Nat. Prod., 69, 255–260.
Wellsow, J., Grayer, R.J., Veitch, N.C.,
Kokubun, T., Lelli, R., Kite, G.C., and
Simmonds, M.S.J. (2006) Phytochemistry, 67, 1818–1825.
Kofujita, H., Fujino, Y., Ota, M., and
Takahashi, K. (2006) Holzforschung, 60,
20–23.
291
292
9 Natural Products as Antimicrobial Agents – an Update
203. Lee, S., Shin, D.-S., Oh, K.-B., and
204.
205.
206.
207.
208.
209.
210.
211.
212.
213.
214.
215.
Shin, K.H. (2003) Arch. Pharm. Res.,
26, 40–42.
Cardoso Sá, N., Cavalcante, T.T.A.,
Araujo, A.X., dos Santos, H.S.,
Albuquerque, M.R.J.R., Bandeira, P.N.,
da Cunha, R.M.S., Cavada, B.S., and
Teixeira, E.H. (2012) Arch. Oral Biol.,
57, 550–555.
Saleem, M., Nazir, M., Ali, M.S.,
Hussain, H., Lee, Y.S., Riaz, N., and
Jabbar, A. (2010) Nat. Prod. Rep., 27,
238–254.
Chiozem, D.D., Trinh-Van-Dufat, H.,
Wansi, J.D., Djama, C.M.,
Fannang, V.S., Seguin, E., Tillequin, F.,
and Wandji, J. (2009) Chem. Pharm.
Bull., 57, 1119–1122.
Awanchiri, S.S., Trinh-Van-Dufat, H.,
Shirri, J.C., Dongfack, M.D.J.,
Nguenang, G.M., Boutefnouchet, S.,
Fomum, Z.T., Seguin, E., Verite, P.,
Tillequin, F., and Wandji, J. (2009)
Phytochemistry, 70, 419–522.
Severiano, M.E., Simao, M.R.,
Porto, T.S., Martins, C.H.G.,
Veneziani, R.C.S., Furtado, N.A.J.C.,
Arakawa, N.S., Said, S., de
Oliveira, D.C.R., Cunha, W.R.,
Gregorio, L.E., and Ambrosio, S.R.
(2010) Molecules, 15, 8553–8566.
Choudhary, M.I., Siddiqui, Z.A.,
Hussain, S., and Rahman, A. (2006)
Chem. Biodivers., 3, 54–61.
Evans, L., Hedger, J.N., Brayford, D.,
Stavri, M., Smith, E., O’Donnell, G.,
Gray, A., Grith, G.W., and Gibbons, S.
(2006) Phytochemistry, 67, 2110–2114.
Vicente, F., Basilio, A., Platas, G.,
Collado, J., Bills, G.F., Del Val, A.G.,
Martin, J., Tormo, J., Harris, G.H.,
Zink, D.L., Justice, M., Kahn, J.N.,
and Pelaez, F. (2009) Mycol. Res., 113,
754–770.
Vairappan, C.S. (2003) Biomol. Eng., 20,
255–259.
Tsukamoto, S., Yamashita, Y., and
Ohta, T. (2005) Mar. Drugs, 3, 22–28.
Bishara, A., Rudi, A., Goldberg, I.,
Benayahu, Y., and Kashman, Y. (2006)
Tetrahedron, 62, 12092–12097.
Williams, R.B., Martin, S.M., Hu, J.-F.,
Norman, V.L., Goering, M.G., Loss, S.,
O’Neil-Johnson, M., Eldridge, G.R., and
216.
217.
218.
219.
220.
221.
222.
223.
224.
225.
226.
227.
228.
229.
Starks, C.M. (2012) J. Nat. Prod., 75,
1319–1325.
Dong, L.-P., Liu, H.-Y., Nia, W.,
Li, J.-Z., and Chen, C.-X. (2006) Chem.
Biodivers., 3, 791–798.
Momin, R.A. and Nair, M.G. (2001)
J. Agric. Food Chem., 49, 142–145.
Tani, H., Koshino, H., Sakuno, E.,
Cutler, H.G., and Nakajima, H. (2006)
J. Nat. Prod., 69, 722–725.
Heilmann, J., Brun, R., Mayr, S.,
Rali, T., and Sticher, O. (2001) Phytochemistry, 57, 1281–1285.
Quang, D.N., Stadler, M., Fournier, J.,
and Asakawa, Y. (2006) J. Nat. Prod.,
69, 1198–1202.
Teles, H.L., Sordi, R., Silva, G.H.,
Castro-Gamboa, I., Bolzani, V.d.S.,
Pfenning, L.H., de Abreu, L.M.,
Costa-Neto, C.M., Young, M.C.M.,
and Araujo, A.R. (2006) Phytochemistry,
67, 2686–2690.
Oueslati, M.H., Jannet, H.B., Mighri, Z.,
Chriaa, J., and Abreu, P.M. (2006) J.
Nat. Prod., 69, 1366–1369.
Perianayagam, J.B., Sharma, S.K.,
Pillai, K.K., Pandurangan, A., and
Kesavan, D. (2012) J. Ethnopharmacol.,
142, 283–286.
Trippe, K., McPhail, K., Armstrong, D.,
Azevedo, M., and Banowetz, G. (2013)
BMC Microbiol., 13, 111.
Katagiri, K., Tori, K., Kimura, Y.,
Yoshida, T., Nagasaki, T., and
Minato, H. (1967) J. Med. Chem.,
10, 1149–1154.
Chen, C., Wang, J., Guo, H., Hou, W.,
Yang, N., Ren, B., Liu, M., Dai, H.,
Liu, X., Song, F., and Zhang, L.
(2013) Appl. Microbiol. Biotechnol.,
97, 3885–3892.
Tadesse, M., Strom, M.B., Svenson, J.,
Jaspars, M., Milne, B.F., Torfoss, V.,
Andersen, J.H., Hansen, E.,
Stensvag, K., and Haug, T. (2010)
Org. Lett., 12, 4752–4755.
Kanokmedhakul, S.,
Kanokmedhakul, K., Nasomjai, P.,
Louangsysouphanh, S., Soytong, K.,
Isobe, M., Kongsaeree, P., Prabpai, S.,
and Suksamrarnr, A. (2006) J. Nat.
Prod., 69, 891–895.
Mierau, V., Sterner, O., and Anke, T.
(2004) J. Antibiot., 57, 311–315.
References
230. Mudur, S.V., Swenson, D.C., Gloer, J.B.,
234. Saleem, M., Hussain, H., Ahmed, I.,
Campbell, J., and Shearer, C.A. (2006)
Org. Lett., 8, 3191–3194.
231. Thomas, L.V., Yu, S., Ingram, R.E.,
Refdahl, C., Elsser, D., and
Delves-Broughton, J. (2002) J. Appl.
Microbiol., 93, 697–705.
232. Hwang, B.K., Lim, S.W., Kim, B.S.,
Lee, J.Y., and Moon, S.S. (2001) Appl.
Environ. Micobiol., 67, 3739–3745.
233. Lavermicocca, P., Valerio, F.,
Evidente, A., Lazzaroni, S., Corsetti, A.,
and Gobbetti, M. (2000) Appl. Environ.
Micobiol., 66, 4084–4090.
van Ree, T., and Krohn, K. (2011) Nat.
Prod. Rep., 28, 1534–1579.
235. Singh, S.B., Jayasuriya, H.,
Ondeyka, J.G., Herath, K.B., Zhang, C.,
Zink, D.L., Tsou, N.N., Ball, R.G.,
Basilio, A., Genilloud, O., Diez, M.T.,
Vicente, F., Pelaez, F., Young, K., and
Wang, J. (2006) J. Am. Chem. Soc., 128,
11916–11920.
236. Cragg, G.M., Newman, D.J., and
Snader, K.M. (1997) Nat. Prod. Drug
Discovery Dev., 60, 52–60.
293
295
10
Photodynamic Antimicrobial Chemotherapy
David A. Phoenix, Sarah R. Dennison, and Frederick Harris
10.1
Introduction
Infectious diseases remain the leading cause of mortality on a global scale, primarily because of the emergence of microbial pathogens with multiple-drug resistance
(MDR) [1–5], and are increasingly being recognized as life-threatening, public
health issues in countries across the world [6–10], including the United Kingdom
[11]. This situation has led to a series of recommendations by the World Health
Organization [12], and based on these recommendations, a number of strategies
to minimize the threat posed by MDR pathogens have been developed [13] such
as stewardship programs and infection control regimes [14–19]. However, while
these strategies may contain and reduce this threat, its ongoing nature necessitates an urgent search for new antimicrobial agents with novel mechanisms of
action to supplement the generally dwindling supply of conventional antimicrobial drugs currently available [20, 21]. In response, there have been concerted
attempts to develop such agents, as so well described in the other chapters of this
book, along with renewed interest in strategies such as photodynamic antimicrobial chemotherapy (PACT) that have fallen into relative disuse [22, 23]. PACT is a
minimally invasive modality that involves the utilization of photosensitizers (PSs),
which are molecules that can be selectively taken up by microbes and activated by
subsequent exposure to visible light to eradicate the host microbe via the generation of reactive cytotoxic oxygen products [24]. Introduced in the 1930s, PACT
was relinquished in favor of antibiotic therapy around the middle of the last century [25], but given the current dearth of antimicrobials [20, 21], this modality
has re-emerged to become a strong, potential alternative or adjunct therapy for
the treatment of infections that are difficult to treat with conventional drugs and
regimes [24, 26–29]. This modality involves the selective targeting of microbes
by molecules such as methylene blue (MB), which is a heteroaromatic, photoactive dye (Figure 10.1) [30, 31], and the archetypical PS used in PACT [30, 32–39].
In addition to these PACT regimes, an alternate form of this therapy is that in
which 5-aminolevulinic acid (ALA; Figure 10.1) is selectively targeted at microbes
Novel Antimicrobial Agents and Strategies, First Edition.
Edited by David A. Phoenix, Frederick Harris and Sarah R. Dennison.
© 2015 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2015 by Wiley-VCH Verlag GmbH & Co. KGaA.
296
10
Photodynamic Antimicrobial Chemotherapy
N
H3C
S
N
CH3
+ CH
3
N
CI− CH3
NH
(a)
N
O
HN
N
OH
H2N
O
(b)
(c)
O
Figure 10.1 The structures of PS used in
PACT. This figure shows the structures of: (a)
methylene blue (MB), (b) 5-aminolevulinic
acid (ALA), and (c) protoporphyrin IX (PPIX).
The core structure of MB is a planar tricyclic
heteroaromatic ring system [31, 41] while
that of PPIX is a planar molecule with four
conjugated pyrrole rings [41, 42]. These
macrocyclic molecules possess π-electron
OH
O
OH
systems that allow for long-lived triplet or
states and thereby the ability of the
parent PS to act as a PACT agent [31, 41,
42]. ALA (c) is effectively the substrate for
the production of PPIX in the heme biosynthetic pathway, which underpins the ability
of PACT based on the introduction of exogenous ALA to induce antimicrobial phototoxicity, as depicted in Figure 10.2 [29].
3 PS*
to induce the accumulation of protoporphyrin IX (PPIX), a macrocyclic, photoactive molecule (Figure 10.1) capable of serving as an endogenous PS within these
microbes [23, 29, 38, 40]. In this chapter, we give an overview of mainstream PACT
using MB as a major example and focus on recent progress in the development of
PACT based on ALA, which is increasingly gaining recognition as a major future
antimicrobial modality in the clinical and biotechnical arenas.
10.2
The Administration and Photoactivation of PS
The ability of ALA administration to initiate a photodynamic response derives
from the fact that this molecule is synthesized as the first intermediate in the heme
biosynthetic pathway, which is highly conserved across organisms and is well characterized in non-plant eukaryotes and non-phototrophic bacteria (Figure 10.2a),
[43, 44]. The heme biosynthetic pathways of these two groups of organisms differ
only in their initial step of ALA synthesis (Figure 10.2b) and in both cases, this synthesis is regulated by negative feedback from heme levels (Figure 10.2a). The last
step of these pathways involves the incorporation of Fe2+ into PPIX to form nonphotodynamically active heme, which is catalyzed by the enzyme ferrochelatase
and is the rate-limiting step of the biosynthetic route [45, 46]. It is this limited
capacity of ferrochelatase to transform PPIX to heme that underlies the ability of
ALA to induce the cell death of microbes directly [47, 48]. It was found that by the
addition of exogenous ALA to target cells, the feedback inhibition of ALA synthesis in the heme biosynthetic pathway was bypassed, which led to the accumulation
10.2
ALA
synthesis
The Administration and Photoactivation of PS
Negative feedback
regulation
δ -Aminolevulinic acid
(ALA)
ALA dehydratase
Porphobilinogen (PBG)
PBG deaminase
Hydroxymethylbilane
Uroporphyrinogen
synthase
Uroporphyrinogen
Uroporphyrinogen
decarboxylase
Coproporphyrinogen III
Coproporphyrinogen
oxidase
Protoporphyrinogen IX
Protoporphyrinogen
oxidase
Protoporphyrin IX (PPIX)
Ferrochelatase
Protoheme (HEME)
(a)
(1) Glutamyl-tRNAGlu
Gluatamate-1semialdehyde (GSA)
GSA
transaminase
Glutamyl-tRNA
reductase
(2)
(b)
ALA
Glycine + succinyl-CoA
ALA
δ -Aminolevulinate
synthase
Figure 10.2 Panel (a) shows the heme
biosynthetic pathway, which is highly conserved across organisms except for the initial step of ALA synthesis. In prokaryotes,
glutamate-1-semialdehyde transferase catalyzes the formation of ALA from glutamate1-semialdehyde (b(1)) while in non-plant
eukaryotes, ALA-synthetase is used to
catalyze the production of ALA from glycine
and succinyl CoA (b(2)). The introduction of
exogenous ALA into the heme biosynthetic
pathway bypasses its negative feedback regulation step and promotes the production of
protoporphyrin IX, a potent PS represented
in Figure 10.1. (This figure was adapted from
Ref. [29].)
297
298
10
Photodynamic Antimicrobial Chemotherapy
of photodynamically active PPIX and other porphyrin intermediates. Subsequent
irradiation by light led to the activation of these endogenous PS, which then served
as in situ photodynamic agents and inactivated the host cell [49]. These observations inspired the use of photodynamic therapy based on ALA for the diagnosis
and treatment of a variety of conditions [38, 50, 51], including cancers [52–55]
and inflammatory diseases of the skin [56–58]. It was first suggested by Kennedy
et al. [59] that this form of photodynamic therapy may also serve as an antimicrobial strategy [59].
It is therapeutically desirable that PS used for PACT application have strong
absorption in the red region of the visible spectrum as biological molecules show
minimal adsorption of light in this region, thus generally allowing deeper light
penetration into human tissues [41, 60, 61]. PPIX [62], MB [30], and most other
PS involved in PACT application are generally activated by red light in the range
of 630–700 nm [63] and a variety of light sources, including lamps, light-emitting
diodes, intense light pulse systems, and lasers, have been used for this purpose
[23, 56, 58, 61, 64–69], primarily for the inactivation of bacteria [70], fungi [71],
viruses [38], and parasites [34, 72]. However, light with wavelengths between 630
and 700 nm is only able to effectively penetrate tissue up to depths of 3 mm [61, 67],
which generally limits PACT to topical application and to accessible areas of the
body such as the skin, nails, oral cavity, and lower female reproductive tract [38,
56, 73]. In response, there have been numerous investigations into optimizing the
topical administration of MB and ALA through the production of functionalized
derivatives and the design of improved delivery methods for these compounds
[29, 30, 74, 75]. A comprehensive discussion of these investigations is beyond the
scope of this chapter, but as examples derivatives of ALA such as its methyl and
hexyl esters have been prepared, which possess higher lipophilicity than the parent
drug and exhibit improved skin penetration, resulting in elevated levels of PPIX
production and microbial death [76–81]. To improve the topical localization of
ALA and its derivatives in PACT, a variety of factors that influence the efficacy
of this administrative route have been recently investigated [82–87] and led to
the development of a number of novel delivery methods, including sprays [88],
microemulsions [89, 90], liposomal encapsulation [91], micro-needle arrays [92,
93], cream formulations with penetration enhancers [86, 94, 95], and pressuresensitive, bioadhesive patches [86, 91, 94–97]. Innovative strategies have also been
reported for optimizing the topical administration of MB and its derivatives [69,
73, 75, 98–101] such as the use of the nebulized PS in conjunction with fiber-optic
light delivery to treat infection in the lungs of cystic fibrosis patients [102] and the
use of MB-loaded liposomes [91, 103] and microspheres [104] to specifically target the pilosebaceous units, which play a major role in the pathogenesis of the skin
disorder, acne vulgaris [105]. More recently, liposomal MB has also been used to
develop a version of PACT that eliminates the requirement for external irradiation by encapsulating a chemiluminescent light source with the dye [106, 107].
Termed chemiluminescence photodynamic antimicrobial therapy, this methodology showed activity against MDR bacterial pathogens that was comparable to that
10.2
The Administration and Photoactivation of PS
of PACT and it was suggested that it has the potential to treat deep-seated microbial infections that may be inaccessible to PACT [106, 108, 109].
It is generally perceived that PS have some selectivity for microbial cells over
healthy mammalian host cells although it is recognized these molecules are effective in killing both cell types [38, 73, 110]. In part, this selectivity derives from
the topical administration of PACT PS as compared to systemic delivery in that
they come directly into contact with the microbes instead of being delivered via
the capillaries and coming into contact with host cells first [111]. A major contributor to the selectivity of MB and most other PS for microbes derives from
the fact that most of these molecules are designed to possess cationic structures,
thereby allowing them to rapidly bind bacterial membranes, which are anionic, in
preference to those of healthy human cells, which are zwitterionic [32, 40, 106].
This difference in membrane binding affinity effectively means that PS are only
slowly taken up by host cells using the process of endocytosis as compared to
the uptake mechanisms of bacterial cells, thus giving these molecules temporal
selectivity for microbes [111, 112]. It is generally accepted that the electrostatic
binding of MB to bacterial membranes facilitates self-promoted pathways of cell
uptake where metal ions involved in stabilizing membrane structure are displaced
by the dye to induce the formation of membrane lesions and the internalization of
MB [113, 114]. In contrast, MB appears to be taken up by fungal cells by crossing
the cytoplasmic membrane of these organisms via mechanisms involving specific
molecular transporters of the major facilitator superfamily [115]. The selectivity
of ALA for microbial cells over healthy host cells is manifested as a more pronounced accumulation of PPIX in these former cells and appears to be based on
a preference for cells with high rates of metabolic activity [29] such as bacterial
cells [116] and virally infected cells [117, 118]. Similarly to MB, specific molecular
transporters appear to facilitate the internalization of ALA into target cells [29],
including those of bacteria [119–121], fungi [122–124], and some parasites [72],
although the uptake of ALA by host cells appears to mediate the inactivation of
other parasites and viruses [29].
A number of physical, pharmacokinetic, and photochemical properties are
desirable for PS used in PACT and a review of these properties is generally
beyond the scope of this chapter, but they are summarized in Table 10.1. The
primary photo-physical processes involved in PACT are similar for MB [30,
32, 114, 125], PPIX [126–128], and the vast majority of PS used in this form of
chemotherapy [24, 61, 129], and a schematic representation of the main steps
involved in these processes is shown in Figure 10.1. In the first step, a PS has a
stable electronic configuration in the singlet ground state with the most energetic
electrons in the highest occupied molecular orbital (1 PS0 ). The PS then absorbs
a photon of light (hv) with wavelength 𝜆 that ideally lies within the therapeutic
window of 600–900 nm (Table 10.1). Having absorbed a photon of light, the PS
undergoes an electronic transition, which promotes an electron from the singlet
state to the lowest unoccupied molecular orbital, causing the PS to reach the
short-lived excited singlet state (1 PS* ) [24, 129]. In this state, the PS may rapidly
lose energy by electronic decay (fluorescence) or physical processes (internal
299
300
10
Photodynamic Antimicrobial Chemotherapy
Table 10.1 Desirable characteristics for PS involved in PACT.
1
2
3
4
5
6
7
8
9
10
11
12
Available in pure form of known chemical composition
Synthesizable from available precursors and easily reproduced
High chemical stability
High singlet oxygen quantum yield (ΦΔ). For example, ΦΔ for MB = 1 [32]
Natural fluorescence
Low photobleaching
Low dark toxicity: not harmful to the target tissue until treatment light application
Strong maximal absorption (𝜆max ) in the window 600–900 nm, where tissue is more
transparent. For example, 𝜆max for MB = 656 nm [32]
High extinction coefficient (𝜀max ) of circa 50 000–100 000 M−1 cm−1 . For example,
for MB, 𝜀max = 95 000 [32]
Stable and soluble in the body’s tissue fluids and easily delivered
Preferential uptake in target tissue
Excreted from the body upon completion of treatment
This table was compiled from Refs. [32, 60, 61].
conversion to heat) and thus return to the singlet ground state, 1 PS0 . However, if
1 PS* is relatively stable, the PS may also undergo an electronic rearrangement in
which the spin of the excited electron is reversed (intersystem crossing) to give
the excited triplet state (3 PS* ). At this stage, the PS may again undergo electronic
decay (phosphorescence) back to the singlet ground state, 1 PS0 . However,
although 3 PS* is less energetic than 1 PS* , it has a relatively longer lifetime – many
microseconds as compared to a few nanoseconds [24, 129], allowing 3 PS* to pass
its excitational energy onto other biomolecules by either of two processes, which
are defined as type I and type II mechanisms and both facilitate photodynamic
action [24, 61] (Figure 10.3).
Type I mechanisms involve direct interaction between 3 PS* and a nearby
biomolecule, which acts as a reductant. Hydrogen abstraction or electron
transfer between 3 PS* and this biomolecule yields free radicals, which can then
initiate further redox reactions leading to cellular damage. In contrast, type II
mechanisms involve energy transfer via the collision of 3 PS* with molecular
oxygen, which yields the ground state PS and singlet oxygen, 1 O2 . When formed
in situ within a cellular environment, singlet oxygen is highly cytotoxic owing
to its strong oxidizing activity. Moreover, it has a long lifetime in relation to
other reactive oxygen species (ROS) and may diffuse from its site of generation
before reacting with a biomolecule, again leading to further reactions and cellular
damage [24, 61, 129]. Nonetheless, given a sufficiently populated triplet state
and the availability of oxygen, a PS has the potential to inflict damage on a
variety of biomolecules via either or both of these mechanisms (Table 10.2) [29,
32, 132–135]. In general, this photodamage (Table 10.2) can include the loss
of membrane functions induced by the peroxidation of membrane cholesterol
[136, 137] and lipids [138, 139], the inactivation of essential enzymes and loss
of membrane protein function owing to bimolecular cross-linking reactions
10.3
Applications of PACT Based on MB
O2
Singlet state
1PS
Intersystem
crossing
O
2−
e−
∗
Fluorescence
Type I
∗
Triplet state
Type II
1 O2
1PS0
2
O
3
Ph
os
ph
ore
sce
nc
e
hν
3PS
Ground state
Figure 10.3 This figure shows a schematic
representation of the possible photochemical/photophysical steps involved in the photosensitization of a PACT agent. A photosensitizer in its ground state, PS, absorbs a
photon of light (hv) and undergoes an electronic transition, which promotes an electron
to the excited singlet state, 1 PS* . The electron can then lose energy by fluorescence
or can undergo an electronic rearrangement
to achieve intersystem crossing to a longlived triplet state, 3 PS* . An electron in the
3 PS* state can lose its excitational energy
by phosphorescence or pass it onto other
molecules by type 1 and type II mechanisms,
which can then lead to the local production of reactive oxygen species such as singlet oxygen (Type II) or superoxide (Type
I) that are cytotoxic to microbial and other
target cells [24, 61, 129]. More recently, the
activation of PS through two-photon processes has been described, which offers several advantages over the single photon process presented above although a discussion
of these processes is beyond the scope of
this chapter [61, 130, 131]. (This figure was
adapted from Ref. [38]. Reproduced with permission. Copyright © 2011 Wiley-Liss, Inc.).
[140–143], and mutagenetic effects resulting from RNA and DNA modification
[144, 145]. A particularly important example of such DNA modification is the
oxidation of guanosine to produce 8-oxo-7,8-dihydroguanosine (8-oxodG), which
is a major product of type II photo-attack on cellular nucleic acids and is often
used as an in vivo biomarker of such attack [125, 146, 147].
10.3
Applications of PACT Based on MB
PACT based on MB is considered to be generally safe for human application [31,
148, 149] and the technique is probably best known for its anticancer activity
301
302
10
Photodynamic Antimicrobial Chemotherapy
Table 10.2 Reaction products from Type I and Type II mechanisms in PACT.
Biological target
Type I
Type II
Amino acids
Tyrosine (phenol coupling),
phenylalinine (hydroxylation),
tryptophan
(N-formylkynurenine)
Cross-linking, carbonylation
Radical chain reactions
5-Carboxamido-5-formamido2-iminohydantoin
5- and 6-Hydroperoxides
Histidine (oxygen addition),
tryptophan (oxygen addition)
cystine (disulfide), methionine
(sulfoxide)
Cross-linking, proteolysis
Lipid hydroperoxides
8-Oxo-7,8-dihydroguanosine
Proteins
Lipids
Nucleic acids
Cholesterol
7-Hydroperoxides
This table was compiled from Refs. [24, 32].
[31, 32, 41, 150–152] although it is an often forgotten fact that MB was the first
synthetic antimicrobial compound to be reported, when in the 1890s the dye successfully treated plasmodium infection in humans [149, 153, 154]. Since this time,
the photodynamic antimicrobial activities of MB and its derivatives have been
widely investigated [31, 32, 39, 155–158] and the dye has been shown to have
activity against bacteria [28, 39, 159–165], viruses [145, 166–170], fungi [33, 40,
115, 171–174], and parasites [72, 175–177]. On the basis of this broad spectrum
activity, PACT based on MB and its derivatives have been developed for a multitude of clinical and biotechnical uses [5, 28, 32, 40, 72, 125, 162, 178, 179] and are
well-established antimicrobials for disinfection in dentistry [180–182] and in the
decontamination of blood [170, 183, 184].
Donated blood is fractionated into red blood cell concentrate, platelet concentrate, and plasma, and the transfusion of blood components is an invaluable part
of modern medicine [185] although a major problem is the presence of pathogens
in donated blood, which can lead to transfusion-transmitted disease [186]. In
response, MB and other phenothiaziniums have been shown to be efficacious
blood decontaminants with activity against a variety of major pathogens found in
blood fractions, particularly viruses [30, 167, 184, 185, 187–191]. For example,
the human immunodeficiency virus and hepatitis viruses are major contaminants
of blood fractions [186] and MB is able to photo-inactivate both these organisms
and other viruses via attack on a number of sites, including nucleic acids, internal
proteins, or lipids, along with the envelope and core proteins [145, 166–170,
187]. Currently, MB-mediated techniques are widely employed by European
transfusion services in the viral photo-decontamination of blood products [167,
184, 187] with a major example provided by the Theraflex-MB system, which is
primarily used for the inactivation of viruses in blood plasma [192, 193]. In the
case of dentistry, oral infections are a major problem and MB has been used to
treat a number of these infections [180–182, 194–198], including lichen planus
[198], caries [199] and in particular, periodontal diseases such as gingivalis and
periodontitis, which are associated with the accumulation of plaque biofilms
10.4
The Applications of PACT Based on ALA
[200, 201]. The major etiological agents associated with periodontal diseases
include Porphyromonas gingivalis, Aggregatibacter actinomycetemcomitans, and
Fusobacterium nucleatum [201] and it has been recently shown that PACT based
on MB and other phenothiaziniums is able to kill a variety of these periopathogens
[181, 202]. Clinical trials have also shown that when this methodology was used
by patients with chronic periodontitis to complement traditional therapies,
the treated individuals experienced significant improvements in the alleviation
of this disorder [202–207]. On the basis of these results, PACT based on MB
was developed and marketed as ‘PeriowaveTM ’ for the treatment of periodontal
diseases and led to the view that photo-disinfection would play a major role in
periodontal therapy [208, 209].
10.4
The Applications of PACT Based on ALA
Similarly to MB, PACT based ALA is generally regarded as safe for use on humans
[29, 210] and the technique is commonly associated with anticancer activity
[52, 54]. In contrast to MB, PACT based on ALA is far less well established as
a mainstream PACT modality although known to have broad spectrum activity
against bacteria viruses, fungi, and parasites [29, 38, 40, 51, 71, 87, 145, 211–215].
However, since the antimicrobial use of ALA based PACT was first suggested
by Kennedy et al. [59], it has become clear that this technique kills microbes
via mechanisms other than through direct uptake and the induction of PPIX as
described for bacteria [216, 217], fungi [71, 218], and some parasites [219, 220].
Some microbes have no capacity for PPIX production [29, 118, 221] and it is
now known that the uptake of ALA and PPIX production by host cells induces
an immune response that mediates the inactivation of a number of parasites
[221–224] and viruses [214, 225]. In yet other cases, ALA-mediated production
of PPIX by host cells appears to inactivate viruses via either the induction of a host
cell response such as apoptosis and necrosis [226, 227] or direct photodynamic
attack on viral particles [29, 228]. On the basis of this activity, PACT based on
ALA has been developed for a wide variety of applications [29, 38, 40, 51, 41, 71,
87, 145, 211–215, 229] but two of the major emerging uses of this technique are
decontamination in the food industry [29, 212] and anti-infective action in the
area of dermatology [29, 51, 57, 221, 230, 231].
10.4.1
Food Decontamination Using PACT Based on ALA
The use of PACT based on ALA for food-related applications has appeal because
ALA is colorless and odorless and thus the use of the technique does not involve
strongly colored agents that could compromise the appearance and taste of food
[212]. In response, there has been recent research into using PACT based on ALA
to inactivate bacterial pathogens that contaminate food [232]. Major examples of
303
304
10
Photodynamic Antimicrobial Chemotherapy
these pathogens include Listeria monocytogenes, whose infections are primarily
associated with raw foods and lead to listeriosis in humans [233] and Salmonella
enterica, which often infects cattle and poultry, thereby causing salmonellosis in
humans [234]. Recently, in vitro studies showed that both S. enterica and L. monocytogenes accumulated PPIX in response to ALA uptake with subsequent irradiation by light inducing their photo-inactivation [232, 235, 236]. Both organisms
have since been shown to be inactivated by PACT based on ALA and theoretical analysis of the survival curves of L. monocytogenes and S. enterica after this
treatment led to the production of mathematical models, which were able to predict the optimal combinations of ALA concentration, incubation time, and light
irradiation period for eradication of these organisms [237]. Most recently, both
organisms were shown to be eradicated not only by PACT based on ALA but
also by PACT based on the exogenous PS, chlorophyllin, although showing differing susceptibilities. Using theoretical analysis similar to that used by the latter
authors, combinations of ALA and chlorophyllin were determined, which allowed
synergistic PACT action based on the two PS and led to enhanced killing of these
pathogens [238]. Another major food pathogen is Bacillus cereus, which is a soilborne, Gram-positive bacterium that is able to contaminate a range of fresh vegetables, berries, and cereals, causing illnesses through the production of emetic
and diarrheal enterotoxins [239]. In vitro studies showed that ALA uptake led to
the accumulation of PPIX in cells of the organism with subsequent light activation of the PS inducing cell death [240]. In several studies, theoretical analysis of
the survival curves of B. cereus after treatment with PACT based on ALA led to
mathematical models, which were able to predict the optimal conditions for eradication of the organism using this technique [237, 241]. It has also been shown
that PACT based on ALA is effective against spores of B. cereus [240] and biofilms
of L. monocytogenes [232]. The adhesion of bacterial spores and biofilms to food
packaging surfaces is a major contamination problem, as in these forms, bacteria are highly resistant to the action of established decontamination procedures
[242–244]. Most recently, the ability of PACT based on ALA to inactivate a range
of microfungi such as Penicillium spp., Aspergillus spp. and Rhizopus spp., which
are known to contaminate wheat, was investigated, and in general, it was found
that at low ALA levels, the technique was highly effective against these organisms
although susceptibility varied between species [245]. The photodynamic mechanisms underlying this antifungal activity were not further investigated but it was
found that the application of ALA at low levels stimulated the growth of wheat
seedlings and roots. It is well established that ALA can promote plant growth
[246–249] and it was suggested that PACT based on ALA may have the potential for development as a seed decontamination technology that has no apparent
harmful effects on the food matrices [245]. Currently, the fungal contamination
of seeds for use in the food industry is becoming a serious threat to human health
and there is an urgent need for novel approaches to the problem [250–252]. Taken
with the fact that ALA appears to have no detrimental effects on food quality it
was suggested that PACT based on ALA [245] may be developed as a technique
10.4
The Applications of PACT Based on ALA
for the general microbial decontamination of both foodstuffs and their associated
processing packaging and surfaces [212, 232].
10.4.2
Dermatology Using PACT Based on ALA
Dermatology focuses on the diagnosis of skin diseases and disorders, including
skin cancers and other skin growths along with a variety of conditions associated
with microbial infections [253], many of which have the potential for treatment by
PACT based on ALA (Tables 10.1–10.3) [29, 51, 57, 221, 230, 231]. The technique
has been shown to have efficacy against some bacterial skin infections (Table 10.1)
[29, 40, 87] as in the case of several recent studies, which demonstrated that when
ALA was applied either as a cream [254, 255] or a bioadhesive patch [256], the
technique was effective against Lichen sclerosus, which is a chronic inflammatory skin disease of the anogenital region that has been associated with Borrelia
burgdorferi [257–259]. PACT based on ALA has also been shown to have efficacy
against chronic folliculitis [260], which is believed to be caused by infection due
Table 10.3 Bacteria and associated conditions susceptible to PACT based on ALA.
Gram-positive bacteria
Bacillus cereus [232, 241], including strains with MDR [116]
Enterococcus faecalis with MDR [116]
Listeria monocytogenes [232, 236]
Staphylococcus aureus [283–285], including methicillin-resistant Staphylococcus
aureus (MRSA) and other strains with MDR [81, 116, 286]
Staphylococcus epidermidis [287], including strains with MDR [116]
Streptococcus faecalis with MDR [116]
Mycobacterium marinum [288, 289]
Mycobacterium smegmatis [290]
Mycobacterium phlei [290]
Gram-negative bacteria
Acinetobacter baumannii with MDR [116]
Aeromonas hydrophila with MDR [116]
Escherichia coli [81, 285], including strains with MDR [81, 116]
Helicobacter pylori [291]
Propionibacterium acnes [275–277, 292]
Pseudomonas aeruginosa [293], including strains with MDR [81]
Salmonella enterica [232, 235]
Bacterial infections
Acne vulgaris associated with P. acnes [262, 272–274, 281, 294–296]
Chronic folliculitis associated with P. acnes, S. aureus, and some Gram-negative
bacteria [260]
Gastritis and gastroduodenal ulceration associated with H. pylori infection [291]
Lichen sclerosus associated with Borrelia burgdorferi [254–256]
Skin lesions associated with M. marinum [288, 289]
305
306
10
Photodynamic Antimicrobial Chemotherapy
to organisms such as Propionibacterium acnes and Staphylococcus aureus after
long-term administration of antibiotics [261]. However, a major focus of research
on the use of PACT based on ALA is the treatment of various types of acne [87,
262–266] and most recently, the technique has been patented in this capacity for
both ALA and a number of its derivatives [27, 267, 268]. This skin condition is
also associated with P. acnes, which normally inhabits human sebaceous glands,
and is a chronic disease of the pilosebaceous unit in which excessive hormonally
induced sebum production is followed by abnormal desquamation of follicular
corneocytes. The resulting mixture of cells and sebum leads to blockage of the
sebaceous duct and proliferation of P. acnes with subsequent release of chemotactic factors by the organism generating an inflammatory response [269–271].
Several earlier studies investigated the ability of PACT based on ALA to treat acne
lesions and reported sustained clinical improvement following the treatment [272,
273]. Subsequent results from a major clinical study suggested that the treatment
of acne vulgaris by the technique may involve a number of mechanisms, including the killing of P. acnes, the reduction of follicular obstruction, and the selective
damaging of pilosebaceous units, thereby removing blockages from the sebaceous
duct [274]. A number of later studies reported findings that strongly support these
mechanisms, including the ability of P. acnes to inherently accumulate high levels
of PPIX, which are further elevated after ALA uptake with subsequent illumination, leading to death of the organism [275–280] although other studies have
questioned the role of P. acnes in acne [73, 281, 282]. Currently, PACT based on
ALA is not used as a frontline treatment for acne and is rather employed as alternative treatment for individuals who are recalcitrant to conventional therapies and
for those for whom these therapies have been unsuccessful, although use of the
technique to treat this condition is predicted to increase [29, 87].
Over recent years, a major focus of research into dermatological potential of
PACT based on ALA has been directed toward viral skin infections (Table 10.2)
[29, 118, 297, 294], particularly those with human papillomavirus (HPV) as the
causative agent [117]. HPV is a DNA virus of the Papovaviridae family, which
causes acanthomas of the skin and adjacent mucous membranes in humans, commonly in the form of warts or verrucae [117, 298]. Recent clinical studies have
shown that PACT based on ALA can efficiently treat common warts as well as
other HPV infections [118, 214, 299], including periungual warts, which are associated with nailbeds [300–302], and anogenital warts (Condylomata acuminate),
which is one of the most common sexually transmitted diseases [214, 303–307].
A number of HPV genotypes are known to infect the genital epithelium and to
be associated with cervical and vulval premalignant lesions [308]. A recent clinical study used HPV viral genotyping to identify patients with such lesions and
when treated with ALA-based photodynamic therapy (PDT), follow-up examinations showed that 80% were HPV negative. It was suggested that this strategy
may provide a system for the early detection and treatment of women at risk from
HPV-related cervical dysplasia [309]. This suggestion has been confirmed by a
number of clinical studies that have shown that ALA-based PDT is highly effective in treating these lesions and eradicating the HPV infection [217, 310–314].
10.4
The Applications of PACT Based on ALA
Given that creams and solutions are not the most suitable drug delivery systems
for moist, irregular areas such as the lower female genital tract, an adhesive patch
for topical delivery of ALA to these bodily regions was recently developed and
was found to successfully treat HPV-related vulval dysplasia [225, 315]. ALAbased PDT appears to be of particular use to the HIV-positive community, which
has a high susceptibility to HPV infections [316–318]. It has been shown that
PACT based on ALA is successful in treating giant warts (Buschke–Loewenstein
tumor) in AIDS patients [319] and clinical trials have shown the technique to be
effective against epidermodysplasia verruciformis [320–324], which is strongly
associated with HIV-infected individuals [325, 326] and other groups that exhibit
abnormal susceptibility to HPV [327]. Most recently, it has been shown that PACT
based on ALA used in combination with immunotherapy is able to treat several
HPV-mediated conditions that are associated with the HIV community [328, 329],
including vulval intraepithelial neoplasia [217] and Bowenoid papulosis, which is
a premalignant condition affecting the anogenital area [330, 331] (Table 10.4).
PACT based on ALA has also been shown to be effective against eukaryotic
microbes associated with skin diseases such as parasites of the Leishmania genus
(Table 10.3)[40, 221, 222], including L. tropica [338] and L. major [339–343], thus
providing a potential treatment for cutaneous leishmaniasis [40, 57, 72, 221, 222,
343]. Leishmania spp. are among those parasites that are unable to synthesize
PPIX owing to the absence in their genome of genes that code for enzymes in
the heme biosynthetic pathway and it has been proposed that PACT based on
ALA kills these organisms through host cell-mediated effects, which act by inducing a systemic immune response [222–224]. Most recently, genetically modified
mutants of the Leishmania genus have been produced in which the ability to code
for enzymes in the heme biosynthetic pathway has been restored. These mutants
are able to produce PPIX and other endogenous PS and when delivered in vaccines
and activated by PACT based on ALA, show the potential to treat leishmaniasis
[344–346] as recently demonstrated using animal models [347].
The most prevalent dermatological use of PACT based on ALA against
eukaryotic microbes has been in the treatment of yeast and fungal skin infections
Table 10.4 Viruses and associated conditions susceptible to PACT based on ALA.
Human immunodeficiency virus (HIV) and associated giant warts or
Buschke–Lowenstein tumor [319]
Herpes simplex virus and associated infections [228, 319, 332, 333], including lip
lesions [334]
Human papillomavirus (HPV) and associated common warts or verruca vulgaris [117,
118, 299]
HPV and associated anogenital warts or condylomata acuminate [335, 336]
HPV and associated periungual warts [300]
HPV associated cervical and vulval lesions [217, 311–313]
HPV and associated epidermodysplasia verruciformis [320, 321]
Molluscum contagiosum virus and associated molluscum contagiosum [337]
307
308
10
Photodynamic Antimicrobial Chemotherapy
[29, 38, 40, 211–213] (Table 10.3). With regard to yeasts the technique was
recently shown to have efficacy against recalcitrant Malassezia (Pityrosporum)
folliculitis, which is localized predominantly on the back and chest and due to
invasion of the hair follicle by Malassezia furfur and other species of the genus
[348, 349]. PACT based on ALA has also been shown to efficiently treat pityriasis
versicolor (Tinea versicolor, tinea flava, and dermatomycosis furfuracea) [350],
which is a condition characterized by a rash on the trunk and proximal extremities
and is primarily caused by Malassezia globosa [351]. The technique also showed
the potential to treat candidiasis when it was demonstrated that the technique
was able to inactivate Candida albicans [352] while more recent studies suggested
that this potential could be enhanced by the supplementary use of cofactors such
as inducers and conventional antifungal agents [218]. More recently, clinical trials
showed that the technique was efficacious in treating vulvovaginal candidiasis
[353] and based on these observations, it has been suggested that the technique
has the potential to treat oral candidiasis [354]. This condition is a particular
problem for individuals with HIV/AIDS whose immuno-compromised state
makes them highly susceptible to infection by Candida spp., which can lead to
disorders such as oral thrush and oesophageal candidiasis [355] (Table 10.5).
In relation to fungal infections (Table 10.3), PACT based on ALA was found to
show limited efficacy in the treatment of tinea cruris [361, 368], which is an infection of the groin caused by Trichophyton rubrum [369, 370] and tinea unguium
(onychomycosis) [362], which is an infection of the nails that can be caused by
T. rubrum and Trichophyton interdigitale [370, 371]. In contrast, other studies
have shown that PACT based on ALA is able to efficiently treat tinea unguium
with a very high cure rate [363, 364] and, in response, a bioadhesive patch able to
deliver ALA for the treatment for onychomycosis was developed. However, ALA
showed no significant penetration of human nails or porcine hooves although
it was suggested that modifications to the technique may lead to treatment for
onychomycosis [365, 368]. PACT based on ALA has also been investigated for
the treatment of tinea pedis (interdigital mycosis) [366], which is a foot infection
mainly due to fungi of the Trichophyton genus [368, 370], These organisms were
found to accumulate PPIX in response to the addition of exogenous ALA and
after the application of PACT, some patients experienced remission from infection, although recurrences were observed [366]. Similar results for the treatment
of tinea pedis by PACT based on ALA were reported in a more recent study [367].
10.5
Future Prospects
PACT based on both MB and ALA has use in the biotechnical arena exemplified
here by blood disinfection and decontamination in the food industry. These
techniques also find strong application in the medical arena, illustrated here
by dentistry and dermatology. PACT based on MB is generally regarded as
an established antimicrobial strategy although use of the technique is rapidly
10.5
Future Prospects
Table 10.5 Eukaryotic microbes and associated conditions susceptible to PACT based on
ALA.
Parasites
Plasmodium falciparum [219]
Leishmania tropica and associated leishmaniasis [222, 338]
Leishmania major and associated leishmaniasis [222–224, 339, 340, 343]
Yeasts
Saccharomyces cerevisiae [356, 357]
Candida guilliermondii [356, 357]
Malassezia furfur and associated Malassezia (Pityrosporum) folliculitis [348]
Malassezia globosa and associated pityriasis versicolor [350]
Candida albicans and associated candidiasis [218, 352]
Fungi
Trichophyton rubrum [358–360]
T. rubrum and associated tinea cruris [361], tinea unguium [362–365], and tinea pedis
[366, 367]
Trichophyton interdigitale and associated tinea cruris [358, 365]
Trichophyton dermatophytes and associated tinea pedis [366]
Penicillium spp. [245]
Aspergillus spp. [245]
Rhizopus spp. [245]
Mucor spp. [245]
Acremonium spp. [245]
Fusarium spp. [245]
Alternaria spp. [245]
Mycelia sterilia [245]
expanding as witnessed by the diversity of its applications [5, 28, 32, 71, 72, 125,
162, 170, 178–182]. Similarly, PACT based on ALA shows great diversity in its
applications [29, 38, 40, 51, 57, 41, 71, 87, 145, 211–215, 221, 229–231] although,
currently, the technique is not generally regarded as a mainstream application of
PACT but rather as a standby modality [215, 372, 373]. However, as evidenced
by this chapter, the levels of interest and research into the use of PACT based on
ALA as an antimicrobial strategy are rapidly growing and as research improves
the understanding, operation, and efficacy of the technique, it seems destined
to become a mainstream modality [29]. For example, complete sequencing of
the genome of P. acnes has greatly illuminated the pathophysiology of acne while
techniques have been developed to selectively accumulate ALA in the sebaceous
glands when treating the disorder [87]. Taken overall, this chapter clearly shows
that PACT and particularly that form of the modality based on ALA, is an antimicrobial strategy of not only the present but more importantly perhaps, also of the
future.
309
310
10
Photodynamic Antimicrobial Chemotherapy
References
1. Wise, R. (2011) The urgent need for
2.
3.
4.
5.
6.
7.
8.
9.
new antibacterial agents. J. Antimicrob.
Chemother., 66, 1939–1940.
Livermore, D.M. (2011) Discovery
research: the scientific challenge of
finding new antibiotics. J. Antimicrob.
Chemother., 66, 1941–1944.
Finch, R. (2011) Regulatory opportunities to encourage technology
solutions to antibacterial drug resistance. J. Antimicrob. Chemother., 66,
1945–1947.
White, A.R. (2011) Effective antibacterials: at what cost? The economics
of antibacterial resistance and its control. J. Antimicrob. Chemother., 66,
1948–1953.
Wainwright, M. (2012) Photodynamic medicine and infection control.
J. Antimicrob. Chemother., 67, 787–788.
Boucher, H.W., Talbot, G.H., Bradley,
J.S., Edwards, J.E. Jr., Gilbert, D., Rice,
L.B., Scheld, M., Spellberg, B., and
Bartlett, J. (2009) Bad bugs, no drugs:
no ESKAPE! An update from the infectious diseases society of America. Clin.
Infect. Dis., 48, 1–12.
Berger, R.E. (2011) Emergence of a
new antibiotic resistance mechanism in
India, Pakistan, and the UK: a molecular, biological, and epidemiological
study editorial comment. J. Urol., 185,
154–154.
Guidos, R.J., Spellberg, B., Blaser, M.,
Boucher, H.W., Bradley, J.S., Eisenstein,
B.I., Gerding, D., Lynfield, R., Reller,
L.B., Rex, J., Schwartz, D., Septimus, E.,
Tenover, F.C., Gilbert, D.N., and IDSA
(2011) Combating antimicrobial resistance: policy recommendations to save
lives. Clin. Infect. Dis., 52, S397–S428.
Spellberg, B., Blaser, M., Guidos, R.J.,
Boucher, H.W., Bradley, J.S., Eisenstein,
B.I., Gerding, D., Lynfield, R., Reller,
L.B., Rex, J., Schwartz, D., Septimus,
E., Tenover, F.C., Gilbert, D.N., and
Infectious Diseases Society of America
(2011) Combating antimicrobial resistance: policy recommendations to save
lives. Clin. Infect. Dis., 52 (Suppl. 5),
S397–S428.
10. Cole, J. (2012) Antimicrobial resistance,
11.
12.
13.
14.
15.
16.
17.
18.
19.
infection control and planning for pandemics: the importance of knowledge
transfer in healthcare resilience and
emergency planning. J. Bus. Contin.
Emer. Plan., 6, 122–135.
Department of Health (2013) The UK
Five Year Antimicrobial Resistance
Strategy 2013–2018 (ed. D.O. Health),
HMSO, London, pp. 1–43.
The World Health Organization (2012)
The Evolving Threat of Antimicrobial
Resistance – Options for Action, World
Health Orginaziation, Geneva.
Lee, C.-R., Cho, I.H., Jeong, B.C., and
Lee, S.H. (2013) Strategies to minimize
antibiotic resistance. Int. J. Environ. Res.
Public Health, 10, 4274–4305.
Piacenti, F.J. and Leuthner, K.D. (2013)
Antimicrobial stewardship and clostridium difficile-associated diarrhea.
J. Pharm. Pract., 26, 506–513.
Moro, M.L. and Gagliotti, C. (2013)
Antimicrobial resistance and stewardship in long-term care settings. Future
Microbiol., 8, 1011–1025.
Bogan, C. and Marchaim, D. (2013) The
role of antimicrobial stewardship in
curbing carbapenem resistance. Future
Microbiol., 8, 979–991.
Cairns, K.A., Jenney, A.W.J., Abbott,
I.J., Skinner, M.J., Doyle, J.S., Dooley,
M., and Cheng, A.C. (2013) Prescribing
trends before and after implementation of an antimicrobial stewardship
program. Med. J. Aust., 198, 262–266.
Chandrasiri, P., Elwitigala, J.P.,
Nanayakkara, G., Chandrasiri, S.,
Patabendige, G., Karunanayaka,
L., Perera, J., Somaratne, P., and
Jayathilleke, K. (2013) A multi centre laboratory study of Gram negative
bacterial blood stream infections in
Sri Lanka. Ceylon Med. J., 58, 56–61.
Lesprit, P., Landelle, C., and
Brun-Buisson, C. (2013) Clinical impact
of unsolicited post-prescription antibiotic review in surgical and medical
wards: a randomized controlled trial.
Clin. Microbiol. Infect., 19, E91–E97.
References
20. Pucci, M.J. and Bush, K. (2013) Inves-
21.
22.
23.
24.
25.
26.
27.
28.
tigational antimicrobial agents of 2013.
Clin. Microbiol. Rev., 26, 792–821.
Alvan, G., Edlund, C., and Heddini,
A. (2011) The global need for effective antibiotics-a summary of plenary
presentations. Drug Resist. Updat., 14,
70–76.
Maisch, T. (2007) Revitalized strategies
against multi-resistant bacteria: antimicrobial photodynamic therapy and
bacteriophage therapy. Anti-Infective
Agents Med. Chem., 6, 145–150.
Almeida, A., Cunha, A., Gomes,
N.C.M., Alves, E., Costa, L., and
Faustino, M.A.F. (2009) Phage therapy and photodynamic therapy: low
environmental impact approaches to
inactivate microorganisms in fish farming plants. Mar. Drugs, 7, 268–313.
Vatansever, F., de Melo, W.C.M.A.,
Avci, P., Vecchio, D., Sadasivam, M.,
Gupta, A., Chandran, R., Karimi, M.,
Parizotto, N.A., Yin, R., Tegos, G.P., and
Hamblin, M.R. (2013) Antimicrobial
strategies centered around reactive oxygen species – bactericidal antibiotics,
photodynamic therapy, and beyond.
FEMS Microbiol. Rev., 37, 955–989.
Hamblin, M.R. and Hasan, T. (2004)
Photodynamic therapy: a new antimicrobial approach to infectious disease?
Photochem. Photobiol. Sci., 3, 436–450.
de Melo, W.C.M.A., Avci, P., de
Oliveira, M.N., Gupta, A., Vecchio, D.,
Sadasivam, M., Chandran, R., Huang,
Y.-Y., Yin, R., Perussi, L.R., Tegos, G.P.,
Perussi, J.R., Dai, T., and Hamblin,
M.R. (2013) Photodynamic inactivation of biofilm: taking a lightly colored
approach to stubborn infection. Expert
Rev. Anti Infect. Ther., 11, 669–693.
Nitzan, Y. and Nisnevitch, M. (2013)
Special features of gram-positive bacterial eradication by photosensitizers.
Recent Pat. Antiinfect. Drug Discov., 8,
88–99.
Sperandio, F.F., Huang, Y.-Y., and
Hamblin, M.R. (2013) Antimicrobial
photodynamic therapy to kill Gramnegative bacteria. Recent Pat. Antiinfect.
Drug Discov., 8, 108–120.
29. Harris, F. and Pierpoint, L. (2012)
30.
31.
32.
33.
34.
35.
36.
Photodynamic therapy based on 5aminolevulinic acid and its use as an
antimicrobial agent. Med. Res. Rev., 32,
1292–1327.
Wainwright, M. (2011) Photodynamic
Inactivation of Microbial Pathogens:
Medical and Environmental Applications, The Royal Society of Chemistry,
pp. 19–43.
Tardivo, J.P., Del Giglio, A., de Oliveira,
C.S., Gabrielli, D.S., Junqueira, H.C.,
Tada, D.B., Severino, D., Turchiello,
R.d.F., and Baptista, M.S. (2005) Methylene blue in photodynamic therapy:
from basic mechanisms to clinical
applications. Photodiagnosis Photodyn.
Ther., 2, 175–191.
Harris, F., Chatfield, L.K., and Phoenix,
D.A. (2005) Phenothiazinium based
photosensitisers – photodynamic
agents with a multiplicity of cellular
targets and clinical applications. Curr.
Drug Targets, 6, 615–627.
Rodrigues, G.B., Dias-Baruffi, M.,
Holman, N., Wainwright, M., and
Braga, G.U.L. (2013) In vitro photodynamic inactivation of Candida species
and mouse fibroblasts with phenothiazinium photosensitisers and red light.
Photodiagnosis Photodyn. Ther., 10,
141–149.
Rodrigues, G.B., Ferreira, L.K.S.,
Wainwright, M., and Braga, G.U.L.
(2012) Susceptibilities of the dermatophytes Trichophyton mentagrophytes
and T. rubrum microconidia to photodynamic antimicrobial chemotherapy
with novel phenothiazinium photosensitizers and red light. J. Photochem.
Photobiol., B, 116, 89–94.
Wainwright, M., Brandt, S.D., Smith,
A., Styles, A., Meegan, K., and
Loughran, C. (2010) Phenothiazinium
photosensitisers VII: novel substituted
asymmetric N-benzylphenothiaziniums
as photoantimicrobial agents. J. Photochem. Photobiol., B, 99, 74–77.
Wainwright, M., Meegan, K., and
Loughran, C. (2011) Phenothiazinium
photosensitisers IX. Tetra- and pentacyclic derivatives as photoantimicrobial
agent. Dyes Pigm., 91, 1–5.
311
312
10
Photodynamic Antimicrobial Chemotherapy
37. Wainwright, M., Smalley, H., Scully,
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.
O., and Lotfipour, E. (2012) Comparative photodynamic evaluation of new
phenothiazinium derivatives against
propionibacterium acnes. Photochem.
Photobiol., 88, 523–526.
Kharkwal, G.B., Sharma, S.K., Huang,
Y.-Y., Dai, T., and Hamblin, M.R. (2011)
Photodynamic therapy for infections:
clinical applications. Lasers Surg. Med.,
43, 755–767.
Phoenix, D.A. and Harris, F. (2003)
Phenothiazinium-based photosensitizers: antibacterials of the future? Trends
Mol. Med., 9, 283–285.
Dai, T., Huang, Y.-Y., and Hamblin,
M.R. (2009) Photodynamic therapy
for localized infections-State of the
art. Photodiagnosis Photodyn. Ther., 6,
170–188.
Ormond, A.B. and Freeman, H.S.
(2013) Dye sensitizers for photodynamic therapy. Materials, 6, 817–840.
Sternberg, E. and Dolphin, D. (1996)
Pyrrolic photosensitizers. Curr. Med.
Chem., 3, 239–272.
Ajioka, R.S., Phillips, J.D., and Kushner,
J.P. (2006) Biosynthesis of heme in
mammals. Biochim. Biophys. Acta,
1763, 723–736.
Schobert, M. and Jahn, D. (2002)
Regulation of heme biosynthesis in
non-phototrophic bacteria. J. Mol.
Microbiol. Biotechnol., 4, 287–294.
Heinemann, I.U., Jahn, M., and Jahn,
D. (2008) The biochemistry of heme
biosynthesis. Arch. Biochem. Biophys.,
474, 238–251.
O’Brian, M.R. (2009) Heme Biosynthesis.
Fukuda, H., Casas, A., and Batlle, A.
(2005) Aminolevulinic acid: from its
unique biological function to its star
role in photodynamic therapy. Int. J.
Biochem. Cell Biol., 37, 272–276.
Batlle, A.M. (1993) Porphyrins, porphyrias, cancer and photodynamic
therapy – a model for carcinogenesis.
J. Photochem. Photobiol. B, 20, 5–22.
Kennedy, J.C. (2006) in Photodynamic
Therapy with ALA (eds R. Pottier, B.
Krammer, R. Baumgartner, and H.
Stepp), Royal Society of Chemistry,
Cambridge, pp. 1–15.
50. Zakhary, K. and Ellis, D.A.F. (2005)
51.
52.
53.
54.
55.
56.
57.
58.
59.
Applications of aminolevulinic acidbased photodynamic therapy in
cosmetic facial plastic practices. Facial
Plast. Surg., 21, 110–116.
Darlenski, R. and Fluhr, J.W. (2013)
Photodynamic therapy in dermatology:
past, present, and future. J. Biomed.
Opt., 18, 061208.
Nokes, B., Apel, M., Jones, C., Brown,
G., and Lang, J.E. (2013) Aminolevulinic acid (ALA): photodynamic
detection and potential therapeutic applications. J. Surg. Res., 181,
262–271.
Ishizuka, M., Abe, F., Sano, Y.,
Takahashi, K., Inoue, K., Nakajima,
M., Kohda, T., Komatsu, N., Ogura,
S.-i., and Tanaka, T. (2011) Novel development of 5-aminolevurinic acid (ALA)
in cancer diagnoses and therapy. Int.
Immunopharmacol., 11, 358–365.
Wachowska, M., Muchowicz, A.,
Firczuk, M., Gabrysiak, M., Winiarska,
M., Wanczyk, M., Bojarczuk, K., and
Golab, J. (2011) Aminolevulinic acid
(ALA) as a prodrug in photodynamic therapy of cancer. Molecules,
16, 4140–4164.
Gaal, M., Kui, R., Hunyadi, Z., Kemeny,
L., and Gyulai, R. (2012) Fluorescence
diagnosis of non-melanoma skin cancer.
Orv. Hetil., 153, 1334–1340.
Babilas, P., Schreml, S., Landthaler,
M., and Szeimies, R.-M. (2010) Photodynamic therapy in dermatology:
state-of-the-art. Photodermatol. Photoimmunol. Photomed., 26, 118–132.
Morton, C.A., Szeimies, R.M., Sidoroff,
A., and Braathen, L.R. (2013) European
guidelines for topical photodynamic
therapy part 2: emerging indications – field cancerization, photorejuvenation and inflammatory/infective
dermatoses. J. Eur. Acad. Dermatol.
Venereol., 27, 672–679.
Kim, R.H. and Armstrong, A.W. (2011)
Current state of acne treatment: highlighting lasers, photodynamic therapy,
and chemical peels. Dermatol. Online J.,
17, 2.
Kennedy, J.C., Pottier, R.H., and Pross,
D.C. (1990) Photodynamic therapy with
endogenous protoporphyrin IX: basic
References
60.
61.
62.
63.
64.
65.
66.
67.
68.
69.
principles and present clinical experience. J. Photochem. Photobiol. B, 6,
143–148.
Pushpan, S.K., Venkatraman, S., Anand,
V.G., Sankar, J., Parmeswaran, D.,
Ganesan, S., and Chandrashekar, T.K.
(2002) Porphyrins in photodynamic
therapy: a search for ideal photosensitizers. Curr. Med. Chem., 2, 187–207.
Wilson, B.C. and Patterson, M.S. (2008)
The physics, biophysics and technology
of photodynamic therapy. Phys. Med.
Biol., 53, R61–R109.
Calzavara-Pinton, P.G., Rossi, M.T.,
and Ortel, B. (2011) in Photodynamic
Inactivation of Microbial Pathogens:
Medical and Environmental Applications (eds M.R. Hamblin and
G. Jori), Royal Society for Chemistry,
Cambridge, pp. 361–375.
Ryskova, L., Buchta, V., and Slezak, R.
(2010) Photodynamic antimicrobial
therapy. Centr. Eur. J. Biol., 5, 400–406.
Donnelly, R.F., Cassidy, C.M., and
Tunney, M.M. (2011) in Photodynamic
Inactivation of Microbial Pathogens:
Medical and Environmental Applications (eds M.R. Hamblin and G. Jori),
Royal Society for Chemistry, Cambridge, pp. 185–216.
Boucher, D. (2011) The tools of PDT:
light sources and devices. Can they
help in getting better therapeutic
results?. AIP Conf. Proc., 1364, 73–86,
http://dx.doi.org/10.1063/1.3626915
(accessed 11 July 2014).
Mang, T.S. (2004) Lasers and light
sources for PDT: past, present and
future. Photodiagnosis Photodyn. Ther.,
1, 43–48.
Issa, M.C.A. and Manela-Azulay, M.
(2010) Terapia fotodinâmica: revisão da
literatura e documentação iconográfica.
An. Bras. Dermatol., 85, 501–511.
Allemann, I.B., Goldberg, D.J.,
Allemann, I.B., and Goldberg, D.J.
(2011) Basics in Dermatological Laser
Applications.
Wainwright, M. (2009) Photosensitisers in Biomedicine, Wiley-Blackwell,
Chichester.
70. Calin, M. and Parasca, S.V. (2009) Light
71.
72.
73.
74.
75.
76.
77.
78.
79.
sources for photodynamic inactivation of bacteria. Lasers Med. Sci., 24,
453–460.
Calzavara-Pinton, P., Rossi, M.T., Sala,
R., and Venturini, M. (2012) Photodynamic antifungal chemotherapy†.
Photochem. Photobiol., 88, 512–522.
Baptista, M.S. and Wainwright, M.
(2011) Photodynamic antimicrobial
chemotherapy (PACT) for the treatment of malaria, leishmaniasis and
trypanosomiasis. Braz. J. Med. Biol.
Res., 44, 1–10.
Cassidy, C.M., Tunney, M.M.,
McCarron, P.A., and Donnelly, R.F.
(2009) Drug delivery strategies for photodynamic antimicrobial chemotherapy:
from benchtop to clinical practice.
J. Photochem. Photobiol., B, 95, 71–80.
Zhang, L.W., Fang, Y.P., and Fang, J.Y.
(2011) Enhancement techniques for
improving 5-aminolevulinic acid delivery through the skin. Dermatol. Sin.,
29, 1–7.
Felgentrager, A., Maisch, T., Dobler,
D., and Spath, A. (2013) Hydrogen
bond acceptors and additional cationic
charges in methylene blue derivatives: photophysics and antimicrobial
efficiency. Biomed. Res. Int., 2013,
482167.
Fukuda, H., Casas, A., and Batlle,
A. (2006) Use of ALA and ALA
derivatives for optimizing ALA-based
photodynamic therapy: a review of our
experience. J. Environ. Pathol. Toxicol.
Oncol., 25, 127–143.
Lopez, R.F., Lange, N., Guy, R., and
Bentley, M.V. (2004) Photodynamic
therapy of skin cancer: controlled drug
delivery of 5-ALA and its esters. Adv.
Drug Deliv. Rev., 56, 77–94.
Giuntini, F., Bourre, L., MacRobert,
A.J., Wilson, M., and Eggleston, I.M.
(2009) Improved peptide prodrugs of
5-ALA for PDT: rationalization of cellular accumulation and protoporphyrin
IX production by direct determination
of cellular prodrug uptake and prodrug
metabolization. J. Med. Chem., 52,
4026–4037.
Bourre, L., Giuntini, F., Eggleston, I.M.,
Wilson, M., and MacRobert, A.J. (2009)
313
314
10
80.
81.
82.
83.
84.
85.
86.
Photodynamic Antimicrobial Chemotherapy
Protoporphyrin IX enhancement by 5aminolaevulinic acid peptide derivatives
and the effect of RNA silencing on
intracellular metabolism. Br. J. Cancer,
100, 723–731.
Vallinayagam, R., Schmitt, F., Barge, J.,
Wagnieres, G., Wenger, V., Neier, R.,
and Juillerat-Jeanneret, L. (2008) Glycoside esters of 5-aminolevulinic acid
for photodynamic therapy of cancer.
Bioconjug. Chem., 19, 821–839.
Fotinos, N., Convert, M., Piffaretti,
J.-C., Gurny, R., and Lange, N.
(2008) Effects on gram-negative
and gram-positive bacteria mediated by 5-aminolevulinic acid and
5-aminolevulinic acid derivatives.
Antimicrob. Agents Chemother., 52,
1366–1373.
Gerritsen, M.J.P., Smits, T.,
Kleinpenning, M.M., van de Kerkhof,
P.C.M., and van Erp, P.E.J. (2009) Pretreatment to enhance protoporphyrin
IX accumulation in photodynamic
therapy. Dermatology, 218, 193–202.
Juzeniene, A., Juzenas, P., Kaalhus,
O., Iani, V., and Moan, J. (2002) Temperature effect on accumulation of
protoporphyrin IX after topical application of 5-aminolevulinic acid and its
methylester and hexylester derivatives
in normal mouse skin. Photochem.
Photobiol., 76, 452–456.
Ramstad, S., Le Anh-Vu, N., and
Johnsson, A. (2006) The temperature
dependence of porphyrin production in Propionibacterium acnes after
incubation with 5-aminolevulinic acid
(ALA) and its methyl ester (m-ALA).
Photochem. Photobiol. Sci., 5, 66–72.
Yang, J.B., Chen, A.C.H., Wu, Q.H.,
Jiang, S., Liu, X.M., Xiong, L.Y., and
Xia, Y.M. (2010) The influence of
temperature on 5-aminolevulinic
acid-based photodynamic reaction
in keratinocytes in vitro. Photodermatol. Photoimmunol. Photomed., 26,
83–88.
Donnelly, R.F., McCarron, P.A., and
Woolfson, A.D. (2005) Drug delivery of
aminolevulinic acid from topical formulations intended for photodynamic
therapy. Photochem. Photobiol., 81,
750–767.
87. Hsieh, M.-F. and Chen, C.-H. (2012)
88.
89.
90.
91.
92.
93.
94.
95.
Review: delivery of pharmaceutical
agents to treat acne vulgaris: current
status and perspectives. J. Med. Biol.
Eng., 32, 215–223.
de Leeuw, J., van der Beek, N., Bjerring,
P., and Neumann, H.A.M. (2010) Photodynamic therapy of acne vulgaris
using 5-aminolevulinic acid 0.5% liposomal spray and intense pulsed light
in combination with topical keratolytic agents. J. Eur. Acad. Dermatol.
Venereol., 24, 460–469.
de Campos Araújo, L.M.P., Thomazine,
J.A., and Lopez, R.F.V. (2010) Development of microemulsions to topically
deliver 5-aminolevulinic acid in photodynamic therapy. Eur. J. Pharm.
Biopharm., 75, 48–55.
Winkler, A. and Müller-Goymann,
C.C. (2005) The influence of topical
formulations on the permeation of
5-aminolevulinic acid and its n-butyl
ester through excised human stratum
corneum. Eur. J. Pharm. Biopharm., 60,
427–437.
Dragicevic-Curic, N. and Fahr, A.
(2012) Liposomes in topical photodynamic therapy. Expert Opin. Drug
Deliv., 9, 1015–1032.
Donnelly, R.F., Morrow, D.I.J.,
McCarron, P.A., Woolfson, A.D.,
Morrissey, A., Juzenas, P., Juzeniene,
A., Iani, V., McCarthy, H.O., and Moan,
J. (2008) Microneedle-mediated intradermal delivery of 5-aminolevulinic
acid: Potential for enhanced topical
photodynamic therapy. J. Control.
Release, 129, 154–162.
Donnelly, R.F., Singh, T.R.R., Tunney,
M.M., Morrow, D.I.J., McCarron, P.A.,
O’Mahony, C., and Woolfson, A.D.
(2009) Microneedle arrays allow lower
microbial penetration than hypodermic needles in vitro. Pharm. Res., 26,
2513–2522.
Morrow, D.I.J. and Donnelly, R.F.
(2009) Novel drug delivery strategies for porphyrins and porphyrin
precursors. Paper Presented at the Photodynamic Therapy: Back to the Future,
Seattle, WA.
Morrow, D.I.J., Garland, M.J.,
McCarron, P.A., Woolfson, A.D., and
References
96.
97.
98.
99.
100.
101.
102.
103.
Donnelly, R.F. (2007) Innovative drug
delivery strategies for topical photodynamic therapy using porphyrin
precursors. J. Environ. Pathol. Toxicol.
Oncol., 26, 105–116.
Morrow, D.I.J., McCarron, P.A.,
Woolfson, A.D., Juzenas, P., Juzeniene,
A., Iani, V., Moan, J., and Donnelly, R.F.
(2010) Novel patch-based systems for
the localised delivery of ALA-esters.
J. Photochem. Photobiol., B, 101, 59–69.
Donnelly, R.F., McCarron, P.A.,
Morrow, D.I.J., and Woolfson, A.D.
(2013) Fast-drying multi-laminate
bioadhesive films for transdermal and
topical drug delivery. Drug Dev. Ind.
Pharm., 39, 1818–1831.
Donnelly, R.F., McCarron, P.A., and
Woolfson, D. (2009) Drug delivery systems for photodynamic therapy. Recent
Pat. Drug Deliv. Formul., 3, 1–7.
Fallows, S.J., Garland, M.J., Cassidy,
C.M., Tunney, M.M., Singh, T.R.R.,
and Donnelly, R.F. (2012) Electricallyresponsive anti-adherent hydrogels
for photodynamic antimicrobial
chemotherapy. J. Photochem. Photobiol.,
B, 114, 61–72.
Donnelly, R.F., Cassidy, C.M., Loughlin,
R.G., Brown, A., Tunney, M.M.,
Jenkins, M.G., and McCarron, P.A.
(2009) Delivery of Methylene blue
and meso-tetra (N-methyl-4-pyridyl)
porphine tetra tosylate from crosslinked poly(vinyl alcohol) hydrogels: a
potential means of photodynamic therapy of infected wounds. J. Photochem.
Photobiol., B, 96, 223–231.
Ali, M.H.M., Hashem, M.M., Zaher,
A., Korraa, S., Hamouda, F., Ali, C.M.,
and Al-Saad, K.A. (2013) Photodynamic therapy for hair removal. QSci.
Connect, 16, 1–10.
Cassidy, C.M., Tunney, M.M., Magee,
N.D., Elborn, J.S., Bell, S., Singh, T.R.R.,
and Donnelly, R.F. (2011) Drug and
light delivery strategies for photodynamic antimicrobial chemotherapy
(PACT) of pulmonary pathogens: a
pilot study. Photodiagnosis Photodyn.
Ther., 8, 1–6.
Fadel, M., Salah, M., Samy, N., and
Soliman, M. (2009) Liposomal methylene blue hydrogel for selective
104.
105.
106.
107.
108.
109.
110.
111.
112.
photodynamic therapy of acne vulgaris.
J. Drugs Dermatol., 8, 983–990.
Mordon, S., Sumian, C., and
Devoisselle, J.M. (2003) Site-specific
methylene blue delivery to pilosebaceous structures using highly porous
nylon microspheres: an experimental evaluation. Lasers Surg. Med., 33,
119–125.
Well, D. (2013) Acne vulgaris: a review
of causes and treatment options. Nurse
Pract., 38, 22–31; quiz 32.
Nakonechny, F., Nisnevitch, M., Nitzan,
Y., and Firer, M.A. (2011) New techniques in antimicrobial photodynamic
therapy: scope of application and overcoming drug resistance in nosocomial
infections, in Science Against Microbial
Pathogens: Communicating Current
Research and Technological Advances,
Formatex Research Center Publisher,
Spain.
Nitzan, Y. and Pechatnikov, I. (2011) in
Photodynamic Inactivation of Microbial
Pathogens – Medical and Environmental Applications (eds M.R. Hamblin and
G. Jori), Royal Society of Chemistry
Publishing, pp. 45–67.
Nakonechny, F., Firer, M.A., Nitzan, Y.,
and Nisnevitch, M. (2010) Intracellular
antimicrobial photodynamic therapy: a
novel technique for efficient eradication
of pathogenic bacteria. Photochem.
Photobiol., 86, 1350–1355.
Nisnevitch, M., Nakonechny, F., Firer,
M., and Nitzan, Y. (2009) Photodynamic antimicrobial chemotherapy
with photosensitizers in liposomes
under external and chemiluminescent
excitation. FEBS J., 276, 332–332.
Garland, M.J., Cassidy, C.M., Woolfson,
D., and Donnelly, R.F. (2009) Designing photosensitizers for photodynamic
therapy: strategies, challenges and
promising developments. Future Med.
Chem., 1, 667–691.
Demidova, T.N. and Hamblin, M.R.
(2004) Photodynamic therapy targeted
to pathogens. Int. J. Immunopathol.
Pharmacol., 17, 245–254.
Sharma, S.K., Dai, T., Kharkwal, G.B.,
Huang, Y.-Y., Huang, L., De Arce, V.J.B.,
Tegos, G.P., and Hamblin, M.R. (2011)
315
316
10
113.
114.
115.
116.
117.
118.
119.
120.
121.
122.
Photodynamic Antimicrobial Chemotherapy
Drug discovery of antimicrobial photosensitizers using animal models. Curr.
Pharm. Des., 17, 1303–1319.
George, S., Hamblin, M.R., and Kishen,
A. (2009) Uptake pathways of anionic
and cationic photosensitizers into bacteria. Photochem. Photobiol. Sci., 8,
788–795.
Wainwright, M. and Crossley, K.B.
(2002) Methylene blue – a therapeutic
dye for all seasons? J. Chemother., 14,
431–443.
Prates, R.A., Kato, I.T., Ribeiro, M.S.,
Tegos, G.P., and Hamblin, M.R. (2011)
Influence of multidrug efflux systems
on methylene blue-mediated photodynamic inactivation of Candida
albicans. J. Antimicrob. Chemother., 66,
1525–1532.
Nitzan, Y., Salmon-Divon, M., Shporen,
E., and Malik, Z. (2004) ALA induced
photodynamic effects on gram positive and negative bacteria. Photochem.
Photobiol. Sci., 3, 430–435.
Szeimies, R.-M. (2003) Photodynamic
therapy for human papilloma virusrelated diseases in dermatology. Med.
Laser Appl., 18, 107–116.
Rossi, R., Bruscino, N., Ricceri, F.,
Grazzini, M., Dindelli, M., and Lotti,
T. (2009) Photodynamic treatment for
viral infections of the skin. G. Ital.
Dermatol. Venereol., 144, 79–83.
Verkamp, E., Backman, V.M.,
Bjornsson, J.M., Soll, D., and
Eggertsson, G. (1993) The periplasmic
dipeptide permease system transports
5-aminolevulinic acid in escherichiacoli. J. Bacteriol., 175, 1452–1456.
Elliott, T. (1993) Transport of 5aminolevulinic acid by the dipeptide
permease in salmonella-typhimurium.
J. Bacteriol., 175, 325–331.
King, N.D. and Obrian, M.R. (1997)
Identification of the lrp gene in
Bradyrhizobium japonicum and its role
in regulation of delta-aminolevulinic
acid uptake. J. Bacteriol., 179,
1828–1831.
Bermudez Moretti, M., Perullini,
A.M., Batlle, A., and Correa Garcia,
S. (2005) Expression of the UGA4 gene
encoding the delta-aminolevulinic and
gamma-aminobutyric acids permease in
123.
124.
125.
126.
127.
128.
129.
130.
131.
Saccharomyces cerevisiae is controlled
by amino acid-sensing systems. Arch.
Microbiol., 184, 137–140.
Bermudez Moretti, M., Correa Garcia,
S., and Batlle, A. (2000) Porphyrin
biosynthesis intermediates are not
regulating delta-aminolevulinic acid
transport in Saccharomyces cerevisiae.
Biochem. Biophys. Res. Commun., 272,
946–950.
Moretti, M.B., Batlle, A., and Garcia,
S.C. (2001) UGA4 gene encoding the
gamma-aminobutyric acid permease in
Saccharomyces cerevisiae is an acidexpressed gene. Int. J. Biochem. Cell
Biol., 33, 1202–1207.
Phoenix, D.A. and Harris, F. (2010)
in Froniters in Anti-infective Drug
Discovery (eds A. Rahman and M.I.
Choudhary), Benthams, London, pp.
17–48.
Krammer, B., Malik, Z., Pottier, R., and
Stepp, H. (2006) in Photodynamic Therapy with ALA: A Clinical Handbook
(eds R. Pottier, B. Krammer, H. Stepp,
R. Baumgartner, and H. Stepp), Royal
Society of Chemistry, Cambridge, pp.
15–77.
Liu, B.C., Farrell, T.J., and Patterson,
M.S. (2010) A dynamic model for ALAPDT of skin: simulation of temporal
and spatial distributions of groundstate oxygen, photosensitizer and
singlet oxygen. Phys. Med. Biol., 55,
5913–5932.
Arnaut, L.G. (2011) in Advances in
Inorganic Chemistry, Inorganic Photochemistry, Vol. 63 (eds R. VanEldik and
G. Stochel), Elsevier Academic Press
Inc., San Diego, CA, pp. 187–233.
Denis, T.G.S., Dai, T.H., Izikson, L.,
Astrakas, C., Anderson, R.R., Hamblin,
M.R., and Tegos, G.P. (2011) All you
need is light antimicrobial photoinactivation as an evolving and emerging
discovery strategy against infectious
disease. Virulence, 2, 509–520.
Pawlicki, M., Collins, H.A., Denning,
R.G., and Anderson, H.L. (2009) Twophoton absorption and the design of
two-photon dyes. Angew. Chem. Int.
Ed., 48, 3244–3266.
Ogawa, K. and Kobuke, Y. (2013)
Two-photon photodynamic therapy by
References
132.
133.
134.
135.
136.
137.
138.
139.
140.
141.
142.
143.
water-soluble self-assembled conjugated
porphyrins. Biomed. Res. Int., 2013,
125658–125658.
Kozinska, A., Oles, T., and Sarna, T.
(2012) Photoactivation and detection
of photoexcited molecules and photochemical products. Isr. J. Chem., 52,
745–756.
Ortel, B., Shea, C.R., and
Calzavara-Pinton, P. (2009) Molecular mechanisms of photodynamic
therapy. Front. Biosci., 14, 4157–4172.
Auten, R.L. and Davis, J.M. (2009)
Oxygen toxicity and reactive oxygen
species: the devil is in the details.
Pediatr. Res., 66, 121–127.
Imlay, J.A. (2003) Pathways of oxidative
damage. Annu. Rev. Microbiol., 57,
395–418.
Kerdous, R., Heuvingh, J., and Bonneau,
S. (2011) Photo-dynamic induction
of oxidative stress within cholesterolcontaining membranes: shape transitions and permeabilization. Biochim.
Biophys. Acta, 1808, 2965–2972.
Yoda, T., Vestergaard, M.C.,
Akazawa-Ogawa, Y., Yoshida, Y.,
Hamada, T., and Takagi, M. (2010)
Dynamic response of a cholesterolcontaining model membrane to
oxidative stress. Chem. Lett., 39,
1273–1274.
Girotti, A.W. (2001) Photosensitized
oxidation of membrane lipids: reaction
pathways, cytotoxic effects, and cytoprotective mechanisms. J. Photochem.
Photobiol., B, 63, 103–113.
Girotti, A.W. and Kriska, T. (2004)
Role of lipid hydroperoxides in photooxidative stress signaling. Antioxid.
Redox Signal., 6, 301–310.
Davies, M.J. (2003) Singlet oxygenmediated damage to proteins and its
consequences. Biochem. Biophys. Res.
Commun., 305, 761–770.
Davies, M.J. (2005) The oxidative environment and protein damage. Biochim.
Biophys. Acta, 1703, 93–109.
Davies, M.J. (2004) Reactive species
formed on proteins exposed to singlet
oxygen. Photochem. Photobiol. Sci., 3,
17–25.
Barelli, S., Canellini, G., Thadikkaran,
L., Crettaz, D., Quadroni, M., Rossier,
144.
145.
146.
147.
148.
149.
150.
151.
152.
J.S., Tissot, J.D., and Lion, N. (2008)
Oxidation of proteins: Basic principles
and perspectives for blood proteomics.
Proteomics Clin. Appl., 2, 142–157.
Agnez-Lima, L.F., Melo, J.T.A., Silva,
A.E., Oliveira, A.H.S., Timoteo, A.R.S.,
Lima-Bessa, K.M., Martinez, G.R.,
Medeiros, M.H.G., Di Mascio, P.,
Galhardo, R.S., and Menck, C.F.M.
(2012) DNA damage by singlet oxygen
and cellular protective mechanisms.
Mutation Res. Rev. Mutation Res., 751,
15–28.
Costa, L., Faustino, M.A.F., Neves,
M.G.P.M.S., Cunha, A., and Almeida,
A. (2012) Photodynamic inactivation of
mammalian viruses and bacteriophages.
Viruses-Basel, 4, 1034–1074.
Loft, S., Danielsen, P., Løhr, M.,
Jantzen, K., Hemmingsen, J.G.,
Roursgaard, M., Karotki, D.G., and
Møller, P. (2012) Urinary excretion of
8-oxo-7,8-dihydroguanine as biomarker
of oxidative damage to DNA. Arch.
Biochem. Biophys., 518, 142–150.
Phoenix, D.A. and Harris, F. (2006)
Light activated compounds as antimicrobial agents – patently obvious?
Recent Pat. Antiinfect. Drug Discov., 1,
181–199.
Ginimuge, P.R. and Jyothi, S.D. (2010)
Methylene blue: revisited. J. Anaesthesiology, Clin. Pharmacol., 26, 517–520.
Miclescu, A. and Wiklund, L. (2010)
Methylene blue, an old drug with
new indications? Rom. J. Anaesthesia
Intensive Care, 17, 35–41.
Fayter, D., Corbett, M., Heirs, M., Fox,
D., and Eastwood, A. (2010) A systematic review of photodynamic therapy
in the treatment of pre-cancerous skin
conditions, Barrett’s oesophagus and
cancers of the biliary tract, brain, head
and neck, lung, oesophagus and skin.
Health Technol. Assess., 14, 1–288.
Rice, L., Wainwright, M., and Phoenix,
D.A. (2000) Phenothiazine photosensitizers. III. Activity of methylene
blue derivatives against pigmented
melanoma cell lines. J. Chemother., 12,
94–104.
Wainwright, M., Phoenix, D.A., Burrow,
S.M., and Waring, J. (1999) Cytotoxicity
317
318
10
153.
154.
155.
156.
157.
158.
159.
160.
161.
Photodynamic Antimicrobial Chemotherapy
and adjuvant activity of cationic photosensitizers in a multidrug resistant cell
line. J. Chemother., 11, 61–68.
Schirmer, R.H., Adler, H., Pickhardt,
M., and Mandelkow, E. (2011) Lest
we forget you – methylene blue … .
Neurobiol. Aging, 32, 2325.
Ohlow, M.J. and Moosmann, B. (2011)
Foundation review: phenothiazine:
the seven lives of pharmacology’s first
lead structure. Drug Discov. Today, 16,
119–131.
Wainwright, M. (2001) Acridine – a
neglected antibacterial chromophore.
J. Antimicrob. Chemother., 47, 1–13.
Wainwright, M. (2003) The use of
dyes in modern biomedicine. Biotech.
Histochem., 78, 147–155.
Wainwright, M., Phoenix, D.A.,
Marland, J., Wareing, D.R.A., and
Bolton, F.J. (1997) In-vitro photobactericidal activity of aminoacridines.
J. Antimicrob. Chemother., 40, 587–589.
Wainwright, M., Phoenix, D.A.,
Marland, J., Wareing, D.R.A., and
Bolton, F.J. (1998) A comparison of
the bactericidal and photobactericidal activities of aminoacridines and
bis(aminoacridines). Lett. Appl. Microbiol., 26, 404–406.
Nagata, J.Y., Hioka, N., Kimura, E.,
Batistela, V.R., Suga Terada, R.S.,
Graciano, A.X., Baesso, M.L., and
Hayacibara, M.F. (2012) Antibacterial photodynamic therapy for dental
caries: evaluation of the photosensitizers used and light source properties.
Photodiagnosis Photodyn. Ther., 9,
122–131.
Hasegawa, H., Sato, S., Kawauchi, S.,
Saitoh, D., Shinomiya, N., Ashida, H.,
and Terakawa, M. (2012) Control of
burn wound sepsis in rats by methylene
blue-mediated photodynamic treatment, in Optical Methods for Tumor
Treatment and Detection: Mechanisms
and Techniques in Photodynamic Therapy Xxi (eds D.H. Kessel and T. Hasan),
SPIE-International Society for Optical
Engineering, Bellingham.
Perni, S., Prokopovich, P., Pratten, J.,
Parkin, I.P., and Wilson, M. (2011)
Nanoparticles: their potential use
162.
163.
164.
165.
166.
167.
168.
169.
170.
171.
in antibacterial photodynamic therapy. Photochem. Photobiol. Sci., 10,
712–720.
Sun, G. and Hong, K.H. (2013) Photoinduced antimicrobial and decontaminating agents: recent progresses in
polymer and textile applications. Text.
Res. J., 83, 532–542.
Garcez, A.S., Nunez, S.C., Azambuja,
N. Jr., Fregnani, E.R., Rodriguez,
H.M.H., Hamblin, M.R., Suzuki, H.,
and Ribeiro, M.S. (2013) Effects of photodynamic therapy on gram-positive
and gram-negative bacterial biofilms
by bioluminescence imaging and scanning electron microscopic analysis.
Photomed. Laser Surg., 31, 519–525.
Biel, M.A., Pedigo, L., Gibbs, A., and
Loebel, N. (2013) Photodynamic therapy of antibiotic-resistant biofilms in
a maxillary sinus model. Int. Forum
Allergy Rhinology, 3, 468–473.
Haidaris, C.G., Foster, T.H., Waldman,
D.L., Mathes, E.J., McNamara, J., and
Curran, T. (2013) Effective photodynamic therapy against microbial
populations in human deep tissue
abscess aspirates. Lasers Surg. Med., 45,
509–516.
Nims, R.W. and Plavsic, M. (2013)
Polyomavirus inactivation – a review.
Biologicals, 41, 63–70.
McClaskey, J., Xu, M., Snyder, E.L.,
and Tormey, C.A. (2009) Clinical trials
for pathogen reduction in transfusion
medicine: a review. Transfus. Apher.
Sci., 41, 217–225.
Pelletier, J.P.R., Transue, S., and Snyder,
E.L. (2006) Pathogen inactivation techniques. Best Pract. Res. Clin. Haematol.,
19, 205–242.
Floyd, R.A., Schneider, J.E., and
Dittmer, D.R. (2004) Methylene blue
photoinactivation of RNA viruses.
Antiviral Res., 61, 141–151.
Sobral, P.M., Barros, A.E.d.L., Gomes,
A.M.A.S., and do Bonfim, C.V. (2012)
Viral inactivation in hemotherapy: systematic review on inactivators with
action on nucleic acids. Rev. Bras.
Hematol. Hemoter., 34, 231–235.
Kato, I.T., Prates, R.A., Sabino, C.P.,
Fuchs, B.B., Tegos, G.P., Mylonakis,
E., Hamblin, M.R., and Ribeiro, M.S.
References
172.
173.
174.
175.
176.
177.
178.
179.
180.
(2013) Antimicrobial photodynamic
inactivation inhibits candida albicans
virulence factors and reduces in vivo
pathogenicity. Antimicrob. Agents
Chemother., 57, 445–451.
Lyon, J.P., Moreira, L.M.,
Dutra de Carvalho, V.S., dos Santos,
F.V., de Lima, C.J., and de Resende,
M.A. (2013) In vitro photodynamic
therapy against foncecaea pedrosoi and
cladophialophora carrionii. Mycoses, 56,
157–161.
Lyon, J.P., Rezende, R.R., Rabelo, M.P.,
de Lima, C.J., and Moreira, L.M.
(2013) Synergic effect of photodynamic therapy with methylene blue
and surfactants in the inhibition of
Candida albicans. Mycopathologia, 175,
159–164.
Mehra, T., Borelli, C., Braunsdorf, C.,
Mailaender-Sanchez, D., Koeberle, M.,
and Schaller, M. (2013) Efficacy of
antifungal PACT therapy. Mycoses, 56,
13–13.
Goyal, M., Alam, A., and
Bandyopadhyay, U. (2012) Redox
regulation in malaria: current concepts
and pharmacotherapeutic implications.
Curr. Med. Chem., 19, 1475–1503.
Krafts, K., Hempelmann, E., and
Skorska-Stania, A. (2012) From methylene blue to chloroquine: a brief review
of the development of an antimalarial
therapy. Parasitol. Res., 111, 1–6.
Wainwright, M. and Amaral, L.
(2005) Review: the phenothiazinium
chromophore and the evolution of antimalarial drugs. Trop. Med. Int. Health,
10, 501–511.
Wainwright, M. (2004) Photoantimicrobials – a PACT against resistance and
infection. Drugs Future, 29, 85–93.
Wainwright, M., Smalley, H., and Flint,
C. (2011) The use of photosensitisers in acne treatment. J. Photochem.
Photobiol., B, 105, 1–5.
Gursoy, H., Ozcakir-Tomruk, C.,
Tanalp, J., and Yilmaz, S. (2013) Photodynamic therapy in dentistry: a
literature review. Clin. Oral Investig.,
17, 1113–1125.
181. Rajesh, S., Koshi, E., Philip, K., and
182.
183.
184.
185.
186.
187.
188.
189.
190.
191.
192.
Mohan, A. (2011) Antimicrobial photodynamic therapy: an overview. J. Indian
Soc. Periodontology, 15, 323–327.
Deepak, D., Urmi, D., and Neeraj,
D. (2012) Photodynamic therapy: a
view through light. J. Orofacial Res., 2,
82–86.
Sandler, S.G. (2010) The status of
pathogen-reduced plasma. Transfus.
Apher. Sci., 43, 393–399.
Seltsam, A. and Müller, T.H. (2013)
Update on the use of pathogen-reduced
human plasma and platelet concentrates. Br. J. Haematol., 162, 442–454.
Wainwright, M., Mohr, H., and Walker,
W.H. (2007) Phenothiazinium derivatives for pathogen inactivation in blood
products. J. Photochem. Photobiol., B,
86, 45–58.
Allain, J.P., Bianco, C., Blajchman,
M.A., Brecher, M.E., Busch, M.,
Leiby, D., Lin, L., and Stramer, S.
(2005) Protecting the blood supply
from emerging pathogens: the role of
pathogen inactivation. Transfus. Med.
Rev., 19, 110–126.
Wainwright, M. (2004) Photoinactivation of viruses. Photochem. Photobiol.
Sci., 3, 406–411.
Wainwright, M. and Baptista, M.S.
(2011) The application of photosensitisers to tropical pathogens in the blood
supply. Photodiagnosis Photodyn. Ther.,
8, 240–248.
Wainwright, M. (2002) The emerging
chemistry of blood product disinfection. Chem. Soc. Rev., 31, 128–136.
Wainwright, M. (2002) Pathogen inactivation in blood products. Curr. Med.
Chem., 9, 127–143.
Wainwright, M., Phoenix, D.A., Smillie,
T.E., and Wareing, D.R.A. (2001)
Phenothiaziniums as putative photobactericidal agents for red blood
cell concentrates. J. Chemother., 13,
503–509.
Steinmann, E., Gravemann, U.,
Friesland, M., Doerrbecker, J., Mueller,
T.H., Pietschmann, T., and Seltsam, A.
(2013) Two pathogen reduction technologiesmethylene blue plus light and
shortwave ultraviolet lighteffectively
319
320
10
193.
194.
195.
196.
197.
198.
199.
200.
201.
202.
Photodynamic Antimicrobial Chemotherapy
inactivate hepatitis C virus in blood
products. Transfusion, 53, 1010–1018.
Seghatchian, J. and Tolksdorf, F. (2012)
Characteristics of the THERAFLEX
UV-Platelets pathogen inactivation system – an update. Transfus. Apher. Sci.,
46, 221–229.
Mohammadi, Z., Palazzi, F., Giardino,
L., and Shalavi, S. (2013) Microbial
biofilms in endodontic infections: an
update review. Biomed. J., 36, 59–70.
Javed, F. and Romanos, G.E. (2013)
Does photodynamic therapy enhance
standard antibacterial therapy in dentistry? Photomed. Laser Surg., 31,
512–518.
Oruba, Z., Chomyszyn-Gajewska, M.,
and Macyk, W. (2012) Application of
the photodynamic therapy in medicine
and dentistry: literature review on
photodynamic and antimicrobial photodynamic therapy. BIODEVICES
2012 – Proceedings of the International Conference on Biomedical
Electronics and Devices 2012, pp.
190-195.
Vlahova, A.P., Kisov, C.K., Popova,
E.V., Haydushka, I.A., and Mantareva,
V.N. (2012) A new method for photodynamic disinfection of prosthetic
constructions and impressions in
prosthetic dentistry. Folia Med., 54,
51–57.
Sudhakara, R.R., Ramya, K., Ramesh,
T., Subbarayudu, G., Sai, M.N., and Sai,
K.C.H. (2012) Photodynamic therapy in
oral diseases. Int. J. Biol. Med. Res., 3,
1875–1883.
Guglielmi, C.d.A.B., Simionato, M.R.L.,
Ramalho, K.M., Imparato, J.C.P.,
Pinheiro, S.L., and Luz, M.A.A.C.
(2011) Clinical use of photodynamic
antimicrobial chemotherapy for the
treatment of deep carious lesions.
J. Biomed. Opt., 16, 088003–088003-7.
Pratten, J., Benhamou, V., and Street,
C. (2011) Antimicrobial Photodynamic
Therapy (aPDT) for Oral Infections.
Dumetriscu, A.L. (2010) Etiology and
Pathogenesis of Peridontal Disease,
Springer-Verlag, Berlin.
Street, C.N., Andersen, R., and Loebel,
N.G. (2009) Periowave (TM) demonstrates bactericidal activity against
203.
204.
205.
206.
207.
208.
209.
210.
periopathogens and leads to improved
clinical outcomes in the treatment of
adult periodontitis, in Optical Methods
for Tumor Treatment and Detection:
Mechanisms and Techniques in Photodynamic Therapy Xviii (ed. D.H.
Kessel), SPIE-International Society for
Optical Engineering, Bellingham.
Andersen, R., Loebel, N., Hammond,
D., and Wilson, M. (2007) Treatment
of periodontal disease by photodisinfection compared to scaling and root
planing. J. Clin. Dent., 18, 34–38.
Andersen, R.C., Loebel, N.G., and
Andersen, D.M. (2009) Meta-analysis
of 5 photodisinfection clinical trials for
periodontitis, in 12th World Congress of
the International Photodynamic Association: Photodynamic Therapy: Back
to the Future (ed. D.H. Kessel), SPIE,
Bellingham, WA.
Andersen, R.C., Loebel, N.G., and
Andersen, D.M. (2009) Photodynamic
dosimetry in the treatment of periodontitis, in 12th World Congress of
the International Photodynamic Association: Photodynamic Therapy: Back
to the Future (ed. D.H. Kessel), SPIE,
Bellingham, WA.
Berakdar, M., Callaway, A., Eddin, M.F.,
Ross, A., and Willershausen, B. (2012)
Comparison between scaling-rootplaning (SRP) and SRP/photodynamic
therapy: six-month study. Head Face
Med., 8, 12.
Ge, L., Shu, R., Li, Y., Li, C., Luo, L.,
Song, Z., Xie, Y., and Liu, D. (2011)
Adjunctive effect of photodynamic
therapy to scaling and root planing in
the treatment of chronic periodontitis.
Photomed. Laser Surg., 29, 33–37.
Benhamou, V. (2009) Photodisinfection:
the future of periodontal therapy. Dent.
Today, 28, 108–109.
Parker, K. (2009) Photodynamic antimicrobial chemotherapy in the general
dental practice. J. Laser Dent., 17,
131–138.
Perez, M., Rodriguez, B., Shintani, T.,
Watanabe, K.S.M., and Harrigan, R.
(2010) 5-Aminolevulinic Acid (5-ALA):
analysis of preclinical and safety literature. Food Nutr. Sci., 4, 1009–1013.
References
211. Sharma, S.K., Mroz, P., Dai, T.H.,
212.
213.
214.
215.
216.
217.
218.
219.
Huang, Y.Y., St Denis, T.G., and
Hamblin, M.R. (2012) Photodynamic
therapy for cancer and for infections:
what is the difference? Isr. J. Chem., 52,
691–705.
Luksiene, Z. and Brovko, L. (2013)
Antibacterial photosensitization-based
treatment for food safety. Food Eng.
Rev., 5, 185–199.
Smijs, T.G.M. and Pavel, S. (2011) The
susceptibility of dermatophytes to photodynamic treatment with special focus
on trichophyton rubrum. Photochem.
Photobiol., 87, 2–13.
Giomi, B., Mastrolorenzo, A., Tiradritti,
L., and Zuccati, G. (2012) Off label
treatments of genital warts: the role
of photodynamic therapy. G. Ital.
Dermatol. Venereol., 147, 467–474.
Sakamoto, F.H., Lopes, J.D., and
Anderson, R.R. (2010) Photodynamic
therapy for acne vulgaris: a critical
review from basics to clinical practice
Part I. Acne vulgaris: when and why
consider photodynamic therapy? J. Am.
Acad. Dermatol., 63, 183–193.
Dietel, W., Pottier, R., Pfister, W.,
Schleier, P., and Zinner, K. (2007) 5Aminolaevulinic acid (ALA) induced
formation of different fluorescent porphyrins: a study of the biosynthesis of
porphyrins by bacteria of the human
digestive tract. J. Photochem. Photobiol.,
B, 86, 77–86.
Hillemanns, P., Wang, X., Staehle,
S., Michels, W., and Dannecker, C.
(2006) Evaluation of different treatment
modalities for vulvar intraepithelial
neoplasia (VIN): CO2 laser vaporization, photodynamic therapy, excision
and vulvectomy. Gynecol. Oncol., 100,
271–275.
Oriel, S. and Nitzan, Y. (2010) Photoinactivation of Candida albicans by
its own endogenous porphyrins. Curr.
Microbiol., 60, 117–123.
Smith, T.G. and Kain, K.C. (2004) Inactivation of plasmodium falciparum by
photodynamic excitation of heme-cycle
intermediates derived from deltaaminolevulinic acid. J. Infect. Dis., 190,
184–191.
220. Komatsuya, K., Hata, M., Balogun, E.O.,
221.
222.
223.
224.
225.
226.
227.
Hikosaka, K., Suzuki, S., Takahashi, K.,
Tanaka, T., Nakajima, M., Ogura, S.-I.,
Sato, S., and Kita, K. (2013) Synergy of
ferrous ion on 5-aminolevulinic acid
mediated growth inhibition of plasmodium falciparum. J. Biochem., 154,
501–504.
Lee, Y. and Baron, E.D. (2011) Photodynamic therapy: current evidence
and applications in dermatology. Semin.
Cutan. Med. Surg., 30, 199–209.
van der Snoek, E.M., Robinson, D.J.,
van Hellemond, J.J., and Neumann,
H.A.M. (2008) A review of photodynamic therapy in cutaneous
leishmaniasis. J. Eur. Acad. Dermatol.
Venereol., 22, 918–922.
Akilov, O.E., Kosaka, S., O’Riordan, K.,
and Hasan, T. (2007) Parasiticidal effect
of delta-aminolevulinic acid-based
photodynamic therapy for cutaneous
leishmaniasis is indirect and mediated
through the killing of the host cells.
Exp. Dermatol., 16, 651–660.
Kosaka, S., Akilov, O.E., O’Riordan,
K., and Hasan, T. (2007) A mechanistic study of delta-aminolevulinic
acid-based photodynamic therapy
for cutaneous leishmaniasis. J. Invest.
Dermatol., 127, 1546–1549.
Zawislak, A., Donnelly, R.F.,
McCluggage, W.G., Price, J.H.,
McClelland, H.R., Woolfson, A.D.,
Dobbs, S., Maxwell, P., and McCarron,
P.A. (2009) Clinical and immunohistochemical assessment of vulval
intraepithelial neoplasia following
photodynamic therapy using a novel
bioadhesive patch-type system loaded
with 5-aminolevulinic acid. Photodiagnosis Photodyn. Ther., 6, 28–40.
Fabbrocini, G., Di Costanzo, M.P.,
Riccardo, A.M., Quarto, M., Colasanti,
A., Roberti, G., and Monfrecola, G.
(2001) Photodynamic therapy with
topical [delta]-aminolaevulinic acid
for the treatment of plantar warts.
J. Photochem. Photobiol., B, 61, 30–34.
Smucler, R. and Jatsova, E. (2005)
Comparative study of aminolevulic acid
photodynamic therapy plus pulsed dye
laser versus pulsed dye laser alone in
321
322
10
228.
229.
230.
231.
232.
233.
234.
235.
236.
Photodynamic Antimicrobial Chemotherapy
treatment of viral warts. Photomed.
Laser Surg., 23, 202–205.
Ayala, F., Grimaldi, E., Perfetto, B.,
Donnarumma, M., De Filippis, A.,
Donnarumma, G., and Tufano, M.A.
(2008) 5-aminolaevulinic acid and
photodynamic therapy reduce HSV1 replication in HaCat cells through
an apoptosis-independent mechanism. Photodermatol. Photoimmunol.
Photomed., 24, 237–243.
Li, X., Guo, H., Tian, Q.Z., Zheng,
G., Hu, Y.C., Fu, Y., and Tan, H.L.
(2013) Effects of 5-aminolevulinic
acid-mediated photodynamic therapy
on antibiotic-resistant staphylococcal
biofilm: an in vitro study. J. Surg. Res.,
184, 1013–1021.
Taylor, E. and Brown, S. (2002) The
advantages of aminolevulinic acid photodynamic therapy in dermatology.
J. Dermatol. Treat., 13, s3–s11.
Kostovic, K., Pastar, Z., Ceovic,
R., Mokos, Z.B., Buzina, D.S., and
Stanimirovic, A. (2012) Photodynamic therapy in dermatology: current
treatments and implications. Coll.
Anthropol., 36, 1477–1481.
Luksiene, Z. and Zukauskas, A. (2009)
Prospects of photosensitization in
control of pathogenic and harmful
micro-organisms. J. Appl. Microbiol.,
107, 1415–1424.
Warriner, K. and Namvar, A. (2009)
What is the hysteria with Listeria?
Trends Food Sci. Technol., 20, 245–254.
Castillo, A.D.C., Martinez, L.H.P., and
Apodaca, N.L.C. (2008) Salmonellosis and campylobacteriosis, the most
prevalent zoonosis in the world. Veterinaria Mexico, 39, 81–90.
Buchovec, I., Vaitonis, Z., and Luksiene,
Z. (2009) Novel approach to control
Salmonella enterica by modern biophotonic technology: photosensitization.
J. Appl. Microbiol., 106, 748–754.
Buchovec, I., Paskeviciute, E., and
Luksiene, Z. (2010) Photosensitizationbased inactivation of food pathogen
Listeria monocytogenes in vitro and
on the surface of packaging material.
J. Photochem. Photobiol. B, 99, 9–14.
237. Vaitonis, Z. and Luksiene, Z. (2010)
238.
239.
240.
241.
242.
243.
244.
245.
246.
Led-based light sources for decontamination of food: modelling
photosensitization-based inactivation of
pathogenic bacteria. Lith. J. Phys., 50,
141–145.
Luksiene, Z., Kokstaite, R., Katauskis,
P., and Skakauskas, V. (2013) Novel
approach to effective and uniform
inactivation of gram-positive listeria
monocytogenes and gram-negative
salmonella enterica by photosensitization. Food Technol. Biotechnol., 51,
338–344.
Stenfors Arnesen, L.P., Fagerlund, A.,
and Granum, P.E. (2008) From soil to
gut: bacillus cereus and its food poisoning toxins. FEMS Microbiol. Rev.,
32, 579–606.
Luksiene, Z., Buchovec, I., and
Paskeviciute, E. (2009) Inactivation
of food pathogen Bacillus cereus by
photosensitization in vitro and on the
surface of packaging material. J. Appl.
Microbiol., 107, 2037–2046.
Marc, Y.L., Buchovec, I., George, S.M.,
Baranyi, J., and Luksiene, Z. (2009)
Modelling the photosensitization-based
inactivation of Bacillus cereus. J. Appl.
Microbiol., 107, 1006–1011.
Lemon, K.P., Earl, A.M., Vlamakis,
H.C., Aguilar, C., and Kolter, R. (2008)
Biofilm development with an emphasis
on Bacillus subtilis. Bacterial Biofilms,
322, 1–16.
Brooks, J.D. and Flint, S.H. (2008)
Biofilms in the food industry: problems
and potential solutions. Int. J. Food Sci.
Technol., 43, 2163–2176.
Gandhi, M. and Chikindas, M.L. (2007)
Listeria: a foodborne pathogen that
knows how to survive. Int. J. Food
Microbiol., 113, 1–15.
Luksiene, Z., Danilcenko, H.,
Taraseviciene, Z., Anusevicius, Z.,
Maroziene, A., and Nivinskas, H. (2007)
New approach to the fungal decontamination of wheat used for wheat sprouts:
effects of aminolevulinic acid. Int. J.
Food Microbiol., 116, 153–158.
Memon, S., Hou, X., Wang, L., and
Li, Y. (2009) Promotive effect of 5aminolevulinic acid on chlorophyll,
References
247.
248.
249.
250.
251.
252.
253.
254.
255.
antioxidative enzymes and photosynthesis of Pakchoi (Brassica campestris
ssp. chinensis var. communis Tsen et
Lee). Acta Physiol. Plant., 31, 51–57.
Korkmaz, A. and Korkmaz, Y. (2009)
Promotion by 5-aminolevulenic acid of
pepper seed germination and seedling
emergence under low-temperature
stress. Sci. Hortic., 119, 98–102.
Zhang, W.F., Zhang, F., Raziuddin,
R., Gong, H.J., Yang, Z.M., Lu, L., Ye,
Q.F., and Zhou, W.J. (2008) Effects of
5-aminolevulinic acid on oilseed rape
seedling growth under herbicide toxicity stress. J. Plant Growth Regul., 27,
159–169.
Korkmaz, A., Korkmaz, Y., and
Demirkiran, A.R. (2010) Enhancing
chilling stress tolerance of pepper
seedlings by exogenous application of
5-aminolevulinic acid. Environ. Exp.
Bot., 67, 495–501.
Yiannikouris, A. and Jouany, J.P. (2002)
Mycotoxins in feeds and their fate
in animals: a review. Anim. Res., 51,
81–99.
Kumar, V., Basu, M.S., and Rajendran,
T.P. (2008) Mycotoxin research and
mycoflora in some commercially important agricultural commodities. Crop
Prot., 27, 891–905.
Reddy, K.R.N., Abbas, H.K., Abel,
C.A., Shier, W.T., Oliveira, C.A.F., and
Raghavender, C.R. (2009) Mycotoxin
contamination of commercially important agricultural commodities. Toxin
Rev., 28, 154–168.
Burge, S. and Wallis, D. (2010) Oxford
Handbook of Medical Dermatology,
Oxford University Press, Oxford.
Olejek, A., Kozak-Darmas, I.,
Kellas-Sleczka, S., Steplewska, K.,
Biniszkiewicz, T., Birkner, B., Jarek,
A., Nowak, L., Stencel-Gabriel, K.,
and Sieron, A. (2009) Effectiveness of
photodynamic therapy in the treatment of Lichen sclerosus: cell changes
in immunohistochemistry. Neuroendocrinol. Lett., 30, 547–551.
Sotiriou, E., Panagiotidou, D., and
Ioannidis, D. (2008) An open trial of
5-aminolevulinic acid photodynamic
therapy for vulvar lichen sclerosus. Eur.
256.
257.
258.
259.
260.
261.
262.
263.
264.
265.
J. Obstet. Gynecol. Reprod. Biol., 141,
187–188.
Zawislak, A.A., McCluggage, W.G.,
Donnelly, R.F., Maxwell, P., Price,
J.H., Dobbs, S.P., McClelland, H.R.,
Woolfson, A.D., and McCarron, P.A.
(2009) Response of vulval lichen
sclerosus and squamous hyperplasia to photodynamic treatment using
sustained topical delivery of aminolevulinic acid from a novel bioadhesive
patch system. Photodermatol. Photoimmunol. Photomed., 25, 111–113.
Eisendle, K. and Zelger, B. (2009) The
expanding spectrum of cutaneous borreliosis. G. Ital. Dermatol. Venereol.,
144, 157–171.
Hillemanns, P., Untch, M., Prove, F.,
Baumgartner, R., Hillemanns, M., and
Korell, M. (1999) Photodynamic therapy of vulvar lichen sclerosus with
5-aminolevulinic acid. Obstet. Gynecol.,
93, 71–74.
Gannon, M.J. and Brown, S.B. (1999)
Photodynamic therapy and its applications in gynaecology. Br. J. Obstet.
Gynaecol., 106, 1246–1254.
Horn, M. and Wolf, P. (2007) Topical
methyl aminolevulinate photodynamic
therapy for the treatment of folliculitis. Photodermatol. Photoimmunol.
Photomed., 23, 145–147.
Bernard, P. (2008) Management of
common bacterial infections of the
skin. Curr. Opin. Infect. Dis., 21,
122–128.
Taylor, M.N. and Gonzalez, M.L.
(2009) The practicalities of photodynamic therapy in acne vulgaris. Br. J.
Dermatol., 160, 1140–1148.
Arshdeep, A. (2013) What’s new in the
management of acne? Indian J. Dermatol. Venereol. Leprol., 79, 279–287.
Rai, R. and Natarajan, K. (2013) Laser
and light based treatments of acne.
Indian J. Dermatol. Venereol. Leprol.,
79, 300–309.
Ma, L., Xiang, L.-H., Yu, B., Yin, R.,
Chen, L., Wu, Y., Tan, Z.-J., Liu, Y.-B.,
Tian, H.-Q., Li, H.-Z., Lin, T., Wang,
X.-L., Li, Y.-H., Wang, W.-Z., Yang, H.L., and Lai, W. (2013) Low-dose topical
5-aminolevulinic acid photodynamic
therapy in the treatment of different
323
324
10
266.
267.
268.
269.
270.
271.
272.
273.
274.
Photodynamic Antimicrobial Chemotherapy
severity of acne vulgaris. Photodiagnosis
Photodyn. Ther., 10, 583–590.
Yang, G.-L., Zhao, M., Wang, J.-M., He,
C.-F., Luo, Y., Liu, H.-Y., Gao, J., Long,
C.-Q., and Bai, J.-R. (2013) Short-term
clinical effects of photodynamic therapy with topical 5-aminolevulinic acid
for facial acne conglobata: an open,
prospective, parallel-arm trial. Photodermatol. Photoimmunol. Photomed.,
29, 233–238.
Godal, A., Klaveness, J., and Morris,
H. (2013) Treatment of acne using
derivatives of 5-Aminolevulinic acid.
US Patent 8,546,447 B2, Photocure
ASA.
Andersson, R.R. (2012) Topical
aminolevulinic acid-photodynamic
therapy for the treatment of acne vulgaris. US Patent 2012/0296263 A1, The
General Hospital Corporation D/B/A
Massachusetts General Hospital.
Lomholt, H.B. and Kilian, M. (2008) Is
acne caused by colonization with the
“wrong” strain of Propionibacterium
acnes? A review of the role of propionibacterium acnes in acne. Ugeskr.
Laeger, 170, 1234–1237.
Bhate, K. and Williams, H.C. (2013)
Epidemiology of acne vulgaris. Br. J.
Dermatol., 168, 474–485.
Del Rosso, J.Q. and Kircik, L.H. (2013)
The sequence of inflammation, relevant biomarkers, and the pathogenesis
of acne vulgaris:what does recent
research show and what does it mean
to the clinician? J. Drugs Dermatol., 12,
S109–S115.
Itoh, Y., Ninomiya, Y., Tajima, S., and
Ishibashi, A. (2001) Photodynamic
therapy of acne vulgaris with topical
delta-aminolaevulinic acid and incoherent light in Japanese patients. Br. J.
Dermatol., 144, 575–579.
Itoh, Y., Ninomiya, Y., Tajima, S., and
Ishibashi, A. (2000) Photodynamic
therapy for acne vulgaris with topical
5-aminolevulinic acid. Arch. Dermatol.,
136, 1093–1095.
Hongcharu, W., Taylor, C.R., Chang,
Y.C., Aghassi, D., Suthamjariya, K.,
and Anderson, R.R. (2000) Topical
ALA-photodynamic therapy for the
275.
276.
277.
278.
279.
280.
281.
282.
283.
treatment of acne vulgaris. J. Invest.
Dermatol., 115, 183–192.
Ramberg, K., Melo, T.B., and Johnsson,
A. (2004) In situ assessment of
porphyrin photosensitizers in propionibacterium acnes. Z. Naturforsch. C,
59, 93–98.
Ashkenazi, H., Malik, Z., Harth, Y., and
Nitzan, Y. (2003) Eradication of propionibacterium acnes by its endogenic
porphyrins after illumination with high
intensity blue light. FEMS Immunol.
Med. Microbiol., 35, 17–24.
Fotinos, N., Mikulic, J., Convert, M.,
Campo, M.A., Piffaretti, J.C., Gurny,
R., and Lange, N. (2009) 5-ALA
derivative-mediated photoinactivation of propionibacterium acnes.
J. Dermatol. Sci., 56, 212–214.
Divaris, D.X.G., Kennedy, J.C., and
Pottier, R.H. (1990) Phototoxic damage
to sebaceous glands and hair-follicles
of mice after systemic administration
of 5-aminolevulinic acid correlates with
localized protoporphyrin-Ix fluorescence. Am. J. Pathol., 136, 891–897.
Sakamoto, F.H., Tannous, Z., Doukas,
A.G., Farinelli, W.A., Smith, N.A.,
Zurakowski, D., and Anderson, R.R.
(2009) Porphyrin distribution after
topical aminolevulinic acid in a novel
porcine model of sebaceous skin. Lasers
Surg. Med., 41, 154–160.
Kosaka, S., Kawana, S., Zouboulis, C.C.,
Hasan, T., and Ortel, B. (2006) Targeting of sebocytes by aminolevulinic
acid-dependent photosensitization.
Photochem. Photobiol., 82, 453–457.
Pollock, B., Turner, D., Stringer, M.R.,
Bojar, R.A., Goulden, V., Stables, G.I.,
and Cunliffe, W.J. (2004) Topical
aminolaevulinic acid-photodynamic
therapy for the treatment of acne vulgaris: a study of clinical efficacy and
mechanism of action. Br. J. Dermatol.,
151, 616–622.
Bryld, L.E. and Jemec, G.B.E. (2006)
The bacterial flora of the skin surface following routine MAL-PDT.
J. Dermatol. Treat., 17, 222–223.
Bisland, S.K., Chien, C., Wilson, B.C.,
and Burch, S. (2006) Pre-clinical in
vitro and in vivo studies to examine the
References
284.
285.
286.
287.
288.
289.
290.
291.
potential use of photodynamic therapy in the treatment of osteomyelitis.
Photochem. Photobiol. Sci., 5, 31–38.
Bisland, S.K., Chien, C., Wilson, B.C.,
and Burch, S. (2005) Pre-clinical in
vitro and in vivo studies to examine the
potential use of photodynamic therapy
in the treatment of osteomyelitis. Opt.
Methods Tumor Treat. Detect.: Mech.
Techn. Photodyn. Ther. XIV , 5689,
26–38.
Nitzan, Y. and Kauffman, M. (1999)
Endogenous porphyrin production
in bacteria by delta-aminolaevulinic
acid and subsequent bacterial photoeradication. Lasers Med. Sci., 14,
269–277.
Grinholc, M., Szramka, B., Olender,
K., and Graczyk, A. (2007) Bactericidal
effect of photodynamic therapy against
methicillin-resistant Staphylococcus
aureus strain with the use of various porphyrin photosensitizers. Acta
Biochim. Pol., 54, 665–670.
Karrer, S., Szeimies, R.M., Ernst, S.,
Abels, C., Baumler, W., and Landthaler,
M. (1999) Photodynamic inactivation
of staphylococci with 5-aminolaevulinic
acid or photofrin. Lasers Med. Sci., 14,
54–61.
Wiegell, S.R., Kongshoj, B., and Wulf,
H.C. (2006) Mycobacterium marinum
infection cured by photodynamic therapy. Arch. Dermatol., 142, 1241–1242.
Rallis, E. and Koumantaki-Mathioudaki,
E. (2007) Treatment of Mycobacterium
marinum cutaneous infections. Expert
Opin. Pharmacother., 8, 2965–2978.
Bruce-Micah, R., Huettenberger, D.,
Freitag, L., Cullum, J., and Foth, H.J.
(2009) Pharmacokinetic of ALA and hALA induced porphyrins in the models
Mycobacterium phlei and Mycobacterium smegmatis. J. Photochem.
Photobiol., B, 97, 1–7.
Wilder-Smith, C.H., Wilder-Smith,
P., Grosjean, P., van den Bergh, H.,
Woodtli, A., Monnier, P., Dorta, G.,
Meister, F., and Wagnieres, G. (2002)
Photoeradication of helicobacter pylori
using 5-aminolevulinic acid: preliminary human studies. Lasers Surg. Med.,
31, 18–22.
292. Ramstad, S., Futsaether, C.M., and
293.
294.
295.
296.
297.
298.
299.
300.
301.
302.
Johnsson, A. (1997) Porphyrin sensitization and intracellular calcium
changes in the prokaryote Propionibacterium acnes. J. Photochem. Photobiol.,
B, 40, 141–148.
Lee, C.F., Lee, C.J., Chen, C.T.,
and Huang, C.T. (2004) deltaAminolaevulinic acid mediated photodynamic antimicrobial chemotherapy
on Pseudomonas aeruginosa planktonic
and biofilm cultures. J. Photochem.
Photobiol., B, 75, 21–25.
Mitra, A. and Stables, G.I. (2006)
Topical photodynamic therapy for
non-cancerous skin conditions. Photodiagnosis Photodyn. Ther., 3, 116–127.
Nestor, M.S. (2007) The use of photodynamic therapy for treatment of acne
vulgaris. Dermatol. Clin., 25, 47–57.
Taub, A.F. (2007) Procedural treatments
for acne vulgaris. Dermatol. Surg., 33,
1005–1026.
Buggiani, G., Troiano, M., Rossi, R.,
and Lotti, T. (2008) Photodynamic
therapy: Off-label and alternative use in
dermatological practice. Photodiagnosis
Photodyn. Ther., 5, 134–138.
Orth, G., Mahy, B.W.J., and van
Regenmortel, M.H.V. (2008) Encyclopedia of Virology, Academic Press,
Oxford, pp. 8–18.
O’Riordan, K., Akilov, O.E., and Hasan,
T. (2005) The potential for photodynamic therapy in the treatment of
localized infections. Photodiagnosis
Photodyn. Ther., 2, 247–262.
Schroeter, C.A., Kaas, L., Waterval,
J.J., Bos, P.M., and Neumann, H.A.M.
(2007) Successful treatment of periungual warts using photodynamic therapy:
a pilot study. J. Eur. Acad. Dermatol.
Venereol., 21, 1170–1174.
Yu, Y.E., Kuohung, V., Penrose, C.,
Shim, H., and Gilchrest, B.A. (2012)
Letter: photodynamic therapy for treatment of hand warts. Dermatol. Surg.,
38, 818–820.
Yoo, K.H., Kim, B.J., and Kim, M.N.
(2009) Enhanced efficacy of photodynamic therapy with methyl
5-aminolevulinic acid in recalcitrant
periungual warts after ablative carbon
325
326
10
303.
304.
305.
306.
307.
308.
309.
Photodynamic Antimicrobial Chemotherapy
dioxide fractional laser: a pilot study.
Dermatol. Surg., 35, 1927–1932.
Liang, J., Lu, X.N., Tang, H., Zhang,
Z., Fan, J., and Xu, J.H. (2009) Evaluation of photodynamic therapy using
topical aminolevulinic acid hydrochloride in the treatment of condylomata
acuminata: a comparative, randomized
clinical trial. Photodermatol. Photoimmunol. Photomed., 25, 293–297.
Li, X.L., Wang, X.X., Gu, J.Y., Ma, Y.E.,
Liu, Z.Y., and Shi, Y.L. (2013) Needlefree injection of 5-aminolevulinic acid
in photodynamic therapy for the treatment of condylomata acuminata. Exp.
Ther. Med., 6, 236–240.
Lu, Y.-G., Yang, Y.-D., Wu, J.-J., Lei,
X., Cheng, Q.-H., He, Y., and Yang, W.
(2012) Treatment of perianal condyloma acuminate with topical ALA-PDT
combined with curettage: outcome
and safety. Photomed. Laser Surg., 30,
186–190.
Ying, Z., Li, X., and Dang, H. (2013)
5-aminolevulinic acid-based photodynamic therapy for the treatment of
condylomata acuminata in Chinese
patients: a meta-analysis. Photodermatol. Photoimmunol. Photomed., 29,
149–159.
Szeimies, R.-M., Schleyer, V., Moll,
I., Stocker, M., Landthaler, M., and
Karrer, S. (2009) Adjuvant photodynamic therapy does not prevent
recurrence of condylomata acuminata
after carbon dioxide laser ablation –
a phase III, prospective, randomized,
bicentric, double-blind study. Dermatol.
Surg., 35, 757–764.
Soergel, P., Loning, M., Staboulidou, I.,
Schippert, C., and Hillemanns, P. (2008)
Photodynamic diagnosis and therapy in
gynecology. J. Environ. Pathol. Toxicol.
Oncol., 27, 307–320.
Ekonjo, G.B., Saleh, Y., Kasiak, J.,
Gryboś, M., Teterycz, E., Korzeniewski,
J., Siewiński, M., Da˛browski, A., and
Małgorzata, S. (2006) In vivo application of 5-aminolevulinic acid in
the treatment of papillomavirus
infection in women with cervical
lesions after detection and genotyping
using PCR techniqueI. Int. J. Gynecol.
Obstetrics, 6.
310. Wierrani, F., Kubin, A., Jindra, R.,
311.
312.
313.
314.
315.
316.
317.
Henry, M., Gharehbaghi, K., Grin,
W., Soltz-Szotz, J., Alth, G., and
Grunberger, W. (1999) 5-aminolevulinic
acid-mediated photodynamic therapy of intraepithelial neoplasia and
human papillomavirus of the uterine
cervix – a new experimental approach.
Cancer Detect. Prev., 23, 351–355.
Yamaguchi, S., Tsuda, H., Takemori,
M., Nakata, S., Nishimura, S.,
Kawamura, N., Hanioka, K., Inoue,
T., and Nishimura, R. (2005) Photodynamic therapy for cervical
intraepithelial neoplasia. Oncology, 69,
110–116.
Soergel, P., Wang, X., Stepp, H.,
Hertel, H., and Hillemanns, P. (2008)
Photodynamic therapy of cervical
intraepithelial neoplasia with hexaminolevulinate. Lasers Surg. Med., 40,
611–615.
Winters, U., Daayana, S., Lear, J.T.,
Tomlinson, A.E., Elkord, E., Stern,
P.L., and Kitchener, H.C. (2008) Clinical and immunologic results of a
phase II trial of sequential imiquimod
and photodynamic therapy for vulval
intraepithelial neoplasia. Clin. Cancer
Res., 14, 5292–5299.
Keefe, K.A., Tadir, Y., Tromberg, B.,
Berns, M., Osann, K., Hashad, R.,
and Monk, B.J. (2002) Photodynamic
therapy of high-grade cervical intraepithelial neoplasia with 5-aminolevulinic
acid. Lasers Surg. Med., 31, 289–293.
McCarron, P.A., Donnelly, R.F., and
Woolfson, A.D. (2008) Autoradiographic and scintillation analysis of
5-aminolevulinic acid permeation
through epithelialised tissue: implications for topical photodynamic therapy
of superficial gynaecological neoplasias.
Pharm. Res., 25, 812–826.
Gormley, R.H. and Kovarik, C.L. (2009)
Dermatologic manifestations of HPV
in HIV-infected individuals. Curr.
HIV/AIDS Rep., 6, 130–138.
Beachler, D.C. and D’Souza, G. (2013)
Oral human papillomavirus infection
and head and neck cancers in HIVinfected individuals. Curr. Opin. Oncol.,
25, 503–510.
References
318. Lissouba, P., Van de Perre, P., and
319.
320.
321.
322.
323.
324.
325.
326.
327.
Auvert, B. (2013) Association of genital
human papillomavirus infection with
HIV acquisition: a systematic review
and meta-analysis. Sex. Transm. Infect.,
89, 350–356.
Smetana, Z., Malik, Z., Orenstein, A.,
Mendelson, E., and BenHur, E. (1997)
Treatment of viral infections with 5aminolevulinic acid and light. Lasers
Surg. Med., 21, 351–358.
Orth, G. (2008) Host defenses against
human papillomaviruses: lessons from
epidermodysplasia verruciformis.
Curr. Top. Microbiol. Immunol., 321,
59–83.
Szeimies, R.M., Landthaler, M., and
Karrer, S. (2002) Non-oncologic indications for ALA-PDT. J. Dermatolog.
Treat., 13 (Suppl. 1), S13–S18.
Zampetti, A., Giurdanella, F., Manco,
S., Linder, D., Gnarra, M., Guerriero,
G., and Feliciani, C. (2013) Acquired
epidermodysplasia verruciformis: a
comprehensive review and a proposal
for treatment. Dermatol. Surg., 39,
974–980.
Karrer, S., Szeimies, R.M., Abels, C.,
Wlotzke, U., Stolz, W., and Landthaler,
M. (1999) Epidermodysplasia verruciformis treated using topical
5-aminolaevulinic acid photodynamic therapy. Br. J. Dermatol., 140,
935–938.
Sunohara, M., Ozawa, T., Morimoto,
K., Harada, T., Ishii, M., and Fukai, K.
(2012) Dye laser photodynamic therapy
for Bowen’s disease in a patient with
epidermodysplasia verruciformis. Osaka
City Med. J., 58, 77–82.
Berk, D.R., Bruckner, A.L., and
Lu, D. (2009) Epidermodysplasia
verruciform-like Lesions in an HIV
patient. Dermatol. Online J., 15, 1.
Rogers, H.D., Macgregor, J.L., Nord,
K.M., Tyring, S., Rady, P., Engler,
D.E., and Grossman, M.E. (2009)
Acquired epidermodysplasia verruciformis. J. Am. Acad. Dermatol., 60,
315–320.
Handisurya, A., Schellenbacher, C., and
Kirnbauer, R. (2009) Diseases caused
by human papillomaviruses (HPV).
328.
329.
330.
331.
332.
333.
334.
J. Dtsch. Dermatol. Ges., 7, 453–466;
quiz 466, 467.
Wang, X.-L., Wang, H.-W., Yuan, K.-H.,
Li, F.-L., and Huang, Z. (2011) Combination of photodynamic therapy and
immunomodulation for skin diseasesupdate of clinical aspects. Photochem.
Photobiol. Sci., 10, 704–711.
Wang, X.-L., Wang, H.-W., and Huang,
Z. (2008) Combination of photodynamic therapy and immunotherapy
- evolving role in dermatology.
Proc. SPIE., 6857, 85704–85704,
http://spie.org/Publications/Proceedings/
Paper/10.1117/12.760769 (accessed 11
July 2014).
Wang, X.-L., Wang, H.-W., Guo, M.-X.,
and Huang, Z. (2007) Combination
of immunotherapy and photodynamic
therapy in the treatment of Bowenoid
papulosis. Photodiagnosis Photodyn.
Ther., 4, 88–93.
Wang, X.-L., Wang, H.-W., and
Guo, M.-X. (2007) Imiquimod
immunotherapy and ALA photodynamic therapy combination for the
treatment of genital bowenoid papulosis. Proc. SPIE., 6438, 43809–43809,
http://spie.org/Publications/Proceedings/
Paper/10.1117/12.698774 (accessed 11
July 2014).
Tao, J.-n., Duan, S.-m., and Li, J. (2007)
Experimental studies on treatment of
HSV infections with photodynamic
therapy using 5-aminolevulinic acid.
Zhonghua Shi Yan He Lin Chuang Bing
Du Xue Za Zhi, 21, 79–82.
Kvacheva, Z.B., Lobanok, Y.S.,
Votyakov, V.I., Shukanova, N.A.,
Vorobey, A.V., Nikolayeva, S.N., and
Titov, L.P. (2005) Photodynamic inhibition of infection caused by herpes
simplex virus type 1 in the cultured
cells, by using 5-aminolevulinic acidinduced porphyrins. Vopr. Virusol., 50,
44–47.
Donnelly, R.F., McCarron, P.A., Ma,
L.W., Juzenas, P., Iani, V., Woolfson,
A.D., Zawislak, A.A., and Moan, J.
(2006) Facilitated delivery of ALA
to inaccessible regions via bioadhesive patch systems. J. Environ. Pathol.
Toxicol. Oncol., 25, 389–402.
327
328
10
Photodynamic Antimicrobial Chemotherapy
335. Wang, X.-L., Wang, H.-W., Zhang, L.-
336.
337.
338.
339.
340.
341.
342.
343.
L., Guo, M.-X., Su, L.-N., and Tao, J.-N.
(2008) 5-aminolevulinic acid-mediated
photodynamic diagnosis for condylomata acuminata. Zhonghua Pifuke
Zazhi., 41, 296–300.
Wang, X.L., Wang, H.W., Huang, Z.,
Stepp, H., Baumgartner, R., Dannecker,
C., and Hillemanns, P. (2007) Study
of protoporphyrin IX (PpIX) pharmacokinetics after topical application
of 5-aminolevulinic acid in urethral
condylomata acurninata. Photochem.
Photobiol., 83, 1069–1073.
Gold, M.H. and Moiin, A. (2007)
Treatment of verrucae vulgaris and
molluscum contagiosum with photodynamic therapy. Dermatol. Clin., 25,
75–80.
Sohl, S., Kauer, F., Paasch, U., and
Simon, J.C. (2007) Photodynamic
treatment of cutaneous leishmaniasis.
J. Dtsch. Dermatol. Ges., 5, 128–130.
Asilian, A. and Davami, M. (2006)
Comparison between the efficacy of
photodynamic therapy and topical
paromomycin in the treatment of
Old World cutaneous leishmaniasis: a
placebo-controlled, randomized clinical trial. Clin. Exp. Dermatol., 31,
634–637.
Ghaffarifar, F., Jorjani, O., Mirshams,
M., Miranbaygi, M.H., and Hosseini,
Z.K. (2006) Photodynamic therapy as
a new treatment of cutaneous leishmaniasis. East. Mediterr. Health J., 12,
902–908.
Evangelou, G., Krasagakis, K.,
Giannikaki, E., Kruger-Krasagakis,
S., and Tosca, A. (2011) Successful
treatment of cutaneous leishmaniasis
with intralesional aminolevulinic acid
photodynamic therapy. Photodermatol. Photoimmunol. Photomed., 27,
254–256.
Gardlo, K., Horska, Z., Enk, C.D.,
Rauch, L., Megahed, M., Ruzicka, T.,
and Fritsch, C. (2003) Treatment of
cutaneous leishmaniasis by photodynamic therapy. J. Am. Acad. Dermatol.,
48, 893–896.
Enk, C.D., Fritsch, C., Jonas, F.,
Nasereddin, A., Ingber, A., Jaffe, C.L.,
and Ruzicka, T. (2003) Treatment of
344.
345.
346.
347.
348.
349.
350.
cutaneous leishmaniasis with photodynamic therapy. Arch. Dermatol., 139,
432–434.
Dutta, S., Furuyama, K., Sassa, S.,
and Chang, K.-P. (2008) Leishmania
spp.: Delta-aminolevulinate-inducible
neogenesis of porphyria by genetic
complementation of incomplete heme
biosynthesis pathway. Exp. Parasitol.,
118, 629–636.
Sah, J.F., Ito, H., Kolli, B.K., Peterson,
D.A., Sassa, S., and Chang, K.-P. (2002)
Genetic rescue of leishmania deficiency in porphyrin biosynthesis creates
mutants suitable for analysis of cellular
events in uroporphyria and for photodynamic therapy. J. Biol. Chem., 277,
14902–14909.
Dutta, S., Chang, C., Kolli, B.K., Sassa,
S., Yousef, M., Showe, M., Showe,
L., and Chang, K.P. (2012) Deltaaminolevulinate-induced host-parasite
porphyric disparity for selective photolysis of transgenic leishmania in the
phagolysosomes of mononuclear phagocytes: a potential novel platform for
vaccine delivery. Eukaryot. Cell, 11,
430–441.
Kumari, S., Samant, M., Khare, P.,
Misra, P., Dutta, S., Kolli, B.K., Sharma,
S., Chang, K.P., and Dube, A. (2009)
Photodynamic vaccination of hamsters with inducible suicidal mutants of
Leishmania amazonensis elicits immunity against visceral leishmaniasis. Eur.
J. Immunol., 39, 178–191.
Lee, J.W., Kim, B.J., and Kim, M.N.
(2010) Photodynamic therapy: new
treatment for recalcitrant malassezia
folliculitis. Lasers Surg. Med., 42,
192–196.
Lee, J.W., Lee, H.I., Kim, M.N., Kim,
B.J., Chun, Y.-J., and Kim, D. (2011)
Topical photodynamic therapy with
methyl aminolevulinate may be an
alternative therapeutic option for the
recalcitrant Malassezia folliculitis. Int. J.
Dermatol., 50, 488–490.
Kim, Y.J. and Kim, Y.C. (2007) Successful treatment of pityriasis versicolor
with 5-aminolevulinic acid photodynamic therapy. Arch. Dermatol., 143,
1218–1220.
References
351. Crespo-Erchiga, V. and Florencio, V.D.
352.
353.
354.
355.
356.
357.
358.
359.
360.
(2006) Malassezia yeasts and pityriasis
versicollor. Curr. Opin. Infect. Dis., 19,
139–147.
Monfrecola, G., Procaccini, E.M.,
Bevilacqua, A., Manco, A., Calabro, G.,
and Santoianni, P. (2004) In vitro effect
of 5-aminolaevulinic acid plus visible
light on Candida albicans. Photochem.
Photobiol. Sci., 3, 419–422.
Sukhih, G., Apolikhina, I., Aslanyan,
K., Teterina, T., Ankirskaya, A.,
Vorozhtsov, G., and Kuzmin, S. (2011)
Photodynamic Therapy – Alternative
Method in Treatment of Vulvovaginal
Candidiasis in Women, Medimond S R
L, Bologna.
Calzavara-Pinton, P.G. and Venturini,
M. (2006) Photodynamic Therapy of
Skin Infections, Medimond S R L,
Bologna.
Egusa, H., Soysa, N.S., Ellepola, A.N.,
Yatani, H., and Samaranayake, L.P.
(2008) Oral candidosis in HIVinfected patients. Curr. HIV Res., 6,
485–499.
Strakhovskaya, M.G., Shumarina, A.O.,
Fraikin, G.Y., and Rubin, A.B. (1999)
Endogenous porphyrin accumulation
and photosensitization in the yeast
Saccharomyces cerevisiae in the presence of 2,2′ -dipyridyl. J. Photochem.
Photobiol., B, 49, 18–22.
Strakhovskaya, M.G., Shumarina,
A.O., Fraikin, Y.G., and Rubin, A.B.
(1998) Synthesis of protoporphyin
IX induced by 5-aminolevulinic
acid in yeast cells in the presence of
2,2-dipyridyl. Biochem. Moscow, 63,
859–863.
Calzavara-Pinton, P.G., Venturini, M.,
and Sala, R. (2005) A comprehensive
overview of photodynamic therapy
in the treatment of superficial fungal
infections of the skin. J. Photochem.
Photobiol., B, 78, 1–6.
Donnelly, R.F., McCarron, P.A., and
Tunney, M.M. (2008) Antifungal photodynamic therapy. Microbiol. Res., 163,
1–12.
Kamp, H., Tietz, H.J., Lutz, M., Piazena,
H., Sowyrda, P., Lademann, J., and
Blume-Peytavi, U. (2005) Antifungal
effect of 5-aminolevulinic acid PDT
361.
362.
363.
364.
365.
366.
367.
368.
in Trichophyton rubrum. Mycoses, 48,
101–107.
Sotiriou, E., Panagiotidou, D., and
Ioannides, D. (2009) 5-Aminolevulininic
acid photodynamic therapy treatment for tinea cruris caused by
Trichophyton rubrum: report of 10
cases. J. Eur. Acad. Dermatol. Venereol.,
23, 341–342.
Sotiriou, E., Koussidou-Ermonti,
T., Chaidemenos, G., Apalla, Z.,
and Ioannides, D. (2010) Photodynamic therapy for distal and lateral
subungual toenail onychomycosis
caused by trichophyton rubrum: preliminary results of a single-centre
open trial. Acta Derm. Venereol., 90,
216–217.
Watanabe, D., Kawamura, C.,
Masuda, Y., Akita, Y., Tamada, Y., and
Matsumoto, Y. (2008) Successful treatment of toenail onychomycosis with
photodynamic therapy. Arch. Dermatol.,
144, 19–21.
Piraccini, B.M., Rech, G., and Tosti,
A. (2008) Photodynamic therapy of
onychomycosis caused by Trichophyton
rubrum. J. Am. Acad. Dermatol., 59,
S75–S76.
Donnelly, R.F., McCarron, P.A.,
Lightowler, J.M., and Woolfson, A.D.
(2005) Bioadhesive patch-based delivery of 5-aminolevulinic acid to the
nail for photodynamic therapy of onychomycosis. J. Control. Release, 103,
381–392.
Calzavara-Pinton, P.G., Venturini, M.,
Capezzera, R., Sala, R., and Zane,
C. (2004) Photodynamic therapy of
interdigital mycoses of the feet with
topical application of 5-aminolevulinic
acid. Photodermatol. Photoimmunol.
Photomed., 20, 144–147.
Sotiriou, E., Koussidou, T., Patsatsi, A.,
Apalla, Z., and Ioannides, D. (2009)
5-Aminolevulinic acid-photodynamic
treatment for dermatophytic tinea pedis
of interdigital type: a small clinical
study. J. Eur. Acad. Dermatol. Venereol.,
23, 203–204.
Qiao, J., Li, R., Ding, Y., and Fang,
H. (2010) Photodynamic therapy in
the treatment of superficial mycoses:
329
330
10
Photodynamic Antimicrobial Chemotherapy
an evidence-based evaluation. Mycopathologia, 170, 339–343.
369. Gupta, A.K., Chaudhry, M., and
Elewski, B. (2003) Tinea corporis,
tinea cruris, tinea nigra, and piedra.
Dermatol. Clin., 21, 395–400.
370. Mendoza, N., Arora, A., Arias, C.A.,
Hernandez, C.A., Madkam, V., Tyring,
S.K., Elias, J.A., Michael, R.M., and
Michael, A.P. (2009) Clinical Mycology, 2nd edn, Churchill Livingstone,
Edinburgh, pp. 509–523.
371. McCarthy, D.J. (2004) Origins of onychomycosis. Clin. Podiatr. Med. Surg.,
21, 533–553.
372. Riddle, C.C., Terrell, S.N., Menser,
M.B., Aires, D.J., and Schweiger, E.S.
(2009) A review of Photodynamic
Therapy (PDT) for the treatment of
acne vulgaris. J. Drugs Dermatol., 8,
1010–1019.
373. Issa, M.C.A. and Manela-Azulay, M.
(2010) Photodynamic therapy: a review
of the literature and image documentation. An. Bras. Dermatol., 85,
501–511.
331
11
The Antimicrobial Effects of Ultrasound
Frederick Harris, Sarah R. Dennison, and David A. Phoenix
11.1
Introduction
Human interest in ultrasound has a long and international history [1–4] with the
first detailed experiments performed in Italy by Spallanzani in the 1770s [5] indicating that non-audible sound might exist. In an attempt to explain the ability
of bats to navigate flight in darkness, this author demonstrated that when blindfolded, these creatures could navigate but when their mouths were covered, they
collided with obstacles [6]. These observations became known as Spallanzani’s bat
problem and remained a scientific mystery until the middle of the twentieth century, when the use of a sonic detector by Donald and Griffin recorded directional
ultrasound noises being emitted by bats in navigating flight [7]. In Switzerland in
1826, Colladon [8] discovered that sound traveled faster in water than in air by
sounding a Church bell underwater and measuring the speed of the sound with
a rear trumpet. This experimental system essentially formed the basis of modern
transducers, and Colladon is generally regarded as the founder of ultrasonography [9]. Later that century, in the 1880s, Pierre Curie discovered that an electrical
voltage could be produced by pressure on some types of crystals [10, 11], which
are now better known as piezoelectric crystals and serve as the basis for modern
ultrasound transducers that transform sound pressure into a measurable voltage
[12]. However, the modern era of research into potential uses of ultrasound is generally taken to begin with work by Langevin in France [13] who in 1917 reported
the death of fish near an ultrasound source during the development of sonar to
detect submarines in the first World War [14]. A decade later in the United States,
Woods and Loomis [15] studied the lethal effect of high-power ultrasound on
a variety of organisms, including fish, frogs, and mice. Subsequent examination
revealed intra-abdominal bleeding but with vibrations of lower intensity; tissue
destruction was reduced. These latter authors are often credited as the first to
report the biological effects of this form of radiation on living cells and tissues
[16]. In Germany, in 1938, Ziess [17] studied the effects of ultrasound on the eye,
which is usually taken to be the start of the field of ophthalmic ultrasonography
[18]. Four years later in Austria, Dussik attempted to locate brain tumors and the
Novel Antimicrobial Agents and Strategies, First Edition.
Edited by David A. Phoenix, Frederick Harris and Sarah R. Dennison.
© 2015 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2015 by Wiley-VCH Verlag GmbH & Co. KGaA.
332
11
The Antimicrobial Effects of Ultrasound
cerebral ventricles by measuring the transmission of ultrasound beams through
the head [19, 20], which marks the start ultrasound use as a medical diagnostic
tool [1]. In the same year in the United States, Lynn et al. [21] built and tested a
high-power, focused, ultrasound transducer and produced lesions deep in bovine
liver without damaging the surrounding tissue and on the basis of this work, these
authors are generally recognized as the inventors of focused ultrasound therapy
[22]. In 1944, Lynn and Putnam [23] proposed using ultrasound to destroy tissue, which led to a series of studies over the 1950s by Fry et al. [24–26] in the
United States who investigated the effect of high-power ultrasound on the central nervous system in order to develop a modality for the noninvasive surgery of
diseased tissue. Toward the end of the 1950s in the United States, these investigations led to clinical trials and the treatment of patients with Parkinson’s disease
and other neurological conditions [27]. In a series of classic studies during the
1950s in Great Britain, Wild laid the foundations of ultrasonic tissue diagnosis
with the publication of amplitude mode (A-mode) ultrasound investigations of
surgical specimens of intestinal and breast malignancies, the development of a
linear handheld brightness mode (B-mode) instrument and early descriptions of
endoscopic (transrectal and transvaginal) A-mode scanning transducers in [20,
28]. In 1956, Burov and Andreevskaya [29, 30] in the USSR tested high-power
ultrasound on tumors and reported an enhanced immunological effect, which led
to the suggestion that this form of radiation had the potential for the treatment of
cancer. Two years later in Scotland, having first realized the diagnostic potential of
ultrasound in the Glasgow shipyards where it was to look for flaws in metallurgy,
Donald [31] published “Investigation of Abdominal Masses by Pulsed Ultrasound”
in the Lancet, which became one of the defining publications in the field of medical ultrasound [32]. This and other work paved the way for the use of diagnostic
ultrasound in gynecology and this technique is now an important aid to diagnosing fetal progress during pregnancy [32, 33]. From these many pioneering studies,
interest in ultrasound has expanded greatly and there has been extensive evaluation of its use for a variety of medical purposes, which currently, range from bone
cutting to promoting drug delivery (Table 11.1).
One aspect of ultrasound that has received far less attention than many of its
other applications is the use of its antimicrobial activity (Table 11.1) and it is an
often forgotten fact that some of the earliest experimental work on the biological
effects of this radiation included studies on its ability to kill microorganisms
[15, 48]. In this chapter, we review current research into the antimicrobial
capability of ultrasound either alone or in combination with a variety of other
antimicrobial strategies.
11.2
The Antimicrobial Activity of Ultrasound Alone
The history of research into the antimicrobial capability of ultrasound can be considered as composed of three periods of which the first lasted until the early 1990s
11.2
The Antimicrobial Activity of Ultrasound Alone
Table 11.1 Major therapeutic uses of ultrasound.
Power level
Application of ultrasound
High or intermediate power
Dentistry and surgery (scalpel and
bone cutting)
Synthesis of microcapsules for drug
delivery
Sonodynamic disinfection
Separation technology
Physiotherapy
Destruction of blood clots
Transdermal
drug
delivery
(sonophoresis)
Sonodynamic therapy (SDT)
Sonodynamic
antimicrobial
chemotherapy (SACT)
Sonodynamic antibiotic therapy
High intensity focused ultrasound
Improved cellular uptake of molecules
(sonoporation)
Diagnostic imaging
Intermediate or low power
Low power
References
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
This table shows a range of therapeutic applications of high, intermediate, and low power
ultrasound.
This table was adapted from Ref. [47].
and was mainly concerned with the direct action of the radiation on microorganisms [49]. The first major investigation into this action is generally taken to
be that of Wood and Loomis [15] who, in 1927, reported the destruction of the
algal microorganism, Spirogyra, when subjected to high-power ultrasound. Similar observations were presented over the next few years for algae such as Euglena,
Colpidium, Nitella, and Paramecium, along with bacteria, such as the luminescent Bacillus fischeri, now known as Aliivibrio fischeri [48, 50, 51]. As an example,
in the case of Spirogyra, it was observed by Wood and Loomis [15] in 1927 that
“Filaments of Spirogyra in suspension were torn to pieces and the cells ruptured”
and later by Harvey [48] in 1930 that “ … Spirogyra, there is the break-up of
the chlorophyll spirals with loss of turgor, the chlorophyll becoming diffuse in
many cells”. It is now known that the exposure of Spirogyra to ultrasound induces
the disruption of connections between the plasmalemma and the algal cell walls,
which then leads to the loss of membrane integrity, the leakage of cytoplasmic contents, a decline in the chlorophyll content of cells and eventually, the collapse of the
cell into a dense brown mass [52]. Since these earlier studies, it has been demonstrated that high-power ultrasound possesses efficacy in inactivating a range of
microorganisms [53], including other algae [54, 55] and bacteria [55–62] along
with protozoa [61, 63–65], fungi [55, 56, 66–70], and viruses [56, 61, 71–73].
333
334
11
The Antimicrobial Effects of Ultrasound
It is generally accepted that the major antimicrobial effects produced by
the direct application of ultrasound are due to acoustic cavitation [49, 53],
which occurs when the incident sonic wave meets a liquid medium and creates
longitudinal waves characterized by regions of alternating compression and
expansion [74, 75]. These regions of pressure change cause cavitation to occur
and gas bubbles to form in the medium, which can have different consequences
on microbial cells depending on the stability of these bubbles [76–78]. The cycles
of low and high acoustic pressure in the incident acoustic wave cause the gas
bubbles to expand and contract, which, in turn, creates shear flow around the
oscillating bubbles [76, 77, 79, 80]. These bubbles have a larger surface area during
the expansion cycle, which increases the diffusion of gas, causing the bubble
to expand [74, 75]. Stable cavitation results when power levels are sufficiently
low so that the bubbles do not collapse violently during their contraction cycle
[77, 81]. However, at higher power levels, collapse (or transient) cavitation is
produced, which can cause the oscillating bubbles to violently accelerate the fluid
around them. During bubble collapse, the rapid condensation of gas molecules
within bubbles occurs, reducing their radii to very small values at the end of the
contraction cycle. Violent collisions between these condensed molecules and
the sudden reversal in the motion of the bubble wall can create regions of very
high pressure and temperature [74, 75]. This process is adiabatic and it has been
estimated that, on a nanosecond time scale, the strength of gas compression in
the process can induce temperatures of up to 15 000 K and pressures in excess of
1000 MPa [82, 83] although conditions nearer 5000 K and 250 MPa are those most
frequently cited for collapse cavitation events [84]. These extreme conditions can
generate high liquid shear force, shock waves, localized heating, and free radicals,
all of which would appear to contribute to the antimicrobial effect of ultrasound
by disrupting cellular, structural, and functional components of microbial cells
[54, 85–88].
As described above, high-power ultrasound radiation has been shown to be
able to inactivate a variety of microorganisms [53] and there has been some
research into this ability in relation to controlling algae growth in ponds, lakes,
and reservoirs, and eliminating biofilm formation by these microbes inside water
treatment plants [89]. However, most research into the direct antimicrobial
action of this radiation has centered on bacterial targets, including Gramnegative organisms, such as Escherichia coli, and Gram-positive bacteria, such
as Staphylococcus aureus [49, 55–62, 90–96]. There have been two major foci of
this research, of which, the first is biofilms and agglomerates of bacteria, which
are also encountered in water treatment plants, exemplified by Pseudomonas
aeruginosa and Bacillus subtilis, respectively [49, 54, 97, 98]. The second major
focus of ultrasound-mediated antimicrobial research is bacterial biofilms found
in the food industries, typically formed by Listeria monocytogenes and Salmonella
typhimurium, respectively [86, 87, 91, 93, 99]. In combination, these studies
have suggested that the occurrence of a sufficient number of collapse cavitation
events can induce the antibacterial effects of ultrasound, which can be divided
into a number of stages [60]: (i) Collapse cavitation induces forces due to surface
11.3
The Antimicrobial Activity of Assisted Ultrasound
resonance in the bacterial cell along with pressures and pressure gradients
resulting from the collapse of gas bubbles, which enter the bacterial solution on
or near the bacterial cell wall. In combination, these events lead to bacterial cell
damage through mechanical fatigue. (ii) Shear forces are induced within bacterial
cells by microstreaming of their aqueous interiors, which is the unidirectional
movement of fluids near the inner and outer membranes of bacteria that can alter
their structure, function, and permeability. (iii) The high temperatures produced
by collapse cavitation induce the pyrolysis or sonochemical degradation of water
in the aqueous interior of bacterial cells. Free radicals such as H• and OH•
are produced by this process, which then attack the chemical structure of the
host organism’s cell wall and weaken it to the point of disintegration. (iv) Other
products of this sonochemical degradation of water such as H2 O2 are known to
be strong antibacterial agents and further contribute to the antimicrobial effect of
ultrasound. (v) There is also evidence to suggest that H• radicals produced by the
action of ultrasound are able to cause damage to bacterial DNA via the induction
of breaks in the double strands of these molecules [60, 86, 100]. These acoustic
cavitation events are also believed to contribute to the ability of ultrasound to
inhibit and disrupt the formation of bacterial biofilms [101–103] by limiting
both the motility of bacterial cells [104] and their ability to attach to surfaces
[103, 105], which is vital to the construction of these sessile microbial communities [106, 107]. In addition, ultrasound can induce the detachment of bacterial
cells already adhering to surfaces and kill cells within these communities via
the lysis of their membranes [49]. The extracellular polymeric matrix and water
channels are essential to the architectural stability and growth of these microbial
communities [106, 107] and further enhancing the action of ultrasound against
bacterial biofilms, this radiation is also able to induce the mechanical destruction
of these structures [49].
11.3
The Antimicrobial Activity of Assisted Ultrasound
It has been established that for a variety of reasons, some microorganisms are relatively resistant to the action of high-power ultrasound alone as compared to other
microbes, even between those that are closely related [108]. For example, it has
been reported that Gram-negative, rod-shaped bacteria, such as P. aeruginosa, are
more susceptible to the action of the radiation than Gram-positive, coccus-shaped
bacteria, such as S. aureus [59, 94, 109]. It has been suggested that this difference in
susceptibility is due, at least in part, to the differing characteristics of the bacterial
cell wall possessed by these two bacterial classes [59, 93]. Essentially, most Grampositive bacteria contain a denser, stronger, and thicker layer of peptidoglycan in
their cell wall, as compared to that of Gram-negative bacteria [110], which would
appear to render the former organisms more resistant to the action of ultrasound
than the latter organisms [59, 93]. It has also been found that spore-forming bacteria such as Bacillus spp. are more resistant to the action of ultrasound alone than
335
336
11
The Antimicrobial Effects of Ultrasound
vegetative bacteria [85, 111]. In general, microorganisms with smaller cells, and
therefore lower surface areas, are more resistant to the action of ultrasound alone,
primarily because of their decreased vulnerability to the pressures produced by
acoustic cavitation events as compared to larger microbial cells [112]. In response
to these observations, there have been a number of attempts to couple the antimicrobial capability of other strategies to that of ultrasound, which have shown that
in many cases, the killing of microbes is enhanced.
11.3.1
Synergistic Effects
Numerous studies have shown that a synergistic antimicrobial effect can be
achieved by the co-application of high-power ultrasound with other forms
of energy [36, 93, 113, 114], primarily visible light [72], ultraviolet light [61,
115–117], pulsed electric fields [118, 119], hydrodynamic pressure (osmosonication) [120, 121], atmospheric pressure (manosonication) [122–125], heat
(thermosonication) [125–128], and heat/pressure (manothermosonication) [125,
129]. These applications of high-power ultrasound are generally known as sonodynamic disinfection, and in most cases, the precise mechanisms by which the
radiation and these other forms of energy synergize to produce their antimicrobial
effects are unclear and vary [36]. However, acoustic cavitation events combined
with the effects of these various energy forms, variously induce turbulence,
shear stress, heating effects, electroporation, photon adsorption, and free radical
production, resulting in damage to microbial membranes, proteins, and DNA,
which appear to be significant factors in these antimicrobial mechanisms [36,
61, 72, 93, 113, 114, 120, 130]. For example, in the case of manosonication, it has
been established that the synergy observed between ultrasound and pressure
can be assigned to a number of different causes, including an increase in the
rates of collapse cavitation and free radical production. However, it has also
been established that these effects occur only up to a pressure maximum and
it is therefore important to determine this critical pressure level to achieve
the maximum synergetic antimicrobial effect of manosonication. Above this
pressure, there is a decrease in antimicrobial efficacy, which is associated with a
decrease of collapse cavitation phenomena, essentially because ultrasound waves
are unable to overcome the combined cohesive forces of over-pressure and the
cohesive force of molecules in the liquid medium [36, 125].
Sonodynamic disinfection also includes the antimicrobial effect of high-power
ultrasound when synergized with a variety of chemicals and compounds [36,
53, 86, 131], including ozone (sonozonation) [132–134], plasma [135–137],
chlorine dioxide [138–140], organic acids [139, 141, 142], chelating agents [143,
144], steam [130], naturally occurring antimicrobials, both endogenous [145]
and exogenous [146–149], and enzymatic solutions [144, 150, 151]. Again,
the detailed mechanisms by which ultrasound synergizes the action of these
chemicals and other compounds to produce antimicrobial effects are unclear
and vary. However, in many cases, it appears that acoustic cavitation events not
11.3
The Antimicrobial Activity of Assisted Ultrasound
only kill microbes directly, as described above, but also induce the detachment
of microbial cells from surfaces along with the mechanical disruption of these
cells and their membranes, a process generally known as sonoporation. This
action promotes the ability of chemicals and other compounds with antimicrobial activity to diffuse into microbial cells and biofilms, thereby enhancing
their decontamination and disinfection efficacy [36, 53, 86, 130, 131, 133–136,
138–151]. For example, in the case of sonozonation, it has been suggested that
the transfer of ozone into aqueous solution is enhanced by acoustic cavitation
and increases the rate by which ozone penetrates into microorganisms. The
formation of more active species, such as OH• radicals and nascent oxygen by
the decomposition of ozone is accelerated by acoustic cavitation processes in
aqueous solution. The oxidants thus formed can attack microbial cells, inducing
the rupture of membranes and damage to intracellular structures, such as DNA,
all of which contribute to a loss of cell viability [132, 133].
On the therapeutic front, the most researched antimicrobial effect of ultrasound is the co-application of the radiation with conventional antibiotics [49,
93], and particularly in relation to indwelling medical devices [43, 90]. This
research is generally taken to have begun in 1994 with a seminal study [104],
which is considered to mark the second period of research into the antimicrobial
capability of ultrasound [49]. This study showed that low-power ultrasound was
ineffective when applied alone against planktonic P. aeruginosa and E. coli [104].
However, when this low-power radiation was applied with gentamycin, an aminoglycoside antibiotic, this combination killed these organisms more efficiently
than gentamycin alone [104]. Four years later, similar observations were reported
for the action of a series of antibiotics, including other aminoglycosides along
with tetracyclines and penicillins against both Gram-negative bacteria, such
as Serratia marcescens, and Gram-positive organisms, such as Staphylococcus
epidermidis [152]. This synergistic antimicrobial effect has since been shown for a
variety of other antibiotics, including glycopeptides, lincomycins, macrolides, fluoroquinolones, sulfonamides, cephalosporins [59, 153–157], and most recently,
host defense peptides [158], which are endogenous antibiotics produced across
the eukaryotic kingdom [159]. In addition to planktonic bacteria, low-power
ultrasound and antibiotics have also been shown to exhibit synergistic effects
against sessile forms of these organisms [160–167]. For example, as compared to
vancomycin alone, this antibiotic in combination with ultrasound has been shown
to have increased efficacy against biofilms of E. coli, P. aeruginosa, S. aureus,
and S. epidermidis, respectively, all of which are known to cause problematic
infections on implanted medical devices [153, 166–168].
There have been a number of investigations into the mechanisms underpinning
the synergistic effects of ultrasound and antibiotics, which have suggested that
is primarily based on sonoporation [43, 101, 153, 166, 167]. In the case of planktonic bacteria, earlier studies suggested that low-power ultrasound induced stable
cavitation events that led to changes in bacterial membranes, thereby allowing
antibiotics to penetrate host cells more efficiently [104, 154]. A series of subsequent studies up to present times generally supported these findings and led to
337
338
11
The Antimicrobial Effects of Ultrasound
the view that ultrasound induced lesions in the membranes of bacterial cells. However, these studies also showed that these lesions were sufficiently large to allow the
internalization of antibiotics into bacterial cells to exert their antimicrobial action
but did not induce significant levels of membrane damage [152, 163, 169–172].
In addition, other studies have suggested that damage to the bacterial cell wall
and heat-mediated biochemical reactions within cells of these organisms can be
induced by stable cavitation events and contribute to the synergistic antimicrobial effect of ultrasound and antibiotics [157, 173]. Stable cavitation events also
appear to mediate the synergistic effects of ultrasound and antibiotics against
bacterial biofilms [161, 163, 174], primarily by creating numerous pores in their
extracellular matrix and thereby enhancing the ability of these molecules to diffuse
through the biofilm and localize near resident bacteria [166, 175, 176]. However,
it appear that this action can be enhanced by the ability of low-power ultrasound
to stimulate microbial growth [79] – an effect that was first reported in 2003 and
considered by some to mark the start of third historical period of research into
the antimicrobial capability of the radiation [49]. Essentially, it appears that pores
generated by low-power ultrasound also facilitate the transport of oxygen and
nutrients into biofilms, with the result that its upper layers are killed and subsequently removed by the combined action of the radiation and antibiotics. As
a consequence of this action, increased levels of nutrients are made available to
underlying bacteria in the biofilm, which are normally dormant but under these
conditions can become active in the presence of high concentrations of antibiotics.
Taken in combination, these synergistic events significantly enhance the ability
of ultrasound to eradicate sessile organisms as compared to the action of either
antibiotics or the radiation when applied alone [79, 166, 167, 175, 177, 178].
11.3.2
Sonosensitizers
In what must be considered to herald the dawn of the fourth period of research
into the antimicrobial potential of ultrasound, it was demonstrated in 2013 that
the radiation possessed the ability to activate photosensitizers (PS) [42], confirming predictions made in 2009 [179]. As described in Chapter 10, PS are generally
activated by light to facilitate the underlying mechanisms of cytotoxicity involved
in photodynamic antimicrobial therapy (PACT) [180, 181], but it was found that
ultrasound was able to induce antibacterial activity by Rose Bengal (RB) [42],
which is an established PACT PS [182–184]. These investigations showed that
in the presence of RB, the application of low-power ultrasound led to the highly
efficient eradication of both Gram-positive bacteria, including S. aureus, and
Gram-negative organisms, including E. coli [42]. An ability to kill E. coli was also
demonstrated when low-power ultrasound was found to enhance the antibacterial
efficacy of ciprofloxacin and levofloxacin [156], which are well-established antibiotics [185, 186] and have been shown to act as PS in photodynamic therapy, which
is based on the anticancer capability of these molecules [187, 188]. These studies
appear to be the first major experimental demonstration of what has been termed
11.3
The Antimicrobial Activity of Assisted Ultrasound
sonodynamic antimicrobial chemotherapy (SACT) and it has been suggested that
this modality could of medical importance with uses ranging from the sterilization of surgical instruments to the treatment of microbial infections [42, 179]. The
mechanisms by which ultrasound-activated PS kill microbes are currently poorly
understood [42, 156] although insight into these mechanisms can be gained from
studies on the anticancer activities of these molecules where they are generally
known as sonosensitizers (SSs) and the corresponding therapeutic modality as
sonodynamic therapy (SDT) [189, 190]. On the basis of SDT studies, which also
utilize low-power ultrasound, it is generally accepted that the activation of SS
involves collapse cavitation within or near target cancer cells. However, it has
been variously suggested that the mechanisms underpinning this activation of
SS can include sonoporation, sonochemistry, and sonoluminescence [74, 75, 82,
83, 189–191]. Several studies have suggested that the anticancer activity of ultrasound and SS is essentially based on sonoporation and involves the mechanical
ability of these molecules to permeate and destabilize the integrity of target cell
membranes. The effect on these weakened membranes is able to synergize with
stress induced in host cells by the shearing effect of ultrasound-mediated bubble
pulsations with resulting cancer cell death [192, 193]. However, many studies have
proposed that the damage inflicted in SDT involves the activation of SS to induce
the production of reactive oxygen species (ROS) with cytotoxicity to cellular components such as membrane lipid [194], DNA [195, 196], and proteins [197–200]
along with cancer cells [201–207] and other cell types [208–210]. Indeed, a recent
study predicted that ROS production would be optimal in ultrasound-mediated
bubble collapse temperatures and pressures of circa 5000 K and 250 MPa, which
are the conditions most usually associated with collapse cavitation events [84].
One earlier study proposed that the anticancer activity of SDT resulted from the
activation of SS via sonochemical reactions occurring either inside collapsing
cavitation bubbles or in the heated gas–liquid interface. According to this study,
the bubble interior may be considered to represent a micro-reactor in which SSderived free radicals are produced either by direct pyrolysis or via reactions with
other radicals formed by the pyrolysis of water. These SS-derived free radicals,
which are predominantly carbon-centered, then react with dissolved oxygen to
form peroxyl and alkoxyl radicals, which are able to initiate damage to critical cellular sites (Figure 11.1) [211]. However, the most commonly proposed mechanism
for the anticancer activity of SDT is the production of ROS through the direct electronic excitation of SS by sonoluminescence (Figure 11.2) [189, 190, 203], which
are flashes of light that are produced during the course of bubble collapse in acoustic cavitation [74, 75]. Sonoluminescence can also be induced by some animals
[212–214] and has been observed in some synergistic antimicrobial activities of
ultrasound [36] but currently, the mechanism(s) underpinning the phenomenon
are unknown with proposed explanations ranging from surface black body
radiation to quantum vacuum radiation [215, 216]. Regardless of the mechanism
underpinning sonoluminescence, it is generally accepted that the activation of
339
340
11
The Antimicrobial Effects of Ultrasound
O2
Sonosensitizers
H2 O
·H
·OH
OO·
Removal by
scavengers
Microbe
H2O
Collapsing cavitation bubble
Figure 11.1 This figure represents the activation of sonosensitizers by sonochemistry
in sonodynamic antimicrobial therapy. In this
representation, a sonosensitizer undergoes
pyrolysis inside collapsing cavitation bubbles
or in the heated gas–liquid interface, forming free radical intermediates. These intermediates, which are predominantly carboncentered, react with dissolved O2 to form
peroxyl radicals, capable of attacking critical
cellular sites in microbial cells because of
their ability to diffuse significant distances. In
contrast, free radicals, which are also formed
during cavitational collapse through the
sonochemistry of water, are unable to cause
significant cellular damage to microbes,
because of their extremely high reactivity
and hence short diffusion distances. (This
figure was adapted from Ref. [211].)
Excited state
3O
En
2
erg
Ground state of
sonosensitizer
y
Triplet
oxygen
Sonoluminescent
light
1
O2
Ultrasound
irradiation
Microbe
O2 Singlet
oxygen
1
1
O2
Microbubbles formed
by acoustic cavitation
Figure 11.2 This figure represents the activation of sonosensitizers by sonoluminescence in sonodynamic antimicrobial therapy.
In this representation, a microbe is irradiated
by low-power ultrasound, which induces collapse cavitation around the surface of the
cell and produces sonoluminescent light. A
sonosensitizer, which tends to attach to the
surface of microbial cells, is exposed to the
Sonosensitizer
sonoluminescent light and becomes activated from its ground state into an excited
state. As the activated sonosensitizer returns
to the ground state, the energy released can
generate reactive oxygen species, such as
singlet oxygen, which mediate the photooxidation of cellular components, thereby
inducing cell death. (This figure was adapted
from Ref. [190].)
11.4
Future Prospects
SS by this ultrasound-induced light source uses a photochemical pathway similar
to that involved in the photodynamic activation of these molecules [179, 190, 201,
202]. As can be seen from Figure 11.2, the exposure of SS to sonoluminescent light
promotes these molecules from their ground state to an electronically excited state
and as the activated SS returns to its ground state, the energy released generates
ROS, such as singlet oxygen and free radicals, which mediate the photo-oxidation
of cellular components, thereby inducing cell death [190]. Support for the use of
this mechanism in SACT would appear to be provided by a series of studies on
sonocatalytic disinfection by TiO2, which suggested that this process was synergized by the sonoluminescent activation of the compound [217–221] as recently
reported for other sonocatalysts [199]. TiO2 is known to serve as a photocatalytic
disinfectant with a wide range of antimicrobial activity [222] and it was found
that irradiation of the compound with ultrasound led to the ROS-mediated lysis
of microbes such as E. coli and Legionella pneumophila accompanied by oxidative damage to DNA and membrane lipids [217–220]. Studies have shown that
other sonocatalysts, including gold nanoparticles and Teraftal, a cobalt octacarboxyphthalocyanine, are activated by ultrasound to kill E. coli and other bacteria
such as S. aureus although in these cases, the sonodynamic antimicrobial mechanisms utilized have not yet been elucidated [223, 224]. Indirect support for use
of sonoluminescence in SACT was provided when it was observed that the activation of ciprofloxacin and levofloxacin by ultrasound led to the production of
ROS, which were proposed to induce the lysis of E. coli cells along with oxidative
damage to intracellular components of the organism [156]. The maximum emission intensity of sonoluminescence in water lies between 250 and 600 nm [225]
and this emission range was found to correlate with the absorbance spectra of
both ciprofloxacin (𝜆max at 276, 316, and 328 nm) [187] and levofloxacin (𝜆max at
288 and 331 nm) [188]. Similarly, the absorbance spectrum of RB (𝜆max ∼ 565 nm),
shown above to be a potent SACT agent, was found to exhibit good overlap with
the emission range of sonoluminescence [42]. However, no ultrasound-mediated
antimicrobial activity was detected in corresponding experiments on Methylene
Blue (MB), which is also an established PACT PS (Chapter 10) [226, 227], and
the molecule showed no significant absorbance (𝜆max ∼ 655 nm) in the range 250
and 600 nm [42]. On the basis of these observations, it was suggested by the latter authors that the limited emission range observed for sonoluminescence may
partly explain why MB and other molecules can function predominantly as either
PS or SS while yet others can serve in both capacities [41, 190].
11.4
Future Prospects
In this chapter, it has been shown that ultrasound has an antimicrobial capacity in
its own right, which is effective against a spectrum of microorganisms and finds
application as a disinfection and decontamination agent in many areas, of which
the food industry is a major example. Indeed, the radiation has been considered
341
342
11
The Antimicrobial Effects of Ultrasound
a technology with high potential in this industry not only for its antimicrobial
efficacy but also for reasons such as its eco-friendly properties, its ability to inactivate enzymes responsible for food deterioration, and that its use results in energy
and cost savings [87]. Nonetheless, the use of ultrasound alone for antimicrobial
disinfection and decontamination has a number of drawbacks and in particular,
as shown in this chapter, a variety of microorganisms are relatively resistant to
its action. Owing to this resistance, application of the radiation alone is often
unable to decrease microbial loads sufficiently to meet existing legislation limits
as in the case of the food industry [36]. In response, to increase the disinfection and decontamination efficacy of ultrasound, it has been used in conjunction
with other antimicrobial strategies, including various energy forms, chemicals,
and antimicrobials to elicit a synergistic inactivation of microorganisms. This use
of assisted ultrasound is increasingly embracing emerging technologies and, as an
example, the combined use of this radiation and plasma technology has recently
been shown for the first time to exhibit a synergistic ability to inactivate bacteria and yeasts [136], which is a technology in the pipeline for use in the food
industry [137]. The synergistic antimicrobial use of ultrasound has also been utilized in the therapeutic arena where the radiation has been used to enhance the
action of antibiotics, particularly when directed against bacterial biofilms. Given
the current dearth of novel conventional antibiotics [228], this use of assisted
ultrasound has sought the use of new classes of antimicrobials, exemplified by
the very recent demonstration that the radiation can synergize the antibacterial
efficacy of host defense peptides [158], which are predicted to form the next generation of antibiotics [229]. In this study, it was shown that the application of ultrasound with defensins, which are endogenous human defense peptides, were active
against biofilms formed by Staphylocci, which are a serious concern within the
biomaterial community [158]. Moreover, it was also shown by this latter study that
this synergistic antimicrobial strategy was effective against biofilms of methicillinresistant S. aureus [158], which gives it a major advantage over many conventional
antibiotics, which are ineffective against multidrug-resistant bacteria [228]. This
chapter has also shown what would appear to be a novel synergistic antimicrobial capability of ultrasound in its ability to induce the antibacterial activity of
RB, which usually requires light activation to achieve this effect [42]. The precise
mechanisms underlying this ultrasound-mediated ability are unclear but the most
favored explanation appears to be the activation of the molecule by sonoluminescent light that has been generated by the radiation. However, other mechanisms
based on sonoporation and sonochemistry may contribute to this effect, which
could also indicate the direct involvement of ultrasound action in this effect and
thereby the occurrence of a synergistic antimicrobial interaction between the radiation and activated RB. Regardless of the mechanisms underpinning the antibacterial activity of RB and ultrasound, the demonstration of this activity would appear
to open an entirely new field of antimicrobial research mediated by the radiation as was recently envisaged [179]. Indeed, the potential of this field of research
could be vast; there are numerous light-activated molecules that exhibit photoantimicrobial activity and could potentially be activated by ultrasound [180, 230].
References
Moreover, although the use of light-activated molecules to treat microbial infections is gaining increasing acceptance, it is not a mainstream option because of
the limited penetration of tissue by visible light and it is well established that significantly deeper tissue penetration can be achieved by ultrasound, which would
clearly favor the use of SACT over PACT [179].
In summary, this chapter clearly shows that the antimicrobial use of ultrasound
is a rapidly expanding and successful field, although in many cases, detailed
descriptions of the mechanisms and processes involved in this use are largely
unclear. Indeed, this chapter has also shown that at some power levels, ultrasound
can stimulate the growth of bacteria, which could clearly be undesirable in some
cases and emphasizes a general principle, which is the need for a holistic approach
to the design of ultrasound-mediated strategies. For example, the design of such
a strategy should include consideration of the nature of the target microbes, the
parameters of the incident radiation, the effects of the microbial environment,
and the microbial inactivation kinetics involved. However, with these lessons
taken on board, the antimicrobial use of ultrasound would appear to have a bright
and promising future – one could say quite ultrasonic in fact!!
References
1. Tzabo, T.L. (2014) Diagnostic Ultra-
2.
3.
4.
5.
6.
7.
8.
9.
sound Imaging: Inside Out, Academic
Press, Oxford.
Hunt, F.V. (1978) Origins in Acoustics:
The Science of Sound from Antiquity to
the Age of Newton, Books on Demand.
Beyer, R.T. (1999) Sounds of Our
Times: Two Hundred Years of Acoustics,
Springer-Verlag, New York.
Rossing, T. (2007) in Springer Handbook of Acoustics (ed. T.D. Rossing),
Springer-Verlag, New York, pp. 9–24.
Griffin, D.R. (2001) Return to the magic
well: echolocation behavior of bats and
responses of insect prey. Bioscience, 51,
555–556.
Eisenberg, R. (1995) Radiology: An
Illustrated History, Mosby, St. Louis,
MO.
Griffin, D.R. and Galambos, R. (1941)
The sensory basis of obstacle avoidance by flying bats. J. Exp. Zool., 86,
481–506.
Colladon, J.-D. and Sturm, J.C.F. (1827)
Mémoire sur la Compression des
liquides. Ann. Chim. Phys., 3, 27–38.
Malumbres, M.P., Garrido, P.P.,
Calafate, C.T., and Gil, J.O. (2009)
Encyclopedia of Information Science
10.
11.
12.
13.
14.
15.
and Technology, 2nd edn, IGI Global,
pp. 3859–3864.
Curie, P. and Curie, J. (1881) Contractions et dilatations produites
par des tensions électriques dans les
cristaux hémiédres á faces inclinées.
C.R. (France), 93, 1137–1140.
Curie, P. and Curie, J. (1880)
Développement, par pression, de
l’électricité polaire dans les cristaux
hémiédres á faces inclinées. C.R.
(France), 91, 294–295.
Amir, M. and Cobbold Richard, S.C.
(2011) Development and application
of piezoelectric materials for generation and detection. Ultrasound, 19,
187–196.
Katzir, S. (2012) Who knew piezoelectricity? Rutherford and langevin on
submarine detection and the invention of sonar. Notes Rec. R. Soc., 66,
141–157.
Langevin, M.P. and Ishimoto, M. (1923)
Utilization des phenomen‘es piezoelectrique pour la measure de l’intensit é
des sons en valeur absolute. J. Physiol.,
4, 539–540.
Wood, R.W. and Loomis, A.L. (1927)
The physical and biological effects of
343
344
11
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
The Antimicrobial Effects of Ultrasound
high frequency sound waves of great
intensity. Philos. Mag. Ser. VII, 4,
417–436.
Graff, K.F. (1982) A History of Ultrasonics in Physical Acoustics, vol. XV,
Academic Press, New York.
Zeiss, E. (1938) Ueber Linsenveraenderungen an herausgenommenen
Rinderlinsen durch Ultraschalleinwirkung. Graefes Arch Ophthalmol.,
139, 301–322.
Thijssen, J.M. (1993) The History of
Ultrasound Techniques in Ophthalmology. Ultrasound Med. Biol., 19,
599–618.
Dussik, K.T. (1942) Über die
Möglichkeit, hochfrequente mechanische Schwingungen als diagnostisches
Hilfsmittel zu verwerten. Z. Neurol.
Psychiat., 174, 153–168.
Thomas, A.M.K., Banergee, A.K., and
Busch, U. (2005) Classic Papers in
Modern Diagnostic Radiology, SpringerVerlag, Berlin.
Lynn, J.G., Zwemer, R.L., Chick, A.J.,
and Miller, A.E. (1942) A new method
for the generation and use of focused
ultrasound in experimental biology. J.
Gen. Physiol., 26, 179–193.
Luigi, M. (2013) High intensity focused
ultrasound, liver disease and bridging
therapy. World J. Gastroenterol., 19,
7494–7499.
Lynn, J.G. and Putnam, T.J. (1944) Histology of cerebral lesions produced by
focused ultrasound. Am. J. Pathol., 20,
637–649.
Fry, W.J., Mosberg, W.H. Jr., Barnard,
J.W., and Fry, F.J. (1954) Production of
focal destructive lesions in the central
nervous system with ultrasound. J.
Neurosurg., 11, 471–478.
Wall, P.D., Fry, W.J., Stephens, R.,
Tucker, D., and Lettvin, J.Y. (1951)
Changes produced in the central nervous system by ultrasound. Science
(New York), 114, 686–687.
Fry, W.J., Barnard, J.W., Fry, E.J.,
Krumins, R.F., and Brennan, J.F. (1955)
Ultrasonic lesions in the mammalian
central nervous system. Science (New
York), 122, 517–518.
Fry, F.J. (1958) Precision high intensity focusing ultrasonic machines
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
for surgery. Am. J. Phys. Med., 37,
152–156.
Shampo, M.A. and Kyle, R.A. (1997)
John Julian Wild—pioneer in ultrasonography. Mayo Clin. Proc., 72,
234.
Burov, A.K. and Andreevskaya, G.D.
(1956) Effect of high intensity supersonic oscillations on malignant tumours
in animals and in man. Dokl. Akad.
Nauk SSSR, 106, 445–458.
Burov, V.A., Dmitrieva, N.P., and
Rudenko, O.V. (2002) Nonlinear ultrasound: breakdown of microscopic
biological structures and nonthermal
impact on a malignant tumor. Dokl.
Akad. Nauk, 383, 101–104.
Donald, I., Macvicar, J., and Brown,
T.G. (1958) Investigation of abdominal
masses by pulsed ultrasound. Lancet, 1,
1188–1195.
Nicolson, M. and Fleming, J.E.E. (2013)
Imaging and Imagining the Fetus:
The Development of Obstetric Ultrasound, Johns Hopkins University Press,
Baltimore, MD.
Griffiths, A., terHaar, G., Rivens, I.,
Giussani, D., and Lees, C. (2012)
High-intensity focused ultrasound
in obstetrics and gynecology: the birth
of a new era of noninvasive surgery?
Ultraschall Med. (Stuttgart: 1980), 33,
E8–E15.
Itro, A., Lupo, G., Carotenuto, A.,
Filipi, M., Cocozza, E., and Marra, A.
(2012) Benefits of piezoelectric surgery
in oral and maxillofacial surgery.
Review of literature. Minerva Stomatol.,
61, 213–224.
Zhao, Y.-Z., Du, L.-N., Lu, C.-T., Jin,
Y.-G., and Ge, S.-P. (2013) Potential and
problems in ultrasound-responsive drug
delivery systems. Int. J. Nanomed., 8,
1621–1633.
Sango, D.M., Abela, D., McElhatton, A.,
and Valdramidis, V.P. (2014) Assisted
ultrasound applications for the production of safe foods. J. Appl. Microbiol.,
116, 1067–1083.
Sarvazyan, A.P., Rudenko, O.V., and
Nyborg, W.L. (2010) Biomedical applications of radiation force of ultrasound:
historical roots and physical basis.
Ultrasound Med. Biol., 36, 1379–1394.
References
38. Watson, T. (2008) Ultrasound in con-
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.
temporary physiotherapy practice.
Ultrasonics, 48, 321–329.
Bubulis, A., Adzerikho, I., Stepanenko,
D., Minchenya, V., Valaika, M., Veikutis,
V., Pranskunas, A., Unikas, R., and
Morkunaite, K. (2012) Low frequency
and high intensity ultrasound in vascular surgery: theory, instrumentation
and possibilities of clinical application.
J. Vibroeng., 14, 1833–1851.
Park, D., Park, H., Seo, J., and Lee, S.
(2014) Sonophoresis in transdermal
drug deliverys. Ultrasonics, 54, 56–65.
Shibaguchi, H., Tsuru, H., Kuroki, M.,
and Kuroki, M. (2011) Sonodynamic
cancer therapy: a non-invasive and
repeatable approach using low-intensity
ultrasound with a sonosensitizer. Anticancer Res., 31, 2425–2429.
Nakonechny, F., Nisnevitch, M., Nitzan,
Y., and Nisnevitch, M. (2013) Sonodynamic excitation of rose bengal
for eradication of Gram-positive and
Gram-negative bacteria. BioMed Res.
Int., 2013, 7.
Yu, H., Chen, S., and Cao, P. (2012)
Synergistic bactericidal effects and
mechanisms of low intensity ultrasound and antibiotics against bacteria:
a review. Ultrason. Sonochem., 19,
377–382.
Shehata, I.A. (2012) Treatment with
high intensity focused ultrasound:
secrets revealed. Eur. J. Radiol., 81,
534–541.
Geis, N.A., Katus, H.A., and
Bekeredjian, R. (2012) Microbubbles
as a vehicle for gene and drug delivery:
current clinical implications and future
perspectives. Curr. Pharm. Des., 18,
2166–2183.
Azhari, H. (2012) Ultrasound: medical imaging and beyond (An Invited
Review). Curr. Pharm. Biotechnol., 13,
2104–2116.
Mason, T.J. (2011) Therapeutic
ultrasound an overview. Ultrason.
Sonochem., 18, 847–852.
Harvey, E.N. (1930) Biological aspects
of ultrasonic waves: a general survey.
Biol. Bull., 59, 306–325.
Erriu, M., Blus, C.,
Szmukler-Moncler, S., Buogo, S.,
50.
51.
52.
53.
54.
55.
56.
57.
58.
Levi, R., Barbato, G., Madonnaripa,
D., Denotti, G., Piras, V., and Orrù, G.
(2014) Microbial biofilm modulation
by ultrasound: current concepts and
controversies. Ultrason. Sonochem., 21,
15–22.
Harvey, E.N. and Loomis, A.L. (1929)
The destruction of luminous bacteria by high frequency sound waves. J.
Bacteriol., 17, 373–376.
Harvey, E.N. and Loomis, E.L. (1928)
High frequency sound waves of small
intensity and their biological effects.
Nature, 121, 622–624.
Center for Aquatic Plant Management (2003) Annual Summary Report,
Reading.
Alzamora, S., Guerrero, S., Schenk, M.,
Raffellini, S., and López-Malo, A.
(2011) in Ultrasound Technologies for
Food and Bioprocessing (eds H. Feng,
G. Barbosa-Canovas, and J. Weiss),
Springer, New York, pp. 321–343.
Broekman, S., Pohlmann, O.,
Beardwood, E.S., and de Meulenaer,
E.C. (2010) Ultrasonic treatment
for microbiological control of water
systems. Ultrason. Sonochem., 17,
1041–1048.
Rapuntean, G., Fit, N., Nadas, G.,
Denes, A., and Cuc, C. (2005) Ultrasounds action on microorganisms in
normal culture conditions. Buletinul
USAMV-CN., 62, 281–285.
Scherba, G., Weigel, R.M., and O’Brien,
W.D. Jr., (1991) Quantitative assessment
of the germicidal efficacy of ultrasonic
energy. Appl. Environ. Microbiol., 57,
2079–2084.
Declerck, P., Vanysacker, L., Hulsmans,
A., Lambert, N., Liers, S., and Ollevier,
F. (2010) Evaluation of power ultrasound for disinfection of both
Legionella pneumophila and its
environmental host Acanthamoeba
castellanii. Water Res., 44, 703–710.
Drakopoulou, S., Terzakis, S.,
Fountoulakis, M.S., Mantzavinos, D.,
and Manios, T. (2009) Ultrasoundinduced inactivation of Gram-negative
and Gram-positive bacteria in secondary treated municipal wastewater.
Ultrason. Sonochem., 16, 629–634.
345
346
11
The Antimicrobial Effects of Ultrasound
59. Maleki, A., Shahmoradi, B., Daraei, H.,
60.
61.
62.
63.
64.
65.
66.
and Kalantar, E. (2013) Assessment of
ultrasound irradiation on inactivation
of Gram negative and positive bacteria isolated from hospital in aqueous
solution. J. Adv. Environ. Health Res., 1,
9–14.
Joyce, E., Phull, S.S., Lorimer, J.P., and
Mason, T.J. (2003) The development
and evaluation of ultrasound for the
treatment of bacterial suspensions. A
study of frequency, power and sonication time on cultured Bacillus species.
Ultrason. Sonochem., 10, 315–318.
van der Walt, E. (2002) The effect of
ultraviolet light, cavitational flow and
ultrasound on protozoan cysts and
oocysts, bacteriophages and Clostridium. WISA, 2002, Durban, 19, May
2002.
Ahn, C.Y., Park, M.H., Joung, S.H., Kim,
H.S., Jang, K.Y., and Oh, H.M. (2003)
Growth inhibition of cyanobacteria by
ultrasonic radiation: laboratory and
enclosure studies. Environ. Sci. Technol.,
37, 3031–3037.
Nakanishi, M., Mukai, S., Kimata, I.,
Iseki, M., and Maeda, Y. (2001) Inactivation of cryptosporidium parvum
oocysts in drinking water by highintensity ultrasonic waves. Paper
presented at the Annual Conference
of the American Water Works Association, Washington, DC.
Ashokkumar, M., Vu, T., Grieser, F.,
Weerawardena, A., Anderson, N.,
Pilkington, N., and Dixon, D.R. (2003)
Ultrasonic treatment of Cryptosporidium oocysts. Water Sci. Technol., 47,
173–177.
Olvera, M., Eguia, A., Rodriguez, O.,
Chong, E., Pillai, S.D., and Ilangovan,
K. (2008) Inactivation of Cryptosporidium parvum oocysts in water using
ultrasonic treatment. Bioresour. Technol., 99, 2046–2049.
Martin, J.F.G., Guillemet, L., Feng,
C., and Sun, D.-W. (2013) Cell viability and proteins release during
ultrasound-assisted yeast lysis of light
lees in model wine. Food Chem., 141,
934–939.
67. Dehghani, M.H., Mahvi, A.H., Jahed,
68.
69.
70.
71.
72.
73.
74.
75.
76.
77.
G.R., and Sheikhi, R. (2007) Investigation and evaluation of ultrasound
reactor for reduction of fungi from
sewage. J. Zhejiang Univ. Sci. B, 8,
493–497.
Liu, D., Zeng, X.-A., Sun, D.-W., and
Han, Z. (2013) Disruption and protein
release by ultrasonication of yeast cells.
Innovative Food Sci. Emerg. Technol.,
18, 132–137.
Kapturowska, A., Stolarzewicz,
I., Chmielewska, I., and
Bialecka-Florjanczyk, E. (2011) Ultrasounds – a tool to inactivate yeast
and to extract intracellular protein.
Zywn-Nauk Technol. JA, 18, 160–171.
Iida, Y., Tuziuti, T., Yasui, K., Kozuka,
T., and Towata, A. (2008) Protein
release from yeast cells as an evaluation
method of physical effects in ultrasonic field. Ultrason. Sonochem., 15,
995–1000.
Su, X., Zivanovic, S., and D’Souza, D.H.
(2010) Inactivation of human enteric
virus surrogates by high-intensity
ultrasound. Foodborne Pathog. Dis., 7,
1055–1061.
Chrysikopoulos, C.V., Manariotis, I.D.,
and Syngouna, V.I. (2013) Virus inactivation by high frequency ultrasound in
combination with visible light. Colloids
Surf., B, 107, 174–179.
Zuber, S., Butot, S., and Baert, L.
(2013) in Foodborne Viruses and Prions and Their Significance for Public
Health (eds F. Smulders, B. Nørrung,
and H. Budka), Wageningen Academic,
Wageningen, pp. 113–136.
Leong, T., Ashokkumar, M., and
Kentish, S. (2011) The fundamentals of power ultrasound – a review.
Acoust. Aust., 39, 54–63.
Rooze, J., Rebrov, E.V., Schouten, J.C.,
and Keurentjes, J.T.F. (2013) Dissolved
gas and ultrasonic cavitation – a
review. Ultrason. Sonochem., 20, 1–11.
Baker, K.G., Robertson, V.J., and Duck,
F.A. (2001) A review of therapeutic
ultrasound: biophysical effects. Phys.
Ther., 81, 1351–1358.
Guzman, H.R., McNamara, A.J.,
Nguyen, D.X., and Prausnitz, M.R.
(2003) Bioeffects caused by changes in
References
78.
79.
80.
81.
82.
83.
84.
85.
86.
87.
acoustic cavitation bubble density and
cell concentration: a unified explanation based on cell-to-bubble ratio and
blast radius. Ultrasound Med. Biol., 29,
1211–1222.
ter Haar, G. (2007) Therapeutic applications of ultrasound. Prog. Biophys. Mol.
Biol., 93, 111–129.
Pitt, W.G. and Ross, S.A. (2003) Ultrasound increases the rate of bacterial
cell growth. Biotechnol. Prog., 19,
1038–1044.
Wu, J. and Nyborg, W.L. (2008) Ultrasound, cavitation bubbles and their
interaction with cells. Adv. Drug Delivery Rev., 60, 1103–1116.
Kodama, T., Tomita, Y., Koshiyama,
K.-I., and Blomley, M.J.K. (2006) Transfection effect of microbubbles on
cells in superposed ultrasound waves
and behavior of cavitation bubble.
Ultrasound Med. Biol., 32, 905–914.
Suslick, K.S., Eddingsaas, N.C.,
Flannigan, D.J., Hopkins, S.D., and
Xu, H. (2011) Extreme conditions
during multibubble cavitation: sonoluminescence as a spectroscopic probe.
Ultrason. Sonochem., 18, 842–846.
Suslick, K.S. and Flannigan, D.J. (2008)
Inside a collapsing bubble: sonoluminescence and the conditions during
cavitation. Annu. Rev. Phys. Chem., 59,
659–683.
Merouani, S., Hamdaoui, O., Rezgui, Y.,
and Guemini, M. (2014) Theoretical
estimation of the temperature and
pressure within collapsing acoustical bubbles. Ultrason. Sonochem., 21,
53–59.
Butz, P. and Tauscher, B. (2002) Emerging technologies: chemical aspects.
Food Res. Int., 35, 279–284.
Bilek, S.E. and Turantas, F. (2013)
Decontamination efficiency of high
power ultrasound in the fruit and vegetable industry, a review. Int. J. Food
Microbiol., 166, 155–162.
Chemat, F., Huma, Z., and Khan,
M.K. (2011) Applications of ultrasound in food technology: processing,
preservation and extraction. Ultrason.
Sonochem., 18, 813–835.
88. Fellows, P. (2000) Food Processing Tech-
89.
90.
91.
92.
93.
94.
95.
96.
97.
nology: Principles and Practice, CRC
Press, New York.
Hutchinson, G. (2008) Sound water
practices: ultrasonic technology controls algae and biofilm. Opflow, 34,
18–19.
Dror, N., Mandel, M., Hazan, Z., and
Lavie, G. (2009) Advances in microbial
biofilm prevention on indwelling medical devices with emphasis on usage of
acoustic energy. Sensors, 9, 2538–2554.
Awad, T.S., Moharram, H.A., Shaltout,
O.E., Asker, D., and Youssef, M.M.
(2012) Applications of ultrasound in
analysis, processing and quality control
of food: a review. Food Res. Int., 48,
410–427.
Gao, S., Lewis, G.D., Ashokkumar,
M., and Hemar, Y. (2014) Inactivation
of microorganisms by low-frequency
high-power ultrasound: 1. Effect of
growth phase and capsule properties of
the bacteria. Ultrason. Sonochem., 21,
446–453.
Piyasena, P., Mohareb, E., and
McKellar, R.C. (2003) Inactivation
of microbes using ultrasound: a review.
Int. J. Food Microbiol., 87, 207–216.
Monsen, T., Lovgren, E., Widerstrom,
M., and Wallinder, L. (2009) In vitro
effect of ultrasound on bacteria and
suggested protocol for sonication and
diagnosis of prosthetic infections. J.
Clin. Microbiol., 47, 2496–2501.
Venieri, D., Markogiannaki, E.,
Chatzisymeon, E., Diamadopoulos, E.,
and Mantzavinos, D. (2013) Inactivation
of Bacillus anthracis in water by photocatalytic, photolytic and sonochemical
treatment. Photochem. Photobiol. Sci.,
12, 645–652.
Conner-Kerr, T., Alston, G., Stovall,
A., Vernon, T., Winter, D., Meixner,
J., Grant, K., and Kute, T. (2010)
The effects of low-frequency ultrasound (35 kHz) on methicillin-resistant
staphylococcus aureus (MRSA) in vitro.
Ostomy Wound Manage., 56, 32–43.
Mason, T.J. (2007) Sonochemistry and
the environment – providing a “green”
link between chemistry, physics and
engineering. Ultrason. Sonochem., 14,
476–483.
347
348
11
The Antimicrobial Effects of Ultrasound
98. Lambert, N., Rediers, H., Hulsmans,
99.
100.
101.
102.
103.
104.
105.
106.
107.
A., Joris, K., Declerck, P., De Laedt, Y.,
and Liers, S. (2010) Evaluation of ultrasound technology for the disinfection
of process water and the prevention
of biofilm formation in a pilot plant.
Water Sci. Technol., 61, 1089–1096.
Mukhopadhyay, S. and Ramaswamy, R.
(2012) Application of emerging technologies to control Salmonella in foods:
a review. Food Res. Int., 45, 666–677.
Winward, G.P. (2007) Disinfection of
Grey Water, Cranfield.
Dong, Y., Chen, S., Wang, Z., Peng, N.,
and Yu, J. (2013) Synergy of ultrasound
microbubbles and vancomycin against
Staphylococcus epidermidis biofilm. J.
Antimicrob. Chemother., 68, 816–826.
Xu, J., Bigelow, T.A., Halverson, L.J.,
Middendorf, J., and Rusk, B. (2012)
Mechanical destruction of pseudomonas aeruginosa biofilms by
ultrasound exposure. 11th International
Symposium on Therapeutic Ultrasound
(eds R., Muratore and E.E. Konofagou),
pp. 463–468.
Bigelow, T.A., Northagen, T., Hill, T.M.,
and Sailer, F.C. (2009) The destruction of Escherichia coli biofilms using
high-intensity focused ultrasound.
Ultrasound Med. Biol., 35, 1026–1031.
Pitt, W.G., McBride, M.O., Lunceford,
J.K., Roper, R.J., and Sagers, R.D. (1994)
Ultrasonic enhancement of antibiotic
action on Gram-negative bacteria.
Antimicrob. Agents Chemother., 38,
2577–2582.
Xu, J., Bigelow, T.A., Halverson, L.J.,
Middendorf, J.M., and Rusk, B. (2012)
Minimization of treatment time for in
vitro 1.1 MHz destruction of Pseudomonas aeruginosa biofilms by
high-intensity focused ultrasound.
Ultrasonics, 52, 668–675.
Donlan, R.M. (2002) Biofilms: microbial
life on surfaces. Emerg. Infect. Dis., 8,
881–890.
Kostakioti, M., Hadjifrangiskou, M., and
Hultgren, S.J. (2013) Bacterial biofilms:
development, dispersal, and therapeutic
strategies in the dawn of the postantibiotic era. Cold Spring Harbor Perspect.
Med., 3, a010306.
108. Torley, P.J. and Bhandari, B.R. (2007)
109.
110.
111.
112.
113.
114.
115.
116.
117.
in Handbook of Food Preservation (ed.
M.S. Rahman), CRC Press, Boca Raton,
FL, pp. 713–732.
Villamiel, M. and De Jong, P. (2000)
Inactivation of Pseudomonas fluorescens and Streptococcus thermophilus in Trypticase Soy Broth and
total bacteria in milk by continuousflow ultrasonic treatment and conventional heating. J. Food Eng., 45,
171–179.
Phoenix, D.A., Dennison, S.R., and
Harris, F. (2013) Antimicrobial Peptides, Wiley-VCH Verlag GmbH & Co.
KGaA, pp. 145–180.
Silva, M.P., Pereira, C.A., Junqueira,
J.C., and Jorge, A.O.C. (2013) Methods of destroying bacterial spores,
in Microbial Pathogens and Strategies for Combating Them: Science,
Technology and Education (ed. A.
Méndez-Vilas), Formatex Research
Center, www.formatex.org (accessed 07
April 2014).
Ahmed, F.I.K. and Russell, C. (1975)
Synergism between ultrasonic waves
and hydrogen peroxide in the killing of
micro-organisms. J. Appl. Bacteriol., 39,
31–41.
Rastogi, N.K. (2011) Opportunities and
challenges in application of ultrasound
in food processing. Crit. Rev. Food Sci.
Nutr., 51, 705–722.
Demirdöven, A. and Baysal, T. (2008)
The use of ultrasound and combined
technologies in food preservation. Food
Rev. Int., 25, 1–11.
Naddeo, V., Landi, M., Belgiorno, V.,
and Napoli, R.M.A. (2009) Wastewater
disinfection by combination of ultrasound and ultraviolet irradiation. J.
Hazard. Mater., 168, 925–929.
Char, C.D., Mitilinaki, E.,
Norma Guerrero, S., and
Maris Alzamora, S. (2010) Use of
high-intensity ultrasound and UV-C
light to inactivate some microorganisms
in fruit juices. Food Bioprocess Technol.,
3, 797–803.
Birmpa, A., Sfika, V., and Vantarakis, A.
(2013) Ultraviolet light and Ultrasound as non-thermal treatments for
the inactivation of microorganisms in
References
118.
119.
120.
121.
122.
123.
124.
125.
126.
fresh ready-to-eat foods. Int. J. Food
Microbiol., 167, 96–102.
Palgan, I., Munoz, A., Noci, F., Whyte,
P., Morgan, D.J., Cronin, D.A., and
Lyng, J.G. (2012) Effectiveness of combined Pulsed Electric Field (PEF) and
Manothermosonication (MTS) for
the control of Listeria innocua in a
smoothie type beverage. Food Control,
25, 621–625.
Huang, E., Mittal, G.S., and Griffiths,
M.W. (2006) Inactivation of Salmonella
enteritidis in liquid whole egg using
combination treatments of pulsed electric field, high pressure and ultrasound.
Biosystems Eng., 94, 403–413.
Chapman, J.S., Ferguson, R., Consalo,
C., and Bliss, T. (2013) Bacteriostatic
effect of sequential hydrodynamic and
ultrasound-induced stress. J. Appl.
Microbiol., 114, 947–955.
Wong, E., Vaillant-Barka, F., and
Chaves-Olarte, E. (2012) Synergistic
effect of sonication and high osmotic
pressure enhances membrane damage and viability loss of Salmonella
in orange juice. Food Res. Int., 45,
1072–1079.
Alvarez, I., Manas, P., Virto, R., and
Condon, S. (2006) Inactivation of
Salmonella Senftenberg 775W by ultrasonic waves under pressure at different
water activities. Int. J. Food Microbiol.,
108, 218–225.
Manas, P., Pagan, R., Raso, J., Sala, F.J.,
and Condon, S. (2000) Inactivation
of Salmonella enteritidis, Salmonella
Typhimurium, and Salmonella Senftenberg by ultrasonic waves under
pressure. J. Food Prot., 63, 451–456.
Manas, P., Pagan, R., and Raso, J. (2000)
Predicting lethal effect of ultrasonic
waves under pressure treatments on
Listeria monocytogenes ATCC 15313
by power measurements. J. Food Sci.,
65, 663–667.
Condón, S., Mañas, P., and Cebrián, G.
(2011) in Ultrasound Technologies for
Food and Bioprocessing (eds H. Feng,
G. Barbosa-Canovas, and J. Weiss),
Springer, New York, pp. 287–319.
Bermudez-Aguirre, D. and
Barbosa-Canovas, G.V. (2012) Inactivation of Saccharomyces cerevisiae in
127.
128.
129.
130.
131.
132.
133.
134.
135.
pineapple, grape and cranberry juices
under pulsed and continuous thermosonication treatments. J. Food Eng.,
108, 383–392.
Herceg, Z., Jambrak, A.R., Lelas, V., and
Thagard, S.M. (2012) The effect of high
intensity ultrasound treatment on the
amount of staphylococcus aureus and
Escherichia coli in milk. Food Technol.
Biotechnol., 50, 46–52.
Wordon, B.A., Mortimer, B., and
McMaster, L.D. (2012) Comparative
real-time analysis of Saccharomyces
cerevisiae cell viability, injury and
death induced by ultrasound (20 kHz)
and heat for the application of hurdle
technology. Food Res. Int., 47, 134–139.
Pagan, R., Manas, P., Alvarez, I., and
Condon, S. (1999) Resistance of Listeria monocytogenes to ultrasonic
waves under pressure at sublethal
(manosonication) and lethal (manothermosonication) temperatures. Food
Microbiol., 16, 139–148.
Schultz, A.C., Uhrbrand, K.,
Norrung, B., and Dalsgaard, A. (2012)
Inactivation of norovirus surrogates
on surfaces and raspberries by steamultrasound treatment. J. Food Prot., 75,
376–381.
Jyoti, K.K. and Pandit, A.B. (2003)
Hybrid cavitation methods for water
disinfection: simultaneous use of
chemicals with cavitation. Ultrason.
Sonochem., 10, 255–264.
Dahi, E. (1976) Physicochemical aspects
of disinfection of water by means of
ultrasound and ozone. Water Res., 10,
677–684.
Jyoti, K.K. and Pandit, A.B. (2004)
Ozone and cavitation for water disinfection. Biochem. Eng. J., 18, 9–19.
Gogate, P.R., Mededovic-Thagard, S.,
McGuire, D., Chapas, G., Blackmon,
J., and Cathey, R. (2014) Hybrid reactor based on combined cavitation and
ozonation: from concept to practical reality. Ultrason. Sonochem., 21,
590–598.
Taran, V.S., Garkusha, I.E., Krasnyj,
V.V., Lozina, A.S., Muratov, R.M.,
Shulaev, V.M., and Timoshenko, A.I.
349
350
11
136.
137.
138.
139.
140.
141.
142.
143.
The Antimicrobial Effects of Ultrasound
(2012) Recent developments of plasmabased technologies for medicine and
industry. Nukleonika, 57, 277–282.
Chen, C.W., Lee, H.M., Chen, S.H.,
Chen, H.L., and Chang, M.B. (2009)
Ultrasound-assisted plasma: a novel
technique for inactivation of aquatic
microorganisms. Environ. Sci. Technol.,
43, 4493–4497.
Knorr, D., Froehling, A., Jaeger,
H., Reineke, K., Schlueter, O., and
Schoessler, K. (2011) Emerging technologies in food processing. Annu. Rev.
Food Sci. Technol., 2, 203–235.
Ayyildiz, O., Sanik, S., and Ileri, B.
(2011) Effect of ultrasonic pretreatment on chlorine dioxide disinfection
efficiency. Ultrason. Sonochem., 18,
683–688.
Brilhante Sao Jose, J.F. and
Dantas Vanetti, M.C. (2012) Effect
of ultrasound and commercial sanitizers in removing natural contaminants
and Salmonella enterica Typhimurium
on cherry tomatoes. Food Control, 24,
95–99.
Chen, Z. and Zhu, C. (2011) Combined
effects of aqueous chlorine dioxide and
ultrasonic treatments on postharvest
storage quality of plum fruit (Prunus
salicina L.). Postharvest Biol. Technol.,
61, 117–123.
Sagong, H.-G., Lee, S.-Y., Chang, P.S., Heu, S., Ryu, S., Choi, Y.-J., and
Kang, D.-H. (2011) Combined effect of
ultrasound and organic acids to reduce
Escherichia coli O157:H7, Salmonella
Typhimurium, and Listeria monocytogenes on organic fresh lettuce. Int. J.
Food Microbiol., 145, 287–292.
Zhou, B., Feng, H., and Luo, Y. (2009)
Ultrasound enhanced sanitizer efficacy in reduction of Escherichia coli
O157:H7 population on spinach leaves.
J. Food Sci., 74, M308–M313.
Oulahal, N., Martial-Gros, A., Bonneau,
M., and Blum, L.J. (2004) Combined effect of chelating agents and
ultrasound on biofilm removal from
stainless steel surfaces. Application to
“Escherichia coli milk” and “Staphylococcus aureus milk” biofilms. Biofilms,
1, 65–73.
144. Oulahal, N., Martial-Gros, A., Bonneau,
145.
146.
147.
148.
149.
150.
151.
152.
153.
M., and Blum, L.J. (2007) Removal of
meat biofilms from surfaces by ultrasounds combined with enzymes and/or
a chelating agent. Innovative Food Sci.
Emerg. Technol., 8, 192–196.
Nazari, S.H. and Weiss, J. (2010) Evidence of antimicrobial activity of date
fruits in combination with high intensity ultrasound. Afr. J. Microbiol. Res.,