Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Biometals in Neurodegenerative Diseases: Mechanisms and Therapeutics
Biometals in Neurodegenerative Diseases: Mechanisms and Therapeutics
Biometals in Neurodegenerative Diseases: Mechanisms and Therapeutics
Ebook1,177 pages8 hours

Biometals in Neurodegenerative Diseases: Mechanisms and Therapeutics

Rating: 5 out of 5 stars

5/5

()

Read preview

About this ebook

Biometals in Neurodegenerative Diseases: Mechanisms and Therapeutics is an authoritative and timely resource bringing together the major findings in the field for ease of access to those working in the field or with an interest in metals and their role in brain function, disease, and as therapeutic targets. Chapters cover metals in Alzheimer’s Disease, Parkinson’s Disease, Motor Neuron Disease, Autism and lysosomal storage disorders.

This book is written for academic researchers, clinicians and advanced graduate students studying or treating patients in neurodegeneration, neurochemistry, neurology and neurotoxicology. The scientific literature in this field is advancing rapidly, with approximately 300 publications per year adding to our knowledge of how biometals contribute to neurodegenerative diseases.

Despite this rapid increase in our understanding of biometals in brain disease, the fields of biomedicine and neuroscience have often overlooked this information. The need to bring the research on biometals in neurodegeneration to the forefront of biomedical research is essential in order to understand neurodegenerative disease processes and develop effective therapeutics.

  • Authoritative and timely resource bringing together the major findings in the field for those with an interest in metals and their role in the brain function, disease, and as therapeutic targets
  • Written for academic researchers, clinicians, and advanced graduate students studying, or treating, patients in neurodegeneration, neurochemistry, neurology and neurotoxicology
  • Edited by international leaders in the field who have contributed greatly to the study of metals in neurodegenerative diseases
LanguageEnglish
Release dateApr 28, 2017
ISBN9780128045633
Biometals in Neurodegenerative Diseases: Mechanisms and Therapeutics

Related to Biometals in Neurodegenerative Diseases

Related ebooks

Medical For You

View More

Related articles

Reviews for Biometals in Neurodegenerative Diseases

Rating: 5 out of 5 stars
5/5

1 rating0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Biometals in Neurodegenerative Diseases - Anthony R. White

    Biometals in Neurodegenerative Diseases

    Mechanisms and Therapeutics

    Edited by

    Anthony R. White

    Cell and Molecular Biology

    QIMR Berghofer Medical Research Institute,

    Herston, QLD, Australia

    Michael Aschner

    Department of Molecular Pharmacology

    Albert Einstein College of Medicine

    Bronx, NY, United States

    Lucio G. Costa

    Department of Environmental and Occupational

    Health Sciences University of Washington

    Seattle, WA, United States

    Ashley I. Bush

    Florey Institute of Neuroscience and Mental Health

    University of Melbourne

    Parkville, VIC, Australia

    Table of Contents

    Cover

    Title page

    Copyright

    Contributors

    Preface

    Chapter 1: Biometals and Alzheimer’s Disease

    Abstract

    Introduction

    The Role of Copper in AD

    The Role of Zinc in AD

    The Role of Iron in AD

    Therapeutic Targeting of Biometals in AD

    Conclusions

    Chapter 2: Copper in Alzheimer’s Disease

    Abstract

    Introduction

    The Physiology of Copper

    Copper Toxicity

    Conclusions

    Chapter 3: The Role of Selenium in Neurodegenerative Diseases

    Abstract

    Introduction

    Selenoproteins and the Selenoproteome

    Selenium and Alzheimer’s Disease

    Parkinson’s Disease

    Other Neurodegenerative Diseases

    Conclusions

    Chapter 4: Does HFE Genotype Impact Macrophage Phenotype in Disease Process and Therapeutic Response?

    Abstract

    Iron

    Hemochromatosis

    HFE

    Macrophages

    HFE Animal Models

    Conclusions

    Chapter 5: Chemical Elements and Oxidative Status in Neuroinflammation

    Abstract

    Introduction

    Metal-Induced Neurotoxicity and Multiple Sclerosis

    Metals and Oxidative Status in Multiple Sclerosis

    Metals and Oxidative Status in Clinically Isolated Syndromes

    Conclusions

    Chapter 6: Metals and Neuroinflammation

    Abstract

    Introduction

    Mechanisms by Which Metal Elements Can Incite Immune Activity

    The Relation Between Reactive Oxygen and Nitrogen Species and Inflammation

    Conclusions

    Chapter 7: Metals and Prions: Twenty Years of Mining the Awe

    Abstract

    Prion Diseases

    Prion Protein

    Prion Protein Function

    Copper and PrP

    Zinc and PrP

    Iron and PrP

    Manganese and PrP

    Metals in Prion Disease

    Chelation Therapy and Prion Disease

    Conclusions

    Chapter 8: Manganese and Neurodegeneration

    Abstract

    Background

    Mn Essentiality and Metabolic Functions

    Mn Biokinetics and Homeostatic Control

    Neurotoxicology of Mn

    Biomonitoring of Mn in Patients Undergoing PN

    Conclusions

    Acknowledgment

    Chapter 9: Zinc in Autism

    Abstract

    Introduction

    Zinc Signaling in Autism

    Therapeutic Strategies in Autism Based on Biometals

    Conclusions

    Chapter 10: Metals and Motor Neuron Disease

    Abstract

    List of Abbreviations

    Introduction

    Metal Exposure

    Metals in ALS Cerebrospinal Fluid

    Metals in ALS

    Protection by Metallothionein

    Metal Distribution in ALS

    Genetic Aspects

    Concluding Remarks

    Chapter 11: Metals and Lysosomal Storage Disorders

    Abstract

    Introduction

    Common Pathological Features of Lysosomal Storage Disorders

    Description of Most Common Neurodegenerative LSDs Associated with Biometal Imbalance

    Function and Regulation of Biometals

    Role of Biometals and Biometal Binding Proteins in LSDs

    Targeting Metals to Treat Disease

    Chapter 12: Developmental Exposure to Metals and its Contribution to Age-Related Neurodegeneration

    Abstract

    Introduction

    Developmental Exposure to Toxicants and Late Effects

    Developmental Lead Exposure and Alzheimer's Disease

    Developmental Arsenic Exposure and Alzheimer's Disease

    Conclusions and Future Perspectives

    Acknowledgment

    Chapter 13: Metal Biology Associated with Huntington’s Disease

    Abstract

    Introduction

    The Epidemiology of HD

    The Symptoms of HD

    The Neuropathology of HD

    Biological Function of Wild-type and Pathogenic HTT Proteins

    Autophagy and Metals in Huntington’s Disease

    Exosomes and Metal in Huntington’s Disease

    Environmental Factors Impacting HD

    Metals in HD

    Iron in HD

    Copper in HD

    Calcium in HD

    Manganese in HD

    Manganese Deposition: Brain Regions, Cell Types, and Cellular Organelles

    Manganese Dyshomeostasis in HD

    Mn-Dependent and Mn-Utilizing Enzymes

    Intracellular pH and Metal Biology in HD

    Metal-Related Clinical Interventions in HD

    Conclusions and Future Directions

    Chapter 14: Metal-Binding to Amyloid-β Peptide: Coordination, Aggregation, and Reactive Oxygen Species Production

    Abstract

    Introduction

    Structure of the Metal-Aβ Complexes

    Affinity of Metals to Aβ

    Aggregation

    Reactive Oxygen Species Induced Oxidative Stress

    Conclusions

    Acknowledgments

    Chapter 15: Metals and Mitochondria in Neurodegeneration

    Abstract

    Introduction

    Iron Dyshomeostasis

    Copper Dislocation

    Zinc Deficiency

    Mitochondrial Dysfunction

    Conclusions

    Acknowledgments

    Chapter 16: Metal Transporters in Neurodegeneration

    Abstract

    Iron Transporters and Neurodegeneration

    Zinc Transporters and Neurodegeneration

    Copper Transporters and Neurodegeneration

    Manganese Transporters and Neurodegeneration

    Magnesium Transporters and Neurodegeneration

    Aluminum Transporters and Neurodegeneration

    Conclusions

    Chapter 17: Metal Imaging in the Brain

    Abstract

    Introduction

    Introduction to MRI Physics

    MRI Contrast Agents

    Gadolinium

    Iron

    Copper

    Manganese

    Chapter 18: Metalloregulation of Protein Clearance: New Therapeutic Avenues for Neurodegenerative Diseases

    Abstract

    Introduction

    Metalloregulation of the Ubiquitin Proteasome System: Implication in Neurodegenerative Diseases

    Metals as Mediators of Autophagy-Lysosomal Response

    Conclusions

    Acknowledgments

    Chapter 19: Metals and Autophagy in Neurotoxicity

    Abstract

    Introduction

    Part 1 Metal-Related Neurotoxicity and Neurodegenerative Diseases

    Part 2 Autophagy in Metal Neurotoxicity

    Conclusions

    Acknowledgments

    Chapter 20: An Overview of Multifunctional Metal Chelators as Potential Treatments for Neurodegenerative Diseases

    Abstract

    Introduction

    Parent Metal Chelators

    Multifunctional Metal Chelators

    Conclusions

    Chapter 21: Abnormal Function of Metalloproteins Underlies Most Neurodegenerative Diseases

    Abstract

    Background

    Biometals

    Abnormal Biometal Levels and Distribution Underlie Most Forms of Neurodegeneration

    Abnormal Metalloprotein Function Underlying Neurodegenerative Diseases

    Neurodegenerative Diseases Caused by Mutation in Metalloproteins

    Neurodegenerative Diseases Associated With Abnormal Metalloprotein Function

    Neurodegenerative Diseases Involving Biometal Changes but Without a Clearly Identified Role for Metalloprotein Abnormities

    Neurodegenerative Diseases Where No Major Role for Biometals or Metalloproteins Has Yet Been Identified

    Conclusions

    Index

    Copyright

    Academic Press is an imprint of Elsevier

    125 London Wall, London EC2Y 5AS, United Kingdom

    525 B Street, Suite 1800, San Diego, CA 92101-4495, United States

    50 Hampshire Street, 5th Floor, Cambridge, MA 02139, United States

    The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, United Kingdom

    Copyright © 2017 Elsevier Inc. All rights reserved.

    No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying, recording, or any information storage and retrieval system, without permission in writing from the publisher. Details on how to seek permission, further information about the Publisher’s permissions policies and our arrangements with organizations such as the Copyright Clearance Center and the Copyright Licensing Agency, can be found at our website: www.elsevier.com/permissions.

    This book and the individual contributions contained in it are protected under copyright by the Publisher (other than as may be noted herein).

    Notices

    Knowledge and best practice in this field are constantly changing. As new research and experience broaden our understanding, changes in research methods, professional practices, or medical treatment may become necessary.

    Practitioners and researchers must always rely on their own experience and knowledge in evaluating and using any information, methods, compounds, or experiments described herein. In using such information or methods they should be mindful of their own safety and the safety of others, including parties for whom they have a professional responsibility.

    To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume any liability for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas contained in the material herein.

    Library of Congress Cataloging-in-Publication Data

    A catalog record for this book is available from the Library of Congress

    British Library Cataloguing-in-Publication Data

    A catalogue record for this book is available from the British Library

    ISBN: 978-0-12-804562-6

    For information on all Academic Press publications visit our website at https://www.elsevier.com/books-and-journals

    Publisher: Mara Conner

    Acquisition Editor: Melanie Tucker

    Editorial Project Manager: Kristi Anderson

    Production Project Manager: Karen East and Kirsty Halterman

    Designer: Victoria Pearson

    Typeset by Thomson Digital

    Contributors

    Alessandro Alimonti,     National Institute of Health, Rome, Italy

    Michael Aschner,     Albert Einstein College of Medicine, Bronx, NY, United States

    Terry Jo V. Bichell

    Vanderbilt Brain Institute

    Vanderbilt University Medical Center, Nashville, TN, United States

    Stephen C. Bondy,     Center for Occupational and Environmental Health, University of California, Irvine, CA, United States

    Valentina Borghesani

    CNRS, LCC (Laboratory of Chemical Coordination)

    University of Toulouse, Toulouse, France

    Aaron B. Bowman

    Vanderbilt Brain Institute

    Vanderbilt University Medical Center, Nashville, TN, United States

    Sonia Brescianini,     Center for Epidemiology, Surveillance and Health Promotion, National Institute of Health, Rome, Italy

    David R. Brown,     University of Bath, Bath, United Kingdom

    Maria C. Buscarinu,     Center for Experimental Neurological Therapies, S. Andrea Hospital, Sapienza University of Rome, Rome, Italy

    Ashley I. Bush,     The Florey Institute of Neuroscience and Mental Health, The University of Melbourne, Parkville, VIC, Australia

    Bárbara R. Cardoso

    The Florey Institute of Neuroscience and Mental Health, The University of Melbourne, Parkville, VIC, Australia

    University of São Paulo, São Paulo, Brazil

    Benedetta Cerasoli,     Center for Experimental Neurological Therapies, S. Andrea Hospital, Sapienza University of Rome, Rome, Italy

    Jingyuan Chen,     Fourth Military Medical University, Xi’an, China

    James R. Connor,     The Pennsylvania State University College of Medicine, Hershey, PA, United States

    Lucio G. Costa

    University of Washington, Seattle, WA, United States

    University of Parma, Parma, Italy

    Peter J. Crouch,     University of Melbourne, Melbourne, VIC, Australia

    Melisa del Barrio

    CNRS, LCC (Laboratory of Chemical Coordination)

    University of Toulouse, Toulouse, France

    David C. Dorman,     North Carolina State University, Raleigh, NC, United States

    Peter Faller,     Biometals and Biological Chemistry, Institute of Chemistry, UMR 7177, University of Strasbourg, Strasbourg, France

    Michela Ferraldeschi,     Center for Experimental Neurological Therapies, S. Andrea Hospital, Sapienza University of Rome, Rome, Italy

    Arianna Fornasiero,     Center for Experimental Neurological Therapies, S. Andrea Hospital, Sapienza University of Rome, Rome, Italy

    Andreas M. Grabrucker

    Institute for Anatomy and Cell Biology

    WG molecular Analysis of Synaptopathies, Ulm University, Ulm, Germany

    Stefanie Grabrucker,     Institute for Anatomy and Cell Biology, Ulm University, Ulm, Germany

    Mark A. Greenough,     The Florey Institute of Neuroscience and Mental Health, The University of Melbourne, Melbourne, VIC, Australia

    Timothy C. Halbesma,     Vanderbilt University Medical Center, Nashville, TN, United States

    Dominic J. Hare

    The Florey Institute of Neuroscience and Mental Health, The University of Melbourne, Parkville, VIC

    University of Technology Sydney, Broadway, NSW, Australia

    Christelle Hureau

    CNRS, LCC (Laboratory of Chemical Coordination)

    University of Toulouse, Toulouse, France

    Hong Jiang

    Medical College of Qingdao University

    Shandong Provincial Collaborative Innovation Center for Neurodegenerative Disorders, Qingdao University, Qingdao, China

    Katja M. Kanninen,     A.I. Virtanen Institute for Molecular Sciences, University of Eastern Finland, Kuopio, Finland

    Henna Konttinen,     A.I. Virtanen Institute for Molecular Sciences, University of Eastern Finland, Kuopio, Finland

    Katarína Lejavová,     A.I. Virtanen Institute for Molecular Sciences, University of Eastern Finland, Kuopio, Finland

    Frank W. Lewis,     Northumbria University Newcastle, Newcastle, United Kingdom

    Wenjing Luo,     Fourth Military Medical University, Xi’an, China

    Tarja Malm,     A.I. Virtanen Institute for Molecular Sciences, University of Eastern Finland, Kuopio, Finland

    Dinamene Marques dos Santos,     University of Lisbon, Lisboa, Portugal

    Ana P. Marreilha dos Santos,     University of Lisbon, Lisboa, Portugal

    Carlo Mattei,     Center for Experimental Neurological Therapies, S. Andrea Hospital, Sapienza University of Rome, Rome, Italy

    Rosella Mechelli,     Center for Experimental Neurological Therapies, S. Andrea Hospital, Sapienza University of Rome, Rome, Italy

    Alexandra I. Mot,     University of Melbourne, Melbourne, VIC, Australia

    Anne M. Nixon,     The Pennsylvania State University College of Medicine, Hershey, PA, United States

    Carlos M. Opazo,     The Florey Institute of Neuroscience and Mental Health, The University of Melbourne, Melbourne, VIC, Australia

    George Perry,     The University of Texas at San Antonio (UTSA), San Antonio, TX, United States

    Anna Pino,     National Institute of Health, Rome, Italy

    Germán Plascencia-Villa,     The University of Texas at San Antonio (UTSA), San Antonio, TX, United States

    Alejandra Ramírez Muñoz,     The Florey Institute of Neuroscience and Mental Health, The University of Melbourne, Melbourne, VIC, Australia

    Giovanni Ristori,     Center for Experimental Neurological Therapies, S. Andrea Hospital, Sapienza University of Rome, Rome, Italy

    Silvia Romano,     Center for Experimental Neurological Therapies, S. Andrea Hospital, Sapienza University of Rome, Rome, Italy

    Per M. Roos,     Karolinska Institutet, Institute of Environmental Medicine, Stockholm, Sweden

    Carlo Salustri,     Institute of Cognitive Sciences and Technologies (CNR), Fatebenefratelli Hospital, Rome, Italy

    Marco Salvetti,     Center for Experimental Neurological Therapies, S. Andrea Hospital, Sapienza University of Rome, Rome, Italy

    Mariacristina Siotto,     Don Carlo Gnocchi Foundation ONLUS, Italy

    Rosanna Squitti,     IRCCS, Institute Center St. John of God Fatebenefratelli, Brescia, Italy

    Maria A. Stazi,     National Institute of Health, Rome, Italy

    Peng Su,     Fourth Military Medical University, Xi’an, China

    David Tétard,     Northumbria University Newcastle, Newcastle, United Kingdom

    K. Grace Tipps,     Vanderbilt University Medical Center, Nashville, TN, United States

    Mariacarla Ventriglia,     Institute of Cognitive Sciences and Technologies (CNR), Fatebenefratelli Hospital, Rome, Italy

    Anthony R. White,     QIMR Berghofer Medical Research Institute, Herston, QLD, Australia

    Miguel José-Yacamán,     The University of Texas at San Antonio (UTSA), San Antonio, TX, United States

    Preface

    Biological metals (biometals) have key functions in the brain but can also induce degenerative changes due to abnormalities in homeostatic mechanisms. The scientific literature in this field is advancing rapidly with approximately 300 publications per year adding to our knowledge of how biometals contribute to neurodegenerative diseases, such as Alzheimer’s disease, Parkinson’s disease, motor neuron disease, and others. Despite this rapid increase in our understanding of biometals in brain disease, the fields of biomedicine and neuroscience have often overlooked this information. The need to bring the research on biometals in neurodegeneration to the forefront of biomedical research is essential if we are to understand neurodegenerative disease processes and develop effective therapeutics. There are currently few sources of consolidated research on biometals and neurodegeneration that are available to biomedical researchers, clinicians, students, and others. This book on biometals in neurodegeneration provides an authoritative and timely resource to bring together the major findings in the field for ease of access to those working in neuroscience or biomedicine, or with an interest in metals and their role in the brain function, disease, and as therapeutic targets.

    Overview of chapters included in Biometals in Neurodegenerative Diseases: Mechanisms and Therapeutics.

    Alzheimer’s disease is the leading form of neurodegeneration contributing to the majority of an estimated 36 million cases of dementia worldwide. With an ageing world population, estimates for global Alzheimer’s disease prevalence are in the order of 115 million patients by 2050. It is therefore fitting that the opening chapter of this book covers biometals in Alzheimer’s disease. Mot and Crouch (Chapter 1, pp. 1–18) provide a comprehensive insight into the contribution of biometals including copper, zinc, and iron in several major pathological features of Alzheimer’s disease including amyloid peptide aggregation and tau phosphorylation. The chapter finishes with an insight into where therapeutic approaches to biometal modulation in Alzheimer’s disease may progress in the future. Staying with the role of biometals in Alzheimer’s disease, Squitti et al. (Chapter 2, pp. 19–34) provides a deeper insight into the role of copper in this disorder, describing the pathways of copper metabolism and brain copper handling as well as interesting aspects of copper toxicity and the role of ceruloplasmin and nonceruloplasmin copper in Alzheimer’s disease. This is followed by a very interesting insight into a biometal that receives less coverage than it should, selenium and its role in neurodegeneration. This biometal forms a critical inorganic component of the amino acid selenocysteine, which is involved in at least 25 different proteins. Abnormal selenium stasis is found in many forms of leading neurodegenerative diseases, and selenium-based therapeutics have the potential to make significant impacts in these disorders as discussed by Cardoso et al. (Chapter 3, pp. 35–50).

    The subsequent three chapters by Nixon and Connor (Chapter 4, pp. 51–66), Ferraldeschi et al. (Chapter 5, pp. 67–82) and Stephen Bondy (Chapter 6, pp. 83–94) cover important aspects of biometals in inflammation and how this modulates outcomes in neurodegeneration. Nixon and Connor et al. describe how the high iron (Fe) gene, HFE genotype, affects macrophage phenotype in disease. Mutations in HFE appear to alter macrophage (and microglial) iron distribution with potentially important outcomes for neurodegeneration. Ferraldeschi et al. follow this with an expanded insight into the role of biometals in oxidative-mediated neuroinflammatory processes, with a more detailed focus on biometal effects in multiple sclerosis. Bondy then provides a valuable account of the various mechanisms leading to biometal-mediated neuroinflammation, including the formation of haptens, the production of reactive oxygen and nitrogen species, the sequestering of reducing capacity and the formation of inflammation-provoking colloids.

    David Brown (Chapter 7, pp. 95–116) brings us a very informative and insightful account of 20 years of metals research in prion diseases, dispelling some common myths and providing a comprehensive account of where we have come to in this field. Iron, zinc, copper, and manganese all have key roles to play in prion protein stasis in diseases, such as Creutzfeld–Jakob disease. Staying with the biometal, manganese, Marques dos Santos et al. (Chapter 8, pp. 117–152) provide an in-depth coverage of this element in neurodegeneration, providing valuable information on the role of manganese in neurobiology, environmental exposure in humans, and subsequent pathways to neurotoxicity and neurodegeneration.

    Although autism spectrum disorder (ASD) is not classically categorized as a neurodegenerative disorder, there are shared comorbidities between ASD and neurodegenerative disorders including altered biometal stasis in the brain. Grabrucker and Grabrucker (Chapter 9, pp. 153–174) delve into the role of these biometal changes in ASDs, providing an interesting account of how biometals can affect neuro-synaptic functions resulting in the features that characterize ASD. Although a rare disorder, motor neuron disease (MND) has been the center of recent major international funding efforts. The disease is rapid with most cases fatal in 2–5 years after onset, and no long-term effective treatment exists. Although the cause in most cases remains unknown, Per Roos (Chapter 10, pp. 175–194) describes a key role for possible environmental exposure to biometals including lead, manganese, and mercury in disease etiology, and how these metals may contribute to neuropathological changes. In the last of the chapters covering the general role of biometals in neurodegeneration, Konttinen et al. (Chapter 11, pp. 195–216) cover important aspects of biometals in another group of rare disorders, the lysosomal storage diseases (LSDs). Lysosomes are key sites of biometal homeostasis, and abnormalities in lysosomal function as occurs in LSDs, leads to significant biometal abnormalities with potentially major impacts leading to neurodegeneration in these disorders.

    The second half of this volume covers molecular and cellular aspects of metals in brain health, dysfunction, and neurotoxicity. Lucio G. Costa (Chapter 12, pp. 217–230) starts this section with a review of how environmental exposure to metals, such as lead and arsenic in early life can affect outcomes in later life neurodegeneration including Alzheimer’s disease. The mechanism of this predisposition is unclear but could be related to epigenetic changes induced by biometals. Bichell et al. (Chapter 13, pp. 231–264) follow this with an account of metal biology in Huntington’s disease including potential iron, copper, and manganese interactions with a range of homeostatic enzymes and proteins in the brain. Del Barrio et al. (Chapter 14, pp. 265–282) return us to Alzheimer’s disease again with a critical insight into the fundamental copper coordination by amyloid peptide, the major protein form deposited in Alzheimer’s brains. The review describes how copper drives aggregation and reactive oxygen species generation through interaction with the amyloid peptide.

    Delving deeper into the cells of the brain, Plascencia-Villa et al. (Chapter 15, pp. 283–312) explore the role of biometals in mitochondrial function and how this contributes to a range of neurodegenerative disorders. Abnormalities in biometal handling in the cell powerhouse leads to major outcomes including reactive oxygen species formation, abnormal cell signaling, and altered energy production with important consequences for highly metabolic neurons. Another important aspect of biometal involvement in neurodegeneration is biometal transportation. Changes to key biometal transporters in the brain can have profound effects on the handling and action of copper, zinc, iron, manganese etc. And conversely, altered metal levels can affect transporter expression and localization. Hong Jiang (Chapter 16, pp. 313–348) provides a comprehensive coverage of metal transporters for the main neuro-metals and how changes to these transporters can underlie neurodegenerative outcomes.

    One of the most important techniques used to understand how biometals contribute to neurodegeneration is metal imaging. David Dorman (Chapter 17, pp. 349–362) describes one of the central biometal imaging techniques applied to neurodegenerative disease, magnetic resonance imaging (MRI), and how it can be used to understand and differentiate the roles of iron, manganese, and copper in brain disorders. Brain protein accumulation is a major feature of most neurodegenerative diseases and is often associated with impaired mechanisms for clearance of aggregated or abnormally folded proteins. Ramírez Muñoz et al. (Chapter 18, pp. 363–376) describe how metals play a fundamental role in protein clearance and how this can be affected by abnormal metal homeostasis, contributing further to neurodegenerative disease pathology. Related to this, Su et al. (Chapter 19, pp. 377–398) contribute a key insight to the role of metals in autophagic processes. A key cellular mechanism for clearance of unwanted cell material, autophagy is a complex process involving many proteins and subcellular compartments. Studies have found major impairments to autophagy in neurodegenerative diseases and this is now a common putative therapeutic target. Su et al., explore the role of metals in autophagy and how abnormalities to metal homeostasis can contribute to pathological autophagic changes.

    Of course, one of the major reasons for increasing our understanding of biometals in neurodegeneration is to develop therapeutic approaches for these disorders. In the penultimate chapter, Lewis and Tétard (Chapter 20, pp. 399–414) provide us with a comprehensive overview of metal chelators as potential new treatments for neurodegeneration. The future of these therapeutics appears to be in development of multifunctional agents that target metals and additional pathological features of neurodegeneration, such as oxidative stress. Time will tell if this approach provides much needed advances in neurodegenerative disease therapeutics. The final chapter by Kanninen and White (Chapter 21, pp. 415–438) then takes an overarching view of biometals in neurodegeneration and provides a review of the role for metalloproteins in many forms of neurodegenerative disease, leading us to consider whether neurodegenerative diseases should be categorized as metallopathies.

    Whether student, clinician, or specialized biometals researcher, we hope that the reader will be able to gain exciting new insights into biometals and neurodegeneration. We believe that this volume will have an important place in the medical literature and provide a valuable reference source for many years in this major field of neurodegenerative disease research.

    Anthony R. White

    Michael Aschner

    Lucio G. Costa

    Ashley I. Bush

    Chapter 1

    Biometals and Alzheimer’s Disease

    Alexandra I. Mot

    Peter J. Crouch    University of Melbourne, Melbourne, VIC, Australia

    Abstract

    Maintaining metal ion homeostasis is essential for diverse cellular processes, particularly in the central nervous system. Changes to the transition metals, copper, zinc, and iron in Alzheimer’s disease (AD) have been extensively reported, especially in the brain, blood, and cerebrospinal fluid. Through interactions with disease-associated proteins, including amyloid precursor protein, the proteolytically cleaved peptide amyloid-beta (Aβ), and tau, these biometals play an important role in disease pathogenesis. Currently, no disease modifying therapy exists for AD and clinical trials over the past decade aimed at targeting established disease mechanisms have failed to alter disease progression. Targeting biometals in AD by restoring metal ion homeostasis represents an alternative potential therapeutic avenue.

    Keywords

    Alzheimer’s disease

    biometals

    copper

    zinc

    iron

    amyloid precursor protein

    amyloid-β

    tau

    metal-based therapies

    Outline

    Introduction

    The Role of Copper in AD

    The Role of Zinc in AD

    The Role of Iron in AD

    Therapeutic Targeting of Biometals in AD

    Conclusions

    References

    Introduction

    Alzheimer’s disease (AD) was first described in 1906 by Dr. Alois Alzheimer, and is now the third leading cause of death in industrial countries.¹ The disease is clinically characterized by the progressive loss of memory and other cognitive domains including language, attention, orientation, and problem solving abilities.² The biggest risk factor for AD is age,³ and given the ageing population demographic of our society the incidence of AD is likely to increase in the future. This highlights the urgent need for the development of effective disease-modifying therapies to halt or slow disease progression. Although the aetiology of sporadic AD remains largely unknown, the disease is associated with distinct pathological abnormalities which distinguish the condition from other forms of dementia.⁴ Macroscopically, AD is characterized by progressive cortical atrophy particularly of the frontal, parietal, and temporal lobes.⁵ Microscopically, the disease is characterized by neuronal and synaptic loss, extracellular senile plaques composed of aggregated amyloid beta (Aβ) peptides, and intracellular neurofibrillary tangles (NFTs) composed of hyperphosphorylated forms of the protein tau.⁶,⁷

    According to the amyloid cascade hypothesis which was first suggested in 1992, the Aβ peptide plays a central role in AD pathogenesis.⁸ The amyloid precursor protein (APP) is cleaved through one of two pathways by the metalloproteinases α-secretase, γ-secretase, or the beta-site amyloid precursor protein-cleaving enzyme 1 (BACE 1) which is a β-secretase.⁹ The nonamyloidogenic pathway involves cleavage of the extracellular APP domain by α-secretase followed by cleavage of the intramembranous domain by γ-secretase. By contrast, in the amyloidogenic pathway the initial cleavage of the extracellular APP domain is mediated by BACE 1 resulting in the production of the Aβ peptide.¹⁰ Although Aβ is generated in all brain regions and is ubiquitously expressed throughout the body, not all brain regions are affected.¹¹,¹² This indicates that Aβ expression alone is not sufficient to cause disease, and it is therefore likely that other factors in the affected brain regions are able to either induce Aβ toxicity or to make these brain regions in other ways vulnerable to disease. Consistent with this is the fact that phase three clinical trials performed on drugs that inhibit Aβ production or lower Aβ levels through immunotherapy have failed to alter disease progression.¹³–¹⁶ Alternative therapeutic approaches are therefore urgently needed.

    Biologically functional metal ions (also known as biometals) are tightly regulated in the heathy brain and a breakdown in the homeostasis mechanisms that compartmentalize and regulate these metals can substantially affect brain function.¹⁷ The following sections will examine how copper, zinc, and iron play a role in AD pathogenesis through interactions with disease-associated proteins. Lastly, therapeutic attempts to restore biometal homeostasis in AD will be reviewed.

    The Role of Copper in AD

    Copper is a redox-active metal and exists in either oxidized (Cu²+) or reduced (Cu+) valence states.¹⁸ Many enzymes utilize this change in copper oxidation state, in the presence of oxygen, to catalyze redox chemistry for a wide range of important biochemical reactions. Some important copper-containing enzymes are: Cu/Zn-superoxide dismutase (SOD1), which converts the superoxide free radical into hydrogen peroxide,¹⁹ cytochrome c oxidase (COX) which plays a key role in the mitochondrial electron transport chain,²⁰ and ceruloplasmin (Cp) which functions as a ferroxidase to facilitate iron export from cells.²¹ All of these proteins require copper binding to perform their function, and low copper levels could therefore lead to oxidative stress, mitochondrial dysfunction, and intracellular iron accumulation, all of which are observed in AD.²²–²⁴ The same chemistry which makes copper useful in biology also allows free copper to catalyze the formation of the hydroxyl radical via the Fenton reaction.²⁵ Therefore, because the free form of the metal is potentially damaging, absorption and excretion of copper are tightly regulated in the body. Copper transport across cellular membranes occurs predominantly via the transporters high affinity copper uptake protein 1 (Ctr1) and ATPase copper-transporting alpha and beta polypeptides (ATP7a and ATP7b).²⁶ Neurodegeneration is a feature of both Menkes disease, which is caused by mutations in the gene encoding ATP7a²⁷,²⁸ and Wilson disease, which is caused by mutations in the gene encoding ATP7b.²⁹,³⁰ This demonstrates that copper dysregulation is detrimental to brain health.

    In the AD brain overall copper levels are decreased within affected regions³¹,³² but are enriched within plaques and neurofibrillary tangles.³³ From this a complex picture emerges where abnormal copper distribution in AD leads to copper deficiency within affected brain regions. In the healthy brain, postsynaptic N-methyl-D-aspartic acid (NMDA) neurites release ionic copper into the synaptic cleft, facilitated by ATP7a, at a concentration of around 15 μM.³⁴,³⁵ Within the synaptic cleft copper causes S-nitrosylation of NMDA receptors, which inhibits their activation.³⁶ The sequestration of copper within amyloid plaques has been proposed as a mechanism by which the pool of free copper within the synaptic cleft is depleted leading to increased activation of NMDA receptors.³⁷ In contrast with decreased brain copper levels in AD, the serum and cerebrospinal fluid (CSF) levels of copper are significantly elevated in AD patients when compared to age-matched controls.³⁸,³⁹ This may correlate with increased expression of ceruloplasmin (a major copper-binding protein in serum) in AD, although excess copper is not bound to this carrier protein.⁴⁰ Furthermore, increased serum copper was reported to correlate well with higher levels of serum peroxides in AD patients.⁴¹ Taken together these studies indicate that altered copper distribution in AD brain, serum, and CSF contributes to disease pathogenesis.

    Copper binds to APP in the amino-terminal ectodomain (between residues 142 and 166)⁴² and catalytically reduces copper (II) to copper (I).⁴³ The structure of the APP copper binding domain consist of four ligands (His-147, His-151, Tyr-168, and Met-170),⁴² which show structural homology to other copper chaperones. As a ubiquitously expressed protein, APP may play a role in regulating metal ion homeostasis. This is supported by studies which have shown that chronic copper overload or copper deficiency can both up- and downregulate APP mRNA expression in mouse fibroblasts.⁴⁴,⁴⁵ Furthermore, another study found that a low copper diet in healthy individuals was associated with decreased APP protein expression in platelets.⁴⁶ Conversely, both in vivo and in vitro studies in APP-knockout mice have shown that copper levels are significantly increased in brain and liver tissues as well as in cortical neuron and fibroblast primary cultures derived from these animals.⁴⁷,⁴⁸ One of these studies also found that APP-knockout primary cortical neurons are susceptible to copper-induced toxicity through copper (I) production which caused localized oxidative stress.⁴⁷ Moreover, the overexpression of mutant APP in various transgenic mouse lines decreased copper levels in both in vivo and in vitro contexts.⁴⁹–⁵¹ Additionally, copper promotes an increase in cell surface APP by increasing its exocytosis and reducing its endocytosis, respectively.⁵² Collectively these studies indicate that the interaction between copper and APP may contribute to AD pathogenesis.

    Copper binds to the Aβ peptide in a pH-dependent manner and Aβ1–16 has been shown to be the minimal sequence required for copper binding.⁵³ Between pH 6 and 7, Aβ binds to copper at Asp1, His6, and His13/14; while at pH 8, the binding sites shift to His6, His13, and His14.⁵⁴ At pH 10 or higher, Asp1, Ala2, Glu3, and Phe4 can also form a complex with copper.⁵⁵,⁵⁶ Of particular interest is the fact that Aβ purified from human brain plaques contains fewer histidine residues, which has been explained by copper-mediated oxidation.⁵⁷ Copper modifies Aβ by promoting dityrosine crosslinking of the peptide, which may act as a seed to accelerate Aβ aggregation and induce oligomer formation.⁵⁸,⁵⁹ Aβ toxicity is partially dependant on copper and can be attenuated in cell culture by copper chelation.⁶⁰,⁶¹ The mechanism of copper-Aβ induced cytotoxicity might involve oxidative stress, as the complex catalytically generates hydrogen peroxide.⁶² Furthermore, another study found that the copper-Aβ complex could inhibit COX function thereby impairing mitochondrial energy production.⁶³

    Copper also binds to the tau protein and certain fragments in the four-repeat microtubule-binding domain of tau (residues 256–273, 287–304, and 306–336) were shown to aggregate in the presence of copper in vitro.⁶⁴,⁶⁵ Furthermore, copper binding to tau induces hydrogen peroxide production,⁶⁶ which recapitulates what is observed in AD brains where copper-containing NFTs are a source of oxidative stress.⁶⁷ Chronic copper exposure accelerates tau hyperphosphorylation via abnormal cyclin-dependent kinase 5 (Cdk5) activation, resulting in dissociation of tau from microtubules in an AD mouse model.⁶⁸ However, copper delivery drugs have been shown to decrease glycogen synthase kinase 3 beta (GSK-3β)-dependent tau phosphorylation.⁶⁹ These studies indicate that copper alters tau phosphorylation through diverse mechanisms.

    The Role of Zinc in AD

    Zinc is another transition metal which plays an important role in diverse cellular functions. With the possible exception of pancreatic β islets, the brain contains the highest concentration of zinc within the body.⁷⁰ Within the healthy brain, the majority of zinc is compartmentalized in membrane-bound metalloproteins (particularly metallothioneins MT-I, II, and III).⁷¹ Diverse classes of proteins require bound zinc for normal function, including metalloenzymes (e.g., SOD1), transcription factors, and signaling kinases.⁷²,⁷³ In its free ionic form, zinc in the brain is highly enriched in glutamatergic nerve terminals where it is released (at concentrations of 1–100 μM) upon neuronal activation.⁷⁴ Synaptic zinc has a functional role in signal transmission and acts as an antagonist to NMDA receptors.⁷⁵ Intracellular zinc uptake is facilitated by a number of zinc-importing proteins (ZIPs), particularly ZIPs 1–5 and 7–15.⁷⁶ Zinc uptake can also be mediated by NMDA receptor-dependent voltage-gated L-type Ca²+, Ca²+-permeable AMPA/kainate channels, and Na+/Zn+ exchangers.⁷⁷–⁷⁹ Approximately 50% of zinc uptake requires the AD-linked presenilin protein, however, the mechanism by which presenilin contributes to zinc uptake is unknown.⁸⁰ Intracellular zinc sequestration or export occurs via the zinc transport (ZnT) protein family. ZnT-2, 5, 7, and 8 are expressed at low levels within the brain, while ZnT-3 is found in granule, pyramidal, and interneuron cells of the hippocampus and plays a role in transporting zinc into glutamatergic vesicles.⁸¹ Indeed, a loss of synaptic zinc in ZnT-3-knockout mice causes age-depend cognitive decline,⁸² demonstrating the importance of synaptic zinc in brain function.

    Bulk tissue analyses of postmortem AD brains have yielded inconsistent results. Earlier work showed no difference in brain zinc levels between AD and controls,⁸³,⁸⁴ while latter studies showed a decrease in zinc levels in several AD brain regions including the neocortex, superior frontal and parietal gyri, medial temporal gyrus and thalamus, and the hippocampus.⁸⁵,⁸⁶ Conflicting reports have, however, also shown elevated zinc levels in multiple AD brain regions including the amygdala, hippocampus, cerebellum, olfactory areas, and superior temporal gyrus.³¹,⁸⁷ These discrepancies could be a result of the examination of different brain regions and/or the utilization of different sample preparations (e.g., tissue fixation affects zinc measurement).⁸⁸ Moreover, bulk tissue analyses are unlikely to reveal changes in zinc compartmentalization. Indeed, studies have shown that in AD there is a redistribution of zinc into extracellular plaques and surrounding neuropil.³³,⁸⁹ While the cause of zinc dysregulation in AD remains unknown, changes to several proteins involved in zinc homeostasis including MT-I, MT-II, MT-III, ZnT-1, ZnT-3, ZnT4, and ZnT-6 in AD⁹⁰–⁹³ are likely to contribute to the aberrant compartmentalization of zinc within the brain. Interestingly, estrogen can also modulate levels of ZnT-3 which is of particular significance given that sex is another major risk factor for AD.⁹⁴ The serum and CSF levels of zinc are decreased in AD patients when compared to age-matched controls,³⁸,⁹⁵ which could in part be explained by nutritional deficiency associated with advanced age.¹⁷

    Zinc binds to APP in a conserved region of amino acids between residues 170 and 188 and this domain consists of two key cysteine ligands at position 186 and 187, which are crucial for binding as well as other potential ligands (e.g., Cys174, Met170, Asp177, and Glu184).⁹⁶,⁹⁷ Zinc interferes with APP processing by altering the function of key secretases involved in APP cleavage. A disintegrin and metalloproteinase domain-containing protein 10 (ADAM 10), the α-secretase involved in the nonamyloidogenic processing of APP, is a zinc-dependent enzyme and thus zinc increases APP proteolysis.⁹⁸ Zinc also increases presenilin 1 expression which facilities cellular zinc uptake,⁸⁰ although the activity of the γ-secretase complex is inhibited by zinc.⁹⁹ Furthermore, the binding of zinc to Aβ within the APP protein sequence can mask the proteolytic cleavage site,¹⁰⁰ thus inhibiting degradation of Aβ by matrix metalloproteases.¹⁰¹

    Zinc binds to Aβ between residues 6 and 28 with up to three zinc ions bound to histidines 6, 13, and 14.⁹⁶,⁹⁷,¹⁰⁰ Zinc binding rapidly induces the aggregation of Aβ into insoluble precipitates, which typify AD pathology.¹⁰² Zinc-induced plaque formation in AD is also supported by the anatomical distribution of plaque and zinc in the brain. Although Aβ is ubiquitously expressed, plaque formation only occurs in neocortical regions of AD-affected brains which closely align with the expression of ZnT3.¹⁰³ APP transgenic mice crossed with ZnT3-knockout mice exhibited decreased Aβ burden,¹⁰³ which demonstrates the contribution of endogenous zinc to amyloid burden in AD. Furthermore, zinc sequestration into amyloid deposits induces loss of functional zinc in the synapse.¹⁰² Synaptic zinc deficiency is further exacerbated in AD by concomitant loss of ZnT3 expression.³³ Therefore, by two mechanisms labile zinc is made deficient in the brain neuropil in AD.

    Zinc can directly bind to tau monomers with moderate affinity, altering its confirmation, and can induce both the fibrillization and the aggregation of the protein.¹⁰⁴,¹⁰⁵ Zinc also modulates the translation and phosphorylation of tau by affecting the activities of GSK-3β, protein kinase B, ERK1/2, and c-Jun N-terminal kinase.¹⁰⁶,¹⁰⁷ Furthermore, zinc is elevated in neurons with neurofibrillary tangle pathology.⁸⁹

    The Role of Iron in AD

    Iron is a transition metal which can exist in oxidation states from −2 to +8, but in biological systems only ferrous (Fe²+) and ferric (Fe³+) states exist. The cycling between Fe²+ and Fe³+ is utilized in biology for various electron transfer (redox) reactions essential to life.¹⁰⁸,¹⁰⁹ Furthermore, iron is required for other essential biological processes including: the transport of oxygen (where iron is bound to haemoglobin),¹¹⁰ regulation of protein expression,¹¹¹,¹¹² cell growth¹¹³ and cell differentiation,¹¹⁴ as well as brain development,¹¹⁵ neurotransmitter synthesis,¹¹⁶ and myelin production.¹¹⁷ Although essential to biological processes, when in excess, iron is toxic because it can react with oxygen through the Fenton reaction to generate superoxide anions and hydroxyl radicals,¹¹⁸ which are a source of oxidative stress.¹¹⁹ For these reasons iron levels are tightly regulated within the body and disruption of these homeostatic processes can cause either iron-deficient anemia or iron overload disorders.¹²⁰ Iron homeostasis is maintained via the coordinated action of several proteins including the iron carrier protein transferrin, the transferrin receptor for iron import, the ferroportin channel for iron export, the iron storage protein ferritin, ferrireductases which reduce iron to its ferrous state, and ferroxidases which oxidize iron to its ferric state. Circulation iron, once oxidized to its ferric state using the ferroxidase ceruloplasmin, cannot cross the blood brain barrier (BBB), but requires an iron complexed to transferrin to bind the transferrin receptor on the lumen side of endothelial cells.¹²¹,¹²² The transferrin complex then enters the cell via endocytosis where iron assimilates with a labile iron pool within the cytosol and is available for incorporation into iron-binding proteins, such as ferritin.¹²³ This process is highly regulated by the abundance or the deficiency in both transferrin (with or without iron incorporated) and its receptor.¹²⁴,¹²⁵

    In the AD brain iron levels are elevated¹²⁶,¹²⁷ and especially enriched within neurofibrillary tangles²⁴ and amyloid plaques,¹²⁸ the latter of which has an iron concentration three times higher than that which is found in the surrounding normal neuropil.³³ Iron accumulation occurs in the AD cortex, but not the cerebellum, which is consistent with the anatomical profile of neurodegeneration in AD.¹²⁹,¹³⁰ The iron storage protein ferritin binds most iron within the brain,¹²⁷ and this protein increases with age and in AD.¹³¹ Neuronal iron deposition causes oxidative stress which likely contributes to elevated oxidative stress observed in the AD brain.¹³² Furthermore, iron-induced oxidative stress has been shown to initiate several apoptotic pathways in neurons and damage lipids and proteins (including the NMDA receptor) resulting in synaptic dysfunction and neuronal cell death.¹³³–¹³⁵ A number of iron-associated proteins have an altered expression profile in AD, which could partly explain the cause of iron accumulation in AD. Ferritin has been reported to be elevated in AD and colocalizes with amyloid plaques.¹²⁶ Transferrin, which is normally expressed solely by oligodendrocytes, is also expressed in astrocytes in the AD brain¹²⁶ and was found to be increased in frontal cortex of AD.¹³⁶ Lower ceruloplasmin expression was found in AD brains,¹³⁷ as well as its ferroxidase activity in plasma.¹³⁸ Taken together, it is likely that the iron accumulation observed in AD is a result of multiple failures in its regulatory proteins. Reports of iron levels in serum and CSF have yet to show a consistent change in AD.³⁸,¹³⁹ While serum iron levels may be unchanged, a recent study has shown that CSF ferritin levels can predict AD outcomes.¹⁴⁰

    Iron binds to APP via the iron-responsive element type II located within the 5’ untranslated region of its mRNA sequence.¹⁴¹ Under conditions of high iron levels, such as are seen in the AD brain, restricted translation of APP by iron-responsive proteins is diminished, leading to increased translation of the transcript.¹¹¹ Conversely, the same study also found that iron chelation decreased translation of the transcript. Iron may also alter APP cleavage by modulating furin, a proconvertase involved in the regulation of α-secretase-dependent processing.¹⁴² Since ferrous iron has a low affinity for transferrin, it requires oxidation by a ferroxidase before it can be removed from the cell.¹⁴³ Ceruloplasmin is the classic ferroxidase¹⁴⁴,¹⁴⁵; however, this protein is not expressed in neurons.¹⁴⁶ In one recent study APP was proposed as the analogous neuronal ferroxidase.¹²⁹ This study found that APP-knockout mice exhibit iron accumulation in brain and peripheral tissues, and loss of APP ferroxidases activity in the AD brain is coincident with iron retention in the tissue.

    Iron in both Fe²+ and Fe³+ states binds to the Aβ peptide at residues Asp1, Glu3, His6, His13, and His14.¹⁴⁷,¹⁴⁸ Iron promotes the aggregation of Aβ in vitro,¹⁴⁹ which can be prevented by iron chelation.¹⁵⁰ The iron-aggregated Aβ is also toxic to cultured cells,¹⁵¹,¹⁵² which has been suggested to be mediated by reactive oxygen species,¹⁵³ by Fenton chemistry,¹⁵⁴ or by the activation of the Bcl-2-related apoptosis pathway.¹⁵⁵ Taken together, these studies suggest a role for iron-mediated Aβ toxicity in AD.

    Iron also binds to the tau protein, and the binding of Fe³+, but not Fe²+, causes aggregation of hyperphosphorylated tau that can be reversed by reducing Fe³+ to Fe²+ or by iron chelation.¹⁵⁶,¹⁵⁷ Within the AD brain, iron enrichment colocalizes with NFTs and is a source of reactive oxygen species.²⁴,⁶⁷ Iron also affects the phosphorylation status of tau: Fe³+ decreases tau phosphorylation¹⁵⁸ while Fe²+ increases tau phosphorylation.¹⁵⁹,¹⁶⁰ Furthermore, a recent study has shown that the iron-export capability of APP requires the binding of tau to APP.¹⁶¹,¹⁶² In tau-knockout primary cortical neurons, APP was inappropriately trafficked and not presented on the extracellular surface where it acts as a ferroxidase.¹⁶¹ Total tau levels are decreased in the AD cortex,¹⁶³,¹⁶⁴ and loss of tau expression causes iron- and age-dependent cognitive loss and cortical atrophy in mice.¹⁶¹ These studies demonstrate the apparent interconnection between iron overload, tau, and APP in the pathogenesis of AD.

    Therapeutic Targeting of Biometals in AD

    Given that targeting Aβ in multiple ways has so far failed to confer clinical benefits,¹³–¹⁶ there is a need to target other pathways in AD. Targeting biometals by restoring metal ion homeostasis represents an alternative potential therapeutic avenue. To achieve this, sophisticated compounds are needed which can correct metal miscompartmentalization in AD by redistributing metal ions from extracellular plaques (preventing Aβ aggregation) into metal ion-deficient neurons (to restore normal function). Several potential metal-based AD therapeutics are discussed later.

    5-Chloro-7-iodo-quinolin-8-ol (clioquinol) is a derivative of 8-hydroxyquinoline that was widely used as an antiparasitic agent before it was withdrawn from clinical use owing to a speculated severe side effect, subacute myelo-optic-neuropathy (SMON). This side effect was only observed in Japan,¹⁶⁵ and the association between SMON and clioquinol has since been questioned.¹⁶⁶ Although it was initially considered a moderate chelator of iron, copper, and zinc,¹⁶⁷ clioquinol has more recently been characterized as a copper/zinc ionophore, which functions to redistribute these metals into cells.¹⁶⁸,¹⁶⁹ In addition, clioquinol is believed to confer neuroprotection by iron chelation, as iron binds to clioquinol,¹⁷⁰ and a number of the reported beneficial effects of clioquinol are iron dependent.¹⁵⁴,¹⁷¹ Preclinical studies have shown promising outcomes, including deceased Aβ brain burden,¹⁶⁶ inhibition of Aβ oligomer formation,¹⁷²,¹⁷³ and rescue of memory impairment in clioquinol treated animals.¹⁶⁶ Furthermore, a phase two clinical trial¹⁷⁴ and a case study¹⁷⁵ reported improvement in cognitive outcomes for patients with AD. However, complications with large-scale manufacturing of the compound have hindered further exploration for its use in AD.

    A second-generation 8-hydroxyquinoline derivative, PBT2, has shown even greater therapeutic efficiency in an AD mouse model¹⁷⁶ as well as in a phase two clinical trial.¹⁷⁷,¹⁷⁸ However, in another more recent phase two clinical trial, PBT2 did not show an improved Pittsburgh compound B-PET scan when compared to placebo patients.¹⁷⁹ While patients on PBT2 had a lower Pittsburgh compound B-PET signal, the result was confounded by an inexplicable reduction in the placebo group. The proposed mechanism for neuroprotection of this drug is by acting as a copper-zinc ionophore, redistributing copper and zinc inside the cell, which induces inhibitory phosphorylation of the α- and β-isoforms of GSK-3 and subsequently lowers Aβ levels.¹⁸⁰ More comprehensive clinical studies are needed to further investigate the utility of this compound as an AD therapeutic.

    Given that the mechanisms of action for both clioquinol and PBT2 probably involve their copper ionophore activity, copper-containing bis(thiosemicarbazone) compounds have been explored for their potential to treat AD. One compound, CuII(gtsm), delivers copper to neurons and has been shown to lower Aβ levels, GSK3β activity, and tau phosphorylation levels in cell culture and AD model mice, which was accompanied by improved cognition in these mice.⁶⁹,¹⁸¹ Another compound, CuII(atsm), delivers copper selectively to cells with an impaired electron transport chain¹⁸² and has not been shown to be beneficial in the APP/PS1 mouse model of AD.⁶⁹ Further studies are needed to examine the utility of these copper-containing bis(thiosemicarbazone) ligands for the treatment of AD.

    Given the role for iron in regulating APP translation,¹¹¹ decreasing iron overload in AD via iron chelators has shown promise in both preclinical and clinical trials. The iron chelators epigallocatechin-3-gallate (EGCG) and M-30 have been shown to reduce APP expression in cultured cells.¹⁸³,¹⁸⁴ Furthermore, the iron chelator deferoxamine inhibits amyloidogenic APP processing in cultured cells and in AD model mice, which also attenuated Aβ burden within the brain and reversed spatial memory impairment.¹⁸⁵,¹⁸⁶ Intramuscular injection of deferoxamine was tested in a single-blind clinical trial of 48 AD patients over a 24-month period and deferoxamine treatment reduced the rate of cognitive decline by half.¹⁸⁷ Despite this encouraging trial outcome obtained in 1991, further clinical developments of compounds that target iron in AD have not occurred.

    Conclusions

    Through interactions with disease-associated proteins, including APP, Aβ, and tau, the biometals copper, zinc, and iron appear to play an important role in AD pathogenesis (Table 1.1). Altered biometal homeostasis in AD has opened up new opportunities for the development of disease-modifying therapeutics. Preclinical and clinical data indicate that targeting biometals by restoring metal ion homeostasis remains a promising prospect for the treatment of AD.

    Table 1.1

    Biometal Interactions With Disease-Associated Proteins in Alzheimer’s Disease

    References

    1. Mattson MP. Pathways towards and away from Alzheimer’s disease. Nature. 2004;430(7000):631–639.

    2. Jalbert JJ, Daiello LA, Lapane KL. Dementia of the Alzheimer type. Epidemiol Rev. 2008;30:15–34.

    3. Ferri CP, Prince M, Brayne C, Brodaty H, Fratiglioni L, Ganguli M, Hall K, Hasegawa K, Hendrie H, Huang Y, Jorm A, Mathers C, Menezes PR, Rimmer E, Scazufca M. Global prevalence of dementia: a Delphi consensus study. Lancet. 2005;366(9503):2112–2117.

    4. Selkoe DJ. Alzheimer’s disease: genes, proteins, and therapy. Physiol Rev. 2001;81(2):741–766.

    5. McDonald CR, McEvoy LK, Gharapetian L, Fennema-Notestine C, Hagler Jr DJ, Holland D, Koyama A, Brewer JB, Dale AM. Regional rates of neocortical atrophy from normal aging to early Alzheimer disease. Neurology. 2009;73(6):457–465.

    6. Grundke-Iqbal I, Iqbal K, Tung YC, Quinlan M, Wisniewski HM, Binder LI. Abnormal phosphorylation of the microtubule-associated protein tau (tau) in Alzheimer cytoskeletal pathology. Proc Natl Acad Sci USA. 1986;83(13):4913–4917.

    7. Masters CL, Simms G, Weinman NA, Multhaup G, McDonald BL, Beyreuther K. Amyloid plaque core protein in Alzheimer disease and Down syndrome. Proc Natl Acad Sci USA. 1985;82(12):4245–4249.

    8. Hardy JA, Higgins GA. Alzheimer’s disease: the amyloid cascade hypothesis. Science. 1992;256(5054):184–185.

    9. Tanzi RE, Gusella JF, Watkins PC, Bruns GA, St George-Hyslop P, Van Keuren ML, Patterson D, Pagan S, Kurnit DM, Neve RL. Amyloid beta protein gene: cDNA, mRNA distribution, and genetic linkage near the Alzheimer locus. Science. 1987;235(4791):880–884.

    10. Vassar R, Bennett BD, Babu-Khan S, Kahn S, Mendiaz EA, Denis P, Teplow DB, Ross S, Amarante P, Loeloff R, Luo Y, Fisher S, Fuller J, Edenson S, Lile J, Jarosinski MA, Biere AL, Curran E, Burgess T, Louis JC, Collins F, Treanor J, Rogers G, Citron M. Beta-secretase cleavage of Alzheimer’s amyloid precursor protein by the transmembrane aspartic protease BACE. Science. 1999;286(5440):735–741.

    11. Spillantini MG, Hunt SP, Ulrich J, Goedert M. Expression and cellular localization of amyloid beta-protein precursor transcripts in normal human brain and in Alzheimer’s disease. Brain Res Mol Brain Res. 1989;6(2-3):143–150.

    12. Arai H, Lee VM, Messinger ML, Greenberg BD, Lowery DE, Trojanowski JQ. Expression patterns of beta-amyloid precursor protein (beta-APP) in neural and nonneural human tissues from Alzheimer’s disease and control subjects. Ann Neurol. 1991;30(5):686–693.

    13. Green RC, Schneider LS, Amato DA, Beelen AP, Wilcock G, Swabb EA, Zavitz KH. Effect of tarenflurbil on cognitive decline and activities of daily living in patients with mild Alzheimer disease: a randomized controlled trial. JAMA. 2009;302(23):2557–2564.

    14. Doody RS, Raman R, Farlow M, Iwatsubo T, Vellas B, Joffe S, Kieburtz K, He F, Sun X, Thomas RG, Aisen PS, Siemers E, Sethuraman G, Mohs R. A phase 3 trial of semagacestat for treatment of Alzheimer’s disease. N Engl J Med. 2013;369(4):341–350.

    15. Doody RS, Thomas RG, Farlow M, Iwatsubo T, Vellas B, Joffe S, Kieburtz K, Raman R, Sun X, Aisen PS, Siemers E, Liu-Seifert H, Mohs R. Phase 3 trials of solanezumab for mild-to-moderate Alzheimer’s disease. N Engl J Med. 2014;370(4):311–321.

    16. Salloway S, Sperling R, Fox NC, Blennow K, Klunk W, Raskind M, Sabbagh M, Honig LS, Porsteinsson AP, Ferris S, Reichert M, Ketter N, Nejadnik B, Guenzler V, Miloslavsky M, Wang D, Lu Y, Lull J, Tudor IC, Liu E, Grundman M, Yuen E, Black R, Brashear HR. Two phase 3 trials of bapineuzumab in mild-to-moderate Alzheimer’s disease. N Engl J Med. 2014;370(4):322–333.

    17. Duce JA, Bush AI. Biological metals and Alzheimer’s disease: implications for therapeutics and diagnostics. Prog Neurobiol. 2010;92(1):1–18.

    18. Ridge PG, Zhang Y, Gladyshev VN. Comparative genomic analyses of copper transporters and cuproproteomes reveal evolutionary dynamics of copper utilization and its link to oxygen. PLoS One. 2008;3(1):e1378.

    19. Robinson NJ, Winge DR. Copper metallochaperones. Annu Rev Biochem. 2010;79:537–562.

    20. Horn D, Barrientos A. Mitochondrial copper metabolism and delivery to cytochrome c oxidase. IUBMB Life. 2008;60(7):421–429.

    21. Wong BX, Ayton S, Lam LQ, Lei P, Adlard PA, Bush AI, Duce JA. A comparison of ceruloplasmin to biological polyanions in promoting the oxidation of Fe(2+) under physiologically relevant conditions. Biochim Biophys Acta. 2014;1840(12):3299–3310.

    22. Perry G, Cash AD, Smith MA. Alzheimer disease and oxidative stress. J Biomed Biotechnol. 2002;2(3):120–123.

    23. Moreira PI, Carvalho C, Zhu X, Smith MA, Perry G. Mitochondrial dysfunction is a trigger of Alzheimer’s disease pathophysiology. Biochim Biophys Acta. 2010;1802(1):2–10.

    24. Smith MA, Harris PL, Sayre LM, Perry G. Iron accumulation in Alzheimer disease is a source of redox-generated free radicals. Proc Natl Acad Sci USA. 1997;94(18):9866–9868.

    25. Madsen E, Gitlin JD. Copper and iron disorders of the brain. Annu Rev Neurosci. 2007;30:317–337.

    26. Ip V, Liu JJ, Mercer JF, McKeage MJ. Differential expression of ATP7A, ATP7B and CTR1 in adult rat dorsal root ganglion tissue. Mol Pain. 2010;6:53.

    27. Vulpe C, Levinson B, Whitney S, Packman S, Gitschier J. Isolation of a candidate gene for Menkes disease and evidence that it encodes a copper-transporting ATPase. Nat Genet. 1993;3(1):7–13.

    28. Kodama H, Murata Y, Kobayashi M. Clinical manifestations and treatment of Menkes disease and its variants. Pediatr Int. 1999;41(4):423–429.

    29. Bull PC, Thomas GR, Rommens JM, Forbes JR, Cox DW. The Wilson disease gene is a putative copper transporting P-type ATPase similar to the Menkes gene. Nat Genet. 1993;5(4):327–337.

    30. Merle U, Stremmel W, Encke J. Perspectives for gene therapy of Wilson disease. Curr Gene Ther. 2007;7(3):217–220.

    31. Deibel MA, Ehmann WD, Markesbery WR. Copper, iron, and zinc imbalances in severely degenerated brain regions in Alzheimer’s disease: possible relation to oxidative stress. J Neurol Sci. 1996;143(1-2):137–142.

    32. Magaki S, Raghavan R, Mueller C, Oberg KC, Vinters HV, Kirsch WM. Iron, copper, and iron regulatory protein 2 in Alzheimer’s disease and related dementias. Neurosci Lett. 2007;418(1):72–76.

    33. Lovell MA, Robertson JD, Teesdale WJ, Campbell JL, Markesbery WR. Copper, iron and zinc in Alzheimer’s disease senile plaques. J Neurol Sci. 1998;158(1):47–52.

    34. Hopt A, Korte S, Fink H, Panne U, Niessner R, Jahn R, Kretzschmar H, Herms J. Methods for studying synaptosomal copper release. J Neurosci Methods. 2003;128(1-2):159–172.

    35. Hartter DE, Barnea A. Evidence for release of copper in the brain: depolarization-induced release of newly taken-up 67copper. Synapse. 1988;2(4):412–415.

    36. Schlief ML, Craig AM, Gitlin JD. NMDA receptor activation mediates copper homeostasis in hippocampal neurons. J Neurosci. 2005;25(1):239–246.

    37. Stys PK, You H, Zamponi GW. Copper-dependent regulation of NMDA receptors by cellular prion protein: implications for neurodegenerative disorders. J Physiol. 2012;590(6):1357–1368.

    38. Basun H, Forssell LG, Wetterberg L, Winblad B. Metals and trace elements in plasma and cerebrospinal fluid in normal aging and Alzheimer’s disease. J Neural Transm Park Dis Dement Sect. 1991;3(4):231–258.

    39. Squitti R, Lupoi D, Pasqualetti P, Dal Forno G, Vernieri F, Chiovenda P, Rossi L, Cortesi M, Cassetta E, Rossini PM. Elevation of serum copper levels in Alzheimer’s disease. Neurology. 2002;59(8):1153–1161.

    40. Squitti R, Pasqualetti P, Dal Forno G, Moffa F, Cassetta E, Lupoi D, Vernieri F, Rossi L, Baldassini M, Rossini PM. Excess of serum copper not related to ceruloplasmin in Alzheimer disease. Neurology. 2005;64(6):1040–1046.

    41. Squitti R, Rossini PM, Cassetta E, Moffa F, Pasqualetti P, Cortesi M, Colloca A, Rossi L, Finazzi-Agro A. d-penicillamine reduces serum oxidative stress in Alzheimer’s disease patients. Eur J Clin Invest. 2002;32(1):51–59.

    42. Barnham KJ, McKinstry WJ, Multhaup G, Galatis D, Morton CJ, Curtain CC, Williamson NA, White AR, Hinds MG, Norton RS, Beyreuther K, Masters CL, Parker MW, Cappai R. Structure of the Alzheimer’s disease amyloid precursor protein copper binding domain. A regulator of neuronal copper homeostasis. J Biol Chem. 2003;278(19):17401–17407.

    43. Multhaup G, Schlicksupp A, Hesse L, Beher D, Ruppert T, Masters CL, Beyreuther K. The amyloid precursor protein of Alzheimer’s disease in the reduction of copper(II) to copper(I). Science. 1996;271(5254):1406–1409.

    44. Armendariz AD, Gonzalez M, Loguinov AV, Vulpe CD. Gene expression profiling in chronic copper overload reveals upregulation of Prnp and App. Physiol Genomics. 2004;20(1):45–54.

    45. Bellingham SA, Lahiri DK, Maloney B, La Fontaine S, Multhaup G, Camakaris J. Copper depletion down-regulates expression of the Alzheimer’s disease amyloid-beta precursor protein gene. J Biol Chem. 2004;279(19):20378–20386.

    46. Davis CD, Milne DB, Nielsen FH. Changes in dietary zinc and copper affect zinc-status indicators of postmenopausal women, notably, extracellular superoxide dismutase and amyloid precursor proteins. Am J Clin Nutr. 2000;71(3):781–788.

    47. White AR, Multhaup G, Maher F, Bellingham S, Camakaris J, Zheng H, Bush AI, Beyreuther K, Masters CL, Cappai R. The Alzheimer’s disease amyloid precursor protein modulates copper-induced toxicity and oxidative stress in primary neuronal cultures. J Neurosci. 1999;19(21):9170–9179.

    48. Bellingham SA, Ciccotosto GD, Needham BE, Fodero LR, White AR, Masters CL, Cappai R, Camakaris J. Gene knockout of amyloid precursor protein and amyloid precursor-like protein-2 increases cellular copper levels in primary mouse cortical neurons and embryonic fibroblasts. J Neurochem. 2004;91(2):423–428.

    49. Bayer TA, Schafer S, Simons A, Kemmling A, Kamer T, Tepest R, Eckert A, Schussel K, Eikenberg O, Sturchler-Pierrat C, Abramowski D, Staufenbiel M, Multhaup G. Dietary Cu stabilizes brain superoxide dismutase 1 activity and reduces amyloid Abeta production in APP23 transgenic mice. Proc Natl Acad Sci USA. 2003;100(24):14187–14192.

    50. Maynard CJ, Cappai R, Volitakis I, Cherny RA, White AR, Beyreuther K, Masters CL, Bush AI, Li QX. Overexpression of Alzheimer’s disease amyloid-beta opposes the age-dependent elevations of brain copper and iron. J Biol Chem. 2002;277(47):44670–44676.

    51. Phinney AL, Drisaldi B, Schmidt SD, Lugowski S, Coronado V, Liang Y, Horne P, Yang J, Sekoulidis J, Coomaraswamy J, Chishti MA, Cox DW, Mathews PM, Nixon RA, Carlson GA, St George-Hyslop P, Westaway D. In vivo reduction of amyloid-beta by a mutant copper transporter. Proc Natl Acad Sci USA. 2003;100(24):14193–14198.

    52. Acevedo KM, Hung YH, Dalziel AH, Li QX, Laughton K, Wikhe K, Rembach A, Roberts B, Masters CL, Bush AI, Camakaris J. Copper promotes the trafficking of the amyloid precursor protein. J Biol Chem. 2011;286(10):8252–8262.

    53. Minicozzi V, Stellato F, Comai M, Dalla Serra M, Potrich C, Meyer-Klaucke W, Morante S. Identifying the minimal copper- and zinc-binding site sequence in amyloid-beta peptides. J Biol Chem. 2008;283(16):10784–10792.

    54. Drew SC, Noble CJ, Masters CL, Hanson GR, Barnham KJ. Pleomorphic copper coordination by Alzheimer’s disease amyloid-beta peptide. J Am Chem Soc. 2009;131(3):1195–1207.

    55. Dorlet P, Gambarelli S, Faller P, Hureau C, Pulse EPR. spectroscopy reveals the coordination sphere of copper(II) ions in the 1-16 amyloid-beta peptide: a key role of the first two N-terminus residues. Angew Chem Int Ed Engl. 2009;48(49):9273–9276.

    56. Karr JW, Szalai VA. Role of aspartate-1 in Cu(II) binding to the amyloid-beta peptide of Alzheimer’s disease. J Am Chem Soc. 2007;129(13):3796–3797.

    57. Atwood CS, Huang X, Khatri A, Scarpa RC, Kim YS, Moir RD, Tanzi RE, Roher AE, Bush AI. Copper catalyzed oxidation of Alzheimer Abeta. Cell Mol Biol. 2000;46(4):777–783.

    58. Atwood CS, Perry G, Zeng H, Kato Y, Jones WD, Ling KQ, Huang X, Moir RD, Wang D, Sayre LM, Smith MA, Chen SG, Bush AI. Copper mediates dityrosine cross-linking of Alzheimer’s amyloid-beta. Biochemistry. 2004;43(2):560–568.

    59. Ali FE, Separovic F, Barrow CJ, Cherny RA, Fraser F, Bush AI, Masters CL, Barnham KJ. Methionine regulates copper/hydrogen peroxide oxidation products of Abeta. J Pept Sci. 2005;11(6):353–360.

    60. You H, Tsutsui S, Hameed S, Kannanayakal TJ, Chen L, Xia P, Engbers JD, Lipton SA, Stys PK, Zamponi GW. Abeta neurotoxicity depends on interactions between copper ions, prion protein, and N-methyl-D-aspartate receptors. Proc Natl Acad Sci USA. 2012;109(5):1737–1742.

    61. Wu WH, Lei P, Liu Q, Hu J, Gunn AP, Chen MS, Rui YF, Su XY, Xie ZP, Zhao YF, Bush AI, Li YM. Sequestration of copper from beta-amyloid promotes selective lysis by cyclen-hybrid cleavage agents. J Biol Chem. 2008;283(46):31657–31664.

    62. Huang X, Atwood CS, Hartshorn MA, Multhaup G, Goldstein LE, Scarpa RC, Cuajungco MP, Gray DN, Lim J, Moir RD, Tanzi RE, Bush AI. The A beta peptide of Alzheimer’s disease directly produces hydrogen peroxide through metal ion reduction. Biochemistry. 1999;38(24):7609–7616.

    63. Crouch PJ, Blake R, Duce JA, Ciccotosto GD, Li QX, Barnham KJ, Curtain CC, Cherny RA, Cappai R, Dyrks T, Masters CL, Trounce IA. Copper-dependent inhibition of human cytochrome c oxidase by a dimeric conformer of amyloid-beta1-42. J Neurosci. 2005;25(3):672–679.

    64. Ma QF, Li YM, Du JT, Kanazawa K, Nemoto T, Nakanishi H, Zhao YF. Binding of copper (II) ion to an Alzheimer’s tau peptide as revealed by MALDI-TOF MS, CD, and NMR. Biopolymers. 2005;79(2):74–85.

    65. Zhou LX, Du JT, Zeng ZY, Wu WH, Zhao YF, Kanazawa K, Ishizuka Y, Nemoto T, Nakanishi H, Li YM. Copper (II) modulates in vitro aggregation of a tau peptide. Peptides. 2007;28(11):2229–2234.

    66. Su XY, Wu WH, Huang ZP, Hu J, Lei P, Yu CH, Zhao YF, Li YM. Hydrogen peroxide can be generated by tau in the presence of Cu(II). Biochem Biophys Res Commun. 2007;358(2):661–665.

    67. Sayre LM, Perry G, Harris PL, Liu Y, Schubert KA, Smith MA. In situ oxidative catalysis by neurofibrillary tangles and senile plaques in Alzheimer’s disease: a central role for bound transition metals. J Neurochem. 2000;74(1):270–279.

    68. Kitazawa M, Cheng D, Laferla FM. Chronic copper exposure exacerbates both amyloid and tau pathology and selectively dysregulates cdk5 in a mouse model of AD. J Neurochem. 2009;108(6):1550–1560.

    69. Crouch PJ, Hung LW, Adlard PA, Cortes M, Lal V, Filiz G, Perez KA, Nurjono M, Caragounis A, Du T, Laughton K, Volitakis I, Bush AI, Li QX, Masters CL, Cappai R, Cherny RA, Donnelly PS, White AR, Barnham KJ. Increasing Cu bioavailability inhibits Abeta oligomers and tau phosphorylation. Proc Natl Acad Sci USA. 2009;106(2):381–386.

    70. Frederickson CJ. Neurobiology of zinc and zinc-containing neurons. Int Rev Neurobiol. 1989;31:145–238.

    71. Ebadi M, Elsayed MA, Aly MH. The importance of zinc and metallothionein in brain. Biol Signals. 1994;3(3):123–126.

    72. Vallee BL, Coleman JE, Auld DS. Zinc fingers, zinc clusters, and zinc twists in DNA-binding protein domains. Proc Natl Acad Sci USA. 1991;88(3):999–1003.

    73. Szallasi Z, Bogi K, Gohari S, Biro T, Acs P, Blumberg PM. Non-equivalent roles for the first and second zinc fingers of protein kinase Cdelta. Effect of their mutation on phorbol ester-induced translocation in NIH 3T3 cells. J Biol Chem. 1996;271(31):18299–18301.

    74. Sensi SL, Paoletti P, Bush AI, Sekler I. Zinc in the physiology and pathology of the CNS. Nat Rev Neurosci. 2009;10(11):780–791.

    75. Paoletti P, Ascher P, Neyton J. High-affinity zinc inhibition of NMDA NR1-NR2A receptors. J Neurosci. 1997;17(15):5711–5725.

    76. Lovell MA. A potential role for alterations of zinc and zinc transport proteins in the progression of Alzheimer’s disease. J Alzheimers Dis. 2009;16(3):471–483.

    77. Koh JY,

    Enjoying the preview?
    Page 1 of 1