<<

PHOSPHORYLATION AND SEQUENCE DEPENDENCY

OF NEUROFILAMENT PROTEIN OXIDATIVE

MODIFICATION IN ALZHEIMER DISEASE

By

QUAN LIU

Submitted in partial fulfillment of the requirements

For the degree of Doctor of Philosophy

Dissertation Adviser: Dr. George Perry

Department of Pathology

CASE WESTERN RESERVE UNIVERSITY

January, 2005 CASE WESTERN RESERVE UNIVERSITY

SCHOOL OF GRADUATE STUDIES

We hereby approve the dissertation of

______

candidate for the Ph.D. degree *.

(signed)______(chair of the committee)

______

______

______

______

______

(date) ______

*We also certify that written approval has been obtained for any proprietary material contained therein.

I grant to Case Western Reserve University the right to use this work, irrespective of any copyright, for the University’s own purpose without cost to the University or to its students, agents and employees. I further agree that the

University may reproduce and provide single copies of the work, in any format other than in or from microforms, to the public for cost of reproduction.

Quan Liu

(sign)

iii

DEDICATION

To my parents, my lovely fiancée, and family members

Be proud of me for having double “D”s and partially it is yours.

iv TABLE OF CONTENTS

Title Page i

Signature Sheet ii

Copyright wavier. iii

Dedication. iv

Table of Contents. 1

List of Tables 5

List of Figures 6

Acknowledgements 9

List of Abbreviations 11

Abstract 15

Chapter 1 General Introduction 16

1.1 Introduction of Alzheimer Disease 17

1.1.1 Clinical 17

1.1.2 Pathology 17

1.1.3 Etiology. 18

1.1.4 Pathogenesis and Hypothesis models. 19

1.1.5 Research Relevance: Oxidative Stress and AD Pathogenesis. 24

1.2 Introduction of Neurofilament Proteins 27

1.2.1 General Information 27

1.2.2 History 28

1.2.3 Structure 28

1.2.4 Expression. 30

1 1.2.5 Assembly 32

1.2.6 Transport 34

1.2.7 Posttranslational Modifications 35

1.2.8 Degradation. 37

1.2.9 Functions. 38

1.2.10 Animal Models.. 38

1.2.11 Neurofilaments and Neurodegenerative Diseases. 45

1.2.12 Summary. 52

1.3 Introduction of Tau Protein 54

1.3.1 General Information 54

1.3.2 History. 54

1.3.3 Tau gene and Tau Expression. 55

1.3.4 Tau Protein Structure 56

1.3.5 Posttranslational Modifications of Tau. 56

1.3.6 Tau Degradation 58

1.3.7 Tau Functions. 59

1.3.8 Tau Polymerization. 60

1.3.9 Animal Models 61

1.3.10 Tau and Neurodegenerative Diseases (Tauopathies) 64

1.3.11 Summary. 69

1.4 Introduction of Oxidative Modification. 70

1.4.1 Free radical Theory of Aging 70

1.4.2 Free Radicals. 70

2 1.4.3 The Primary Sources of Free Radicals 71

1.4.4 Physiological Damages Caused by Free Radicals 73

1.4.5 Antioxidant Systems. 73

1.4.6 Oxidative Stress in The Nervous System 76

1.4.7 HNE and HNE Modification in AD 80

1.4.8 Summary. 80

1.5 Research Relevance 83

Chapter 2 Oxidative Modification of Neurofilament Proteins 84

2.1 Introduction 85

2.2 Materials and Methods. 94

2.3 Results 99

2.4 Discussion 111

Chapter 3 Cellular Protective Function of Neurofilament Heavy Subunit 121

3.1 Introduction 122

3.2 Materials and Methods. 128

3.3 Results 132

3.4 Discussion 140

Chapter 4 Oxidative Modification of Tau Protein Contribute to NFT

Formation. 144

4.1 Introduction 145

4.2 Materials and Methods. 155

4.3 Results 159

4.4 Discussion 165

3 Chapter 5 Conclusion and Future Work. 169

References. 176

Appendix Publications, Abstracts and Conferences 235

4 List of Tables

Table 1.1 Comparison of the structures of neurofilaments

and other type IV filaments 29

Table 1.2 The summary and comparison of NF subunits

knock-out and overexpression mouse models 42

Table 2.1 Comparison of all the proteins and peptides in

these studies. 116

Table 2.2 Protein database search for KSP repeats. 117

Table 2.3 All the control peptides used in these studies

did not show high level of HNE adduct. 118

5 List of Figures

Figure 1.1 Oxidative stress is a prominent central event

in AD pathogenesis. 25

Figure 1.2 Comparison of the structures of neurofilam-

ents and other type IV filaments. 31

Figure 1.3 Schematic model of neurofilament assembly 33

Figure 1.4 Regulatory functions of NF phosphorylation 36

Figure 1.5 Hypothetical relationships among neurofil-

ament proteins, oxidative stress, cell injury

and tissue damage 39

Figure 1.6 The structure of HNE and its adducts with amino acids. 81

Figure 2.1 Phosphorylation state of NFH regulates

level of HNE-NFH adduct 100

Figure 2.2 Tau protein did not show the similar effect. 101

Figure 2.3 C-terminus of NFH is very reactive with HNE. 102

Figure 2.4 Nonphospho-20 AA KSP peptides are not

able to be intensively modified by HNE. 104

Figure 2.5 Phospho-AKSPV peptide is modified by HNE. 105

Figure 2.6 Synthetic K-S-P peptides with and without

phosphation show different reactivity with

HNE 106

Figure 2.7 Synthetic K-S peptides with and without

phosphorylation did not show significant

6 difference in the formation of HNE adduct. 108

Figure 2.8 HPLC-ESI-MS/MS confirms the removal

of HNE from the HNE adduct after depho-

sphorylation of the sample 109

Figure 2.9 Hypothetical model for the phosphorylation-

regulated HNE modification in KSP motif. 121

Figure 3.1 HNE treatment induced higher level of

NFH in M17 cells 133

Figure 3.2 HNE treatment induced higher phosphoryl-

ation level of NFH. 134

Figure 3.3 MAPK were activated in M17 cells after

HNE treatment 135

Figure 3.4 M17 differentiated cells with higher level of

NFH showed significant protection form HNE

cytotoxicity with LDH cytotoxicity assay 137

Figure 3.5 M17 differentiated cells with higher level of

NFH showed significant protection form HNE

cytotoxicity with LDH cytotoxicity assay 138

Figure 3.6 N2A cells showed significant protection with

overexpressing NFH using trypan blue staining. 139

Figure 4.1 Summary of the specificity of seven NFT antibodies. 161

Figure 4.2 Recognition of various τ forms with and without

dephosphorylation by antibodies to NFT. 162

7 Figure 4.3 Enhancement (fold increase) of antibody

recognition by HNE modification 163

Figure 4.4 Immunocytochemistry on adjacent serial

sections with the antibodies to NFT. 164

8 ACKNOWLEDGEMENTS

I would like to express my sincere gratitude to my dissertation advisor, Professor

George Perry, for his never-ending guidance and support during my Ph.D. research. He has continuously provided me with enthusiasm, vision, and wisdom, and inspired me from the beginning to the end. He also created an environment that gave me the flexibility to explore new ideas, while helping me to make critical decisions whenever the project was at a crossroad. He is an outstanding researcher to have as a role model for a

Ph.D. student.

I also want to thank my committee chair Professor Mark A. Smith, who provided knowledge and skills for me anytime during my Ph.D. studies, and motivated me to improve my basic knowledge in all aspects. It is he that really made me to write scientific papers and accomplish all the things in my Ph.D. training process.

I have also had the good fortune to work with Professor Lawrence M. Sayre, who has helped me in every aspect of the thesis, which is deeply appreciated. I admire him for his broad knowledge, creative thinking, and deep insight in experimental design and ability for solving problems. Discussions with Professor Sayre directly led to the results of the neurofilament studies in chapter 2.

I also thank Professor Shu G. Chen, who really helped me with the equipment and in producing good data. He also directly helped me with the methods and experimental design, which put me on the right track so many times. He has always been supportive for my research work and contributed extensively to my Ph.D. research.

9 It has been a privilege to work together with these intelligent and friendly people in my laboratory and co-laboratories. They have given me all the support from time to time. Here I send my gratitude to Sandra L. Siedlak, Peggy L.R. Harris, Beth Kumar,

Xiongwei Zhu, Kazuhiro Honda, Paula I. Moreira, Dandan Wang, De Lin, David R. Sell, and Dr. Ricardo B. Maccioni, Dr. Jesus Avila, Dr. Mervyn J. Monteiro, Dr. Don W.

Cleveland, Dr. Robert Salomon, Dr. Vincent Monnier, Dr. Michael Kinter, Dr. Touradj

Solouki, Dr. Michael Strong, and so many other people that have already contributed to my Ph. D. research work. I really had a good time with them and had been honored to have all these colleagues to work with.

Last, I would like to thank my family. All the family members are so supportive for my studies, my personal life and my future plans, in spite of they are thousand miles away or nearby. They are the best family members I know in my life. Surely they are worthy of my most sincerely appreciation.

This work has been supported by the NIH and Alzheimer Association. Their assistance is gratefully acknowledged.

10 List of Abbreviations

Aβ amyloid-β protein

AβPP amyloid-β precursor protein

AD Alzheimer’s disease

ALE advanced lipoxidation end-product

ALS Amyotrophic Lateral Sclerosis

AP alkaline phosphatase

AR alkoxyl radical

BME. β-mercaptoethanol

CaM. calmodulin MARK kinases

Cdk5. Cyclin-dependent kinase 5

CNBr cyanogen bromide

CNS central nervous system

CMT. Charcot-Marie-Tooth Disease

CSF. cerebrospinal fluid

CWRU Case Western Reserve University

DNPH. 2,4-dinitrophenylhydrazine

ECL enhanced chemiluminescence

ERK1/2DAB 3,3’-diaminobenzidine

FRTA free radical theory of aging

FTDP-17 Frontotemporal Dementia with Parkinsonism

Linked to Chromosome 17

GSH glutathione

11 GSK3. glycogen synthase kinase 3

HF hydrofluoric acid

HNE. 4-hydroxy-2-nonenal

H2O2 hydrogen peroxide

HOCl hypochlorous acid

HPLC-ESI-MS/MS high performance liquid chromatography-electrospray

ionization-mass spectrometry/mass spectrometry

IF intermediate filament

JNK Jun N-terminal kinase

KSP Lysine-Serine-Proline

LDH. lactate dehydrogenase

MALDI Matrix-assisted Laser Desorption/Ionization mass

spectrometry

MAPs. microtubule-associated proteins

MAPK Mitogen-Activated Protein Kinase

MF microfilament

MT microtubule

NF-L. neurofilament proteins light subunit

NF-M neurofilament proteins medium subunit

NF-H. neurofilament proteins heavy subunit

NEFL neurofilament proteins light subunit gene

NEFM. neurofilament proteins medium subunit gene

NEFH. neurofilament proteins heavy subunit gene

12 NFPs. neurofilament proteins

NFTs. neurofibrillary tangles

NOR. nitric oxide radical

O-GlcNAc O-linked N-acetylglucosamine

PBS phosphate-buffered saline

PD Parkinson's Disease

PHF paired helical filaments

PKA protein kinase A

PKC protein kinase C

PLC phospholipase C

PNS peripheral nervous system

PP2A protein phosphatase 2A

PR. peroxyl radical

PSP. Progressive supranuclear palsy

RA retinoic acid

ROS Reactive oxygen species sGAGs sulfoglycosaminoglycans

SDS-PAGE. Sodium Dodecyl Sulfate Polyacrylamide

Gel Electrophoresis

SOD superoxide dismutases

SCZ Schizophrenia

TBS Tris-buffered saline

TCA trichloroacetic acid

13 TEM. transmission electron microscopy

TMAO trimethylamine N-oxide

UMLS Unified Medical Language System

14 Phosphorylation and Sequence Dependency of Neurofilament Protein Oxidative

Modification in Alzheimer disease

Abstract

By

QUAN LIU

Protein adducts of the lipid peroxidation product 4-hydroxy-2-nonenal (HNE) are prominent features of oxidative damage in neuronal cell bodies in Alzheimer disease

(AD). Such adducts are also seen in axons of normal as well as diseased individuals.

These HNE adducts are constant throughout the axons and the aging process of humans, mice, and rats, indicating that HNE adduction is physiological and regulated from birth to senility. In axons, HNE is primarily adducted to the neurofilament heavy subunit (NFH) and the neurofilament medium subunit (NFM), and limited to lysine residues.

Interestingly, we found that phosphorylation is essential since formation and removal of

HNE adducts are controlled by the NFH/M phosphorylation state. Studies using immunochemistry, synthetic peptides, mass spectrometry and chemical stabilization of

HNE adducts, demonstrated that XKSPX, the most repeated sequences in NFH/M, are the major component in neurons highly susceptible to phosphorylation regulated aldehyde adduction. To our knowledge, this is the first study to directly show phosphorylation can regulate NFH-HNE levels in axons and the multiple KSP repeats are the critical motifs which preserve the phosphorylation-dependent regulation. This study provides the new evidence that indicates signal transduction could modulate oxidative modification in brain through activation of kinases and phosphatases.

15

Chapter 1

General Introduction

16 1.1 Introduction of Alzheimer Disease

1.1.1 General Information

Alzheimer’s disease (AD) was first reported by Dr. Alois Alzheimer, a Germany

doctor, during the 1907 neurology conference, in which he described a 51-year-old

woman developing rapid memory degeneration and died with severe dementia several

years later (Alzheimer, 1907; Alzheimer et al., 1995). Although the disease was once

considered rare, it is now established as the leading cause of dementia. An estimated 4

million individuals in the United States have Alzheimer’s disease with an annual expense

of about $70-$100 billion. Also as many as 350,000 more individuals develop the disease

each year (Mayeux et al., 2003). There is still no cure for AD, and the only therapy

known as cholinesterase inhibitors or anti-cholinesterase drugs including Reminyl

(galantamine), Aricept (donepezil hydrochloride) and Exelon (rivastigmine) act to slow

the disease progression (Alzheimer’s Association, 2004). In October 2003, Namenda

(Memantine), a new drug as a NMDA antagonist, was approved by FDA to be

used on moderate to severe Alzheimer patients (Alzheimer’s Association, 2004).

Alzheimer’s disease involves the parts of the brain that control thought, memory,

and language. After last two decades of intensive study, more information has been

obtained, but the cause of AD is still unknown. There are several major theories such as beta amyloid toxicity (Selkoe, 2000), tauopathy (Lee et al., 2001), inflammation (McGeer et al., 2001; Weiner et al., 2002), oxidative stress (Markesbery, 1997; Christen, 2000;

Picklo et al., 2002, Perry et al., 2002), all of which have been discussed in the literature

intensively.

1.1.2 Clinical

17 The risk of Alzheimer’s disease varies from 12% to 19% for women over the age

of 65 years and 6% to 10% for men (Seshadri et al., 1997). On average, AD patients live

for about 8 years after they are diagnosed, although the disease can last for as many as 20

years. The areas of the brain that control memory and thinking skills are affected first, but as the disease progresses, cells die in other regions of the brain. Eventually, the person with AD will need complete care. If the individual has no other serious illness, the loss of brain function itself will cause death.

1.1.3 Pathology

The hallmarks of Alzheimer’s disease are two distinctive lesions in the patient’s brain, senile plaques and neurofibrillary tangles (NFTs) (Alzheimer et al., 1995). They were identified by silver staining of the AD patient’s brains. In addition, other pathological changes are detected in the AD patient’s brains, such as neuronal and dendritic loss, neurotrophic threads, dystrophic neuritis, granulovacuolar degeneration,

Hirano bodies, cerebrovascular amyloid, and atrophy of the brain (Smith, 1998).

Senile plaques are described originally as “millary foci” (Alzheimer et al., 1995).

They are spherical extracellular lesions with 10-200 µm in diameter with a central core made up by bundles of 6-10 nm amyloid-β protein (Aβ) (Smith, 1998). In the peripheral region of the senile plaques, Aβ protein, amyloid-β precursor protein (AβPP), τ, and neurofilament proteins are reported (Perry et al., 1988).

NFTs are a major intracellular protein aggregation found in AD brains. They are located primarily in the cerebral cortex, especially in the large pyramidal neurons in the hippocampal and frontotemporal regions (Smith, 1998). NFTs are composed of bundles

of paired helical filaments (PHF), the major component of which is the

18 microtubule-associated protein τ (Grundke-Iqbal and Iqbal, 1989; Lee et al., 1991).

Moreover, neurofilament proteins are also reported (Perry et al., 1988). In PHF, τ is abnormally hyperphosphorylated (Grundke-Iqbal et al., 1986; Trojanowski and Lee,

1995; Lee et al., 2001), ubiquitinated (Iqbal and Grundke-Iqbal, 1997; Mori et al., 1987;

Perry et al., 1987), oxidized (Mattson, 1995; Takeda et al., 2000), proteolytically processed and aggregated into filaments (Lee et al., 2001; Lovestone and McLoughlin,

2002; Perez et al., 2002). Hyperphosphorylation of τ renders it unable to bind to microtubules and therefore unable to promote or maintain microtubule assembly (Baudier and Cole, 1987). Moreover, hyperphosphorylation makes τ more resistant to proteolytic degradation, which may play a key role in neurofibrillary degeneration in AD patients

(Iqbal et al., 1998; Eidenmuller et al., 2000).

1.1.4 Etiology

A Age

Age is the single greatest risk factor in the etiology of AD. AD rarely occurs in people younger than 60 years old, and affects 10-15% of individuals over 65 years old and up to 47% of individuals over age of 80 (Evans et al., 1989). This predominance of age as a major cause in AD etiology indicates that some age-related events are closely involved in the development of the disease. However, the specific aging process related to AD pathogenesis is unknown.

B Oxidative Stress (OS)

As one of the primary theories of aging, free radicals (Oxidative stress) may play a major role in the AD pathogenesis. Oxidative stress is a potential source of damage to

DNA, lipids, sugars and proteins within cells, imbalance between the intracellular

19 production of free radical species and antioxidant defense mechanisms results in oxidative stress (Perry et al., 2002). In a post-mitotic environment such as neurons, oxidative stress can be treated as a marker of age-related deterioration in cellular homeostatic mechanisms. Since the neurons have a diminished capacity to deal with redox imbalance, even minor cellular stresses have the ability to lead to irreversible injury and such contribute to the pathogenesis of neurodegenerative disease. Reactive oxygen species (ROS), which are the primary mediators of oxidative injury, cause damage to lipids, sugars and amino acid side-chains (Picklo et al., 2002). For example, oxidative deamination of lysine and deguanidination of arginine results in protein-based aldehyde groups that can be detected by 2,4-dinitrophenylhydrazine (DNPH) (Levine et al., 1994). However, it appears that most DNPH-detectable carbonyls found on proteins are resulted from modification by lipid and sugar oxidation products, which act as secondary toxins. Such toxic carbonyl species have been shown to play a role in the pathophysiology of AD (Smith et al., 1994a; Yan et al., 1994; Sayre et al., 1997;

Castellani et al., 2001).

C Genetics

Although familial AD only accounts for a small percentage of AD cases and the majority of the AD cases are spontaneous. There is a strong evidence related with the genetic component in AD occurrence. Up to 35% of the AD patients have affected first degree relatives (Breitner and Folstein, 1984). Now the greatest correlated genetic factor is the polymorphism of ApoE gene, which may correlate with 50% of the AD patients especially for ApoE4 alleles, but the mechanisms are unknown (Ashford, 2004; Teter,

20 2004). Mutations in beta amyloid protein and presenilin 1 and 2 genes are considered as critical factors in the AD pathogenesis.

ApoE genes

ApoE genes are found to be related with AD occurrence. ApoE is an abundant

34-kDa glycoprotein that is synthesized and secreted mainly by astrocytes and microglia in the central nervous system (CNS). It is well established that ApoE, specifically the E4 allele of ApoE, is a major genetic risk factor for the more common, late-onset form of

AD (Saunders et al., 1993; Rebeck et al., 1993). ApoE genotype also seems to be a determinant of brain Aβ burden in individuals affected with AD (Strittmatter et al., 1993;

Schmechel et al., 1993). Now the polymorphism of ApoE gene is considered the greatest correlated genetic factor in AD pathogenesis (Ashford, 2004; Teter, 2004).

Amyloid β precursor protein

Aβ protein is the major component of senile plaque cores and it is derived from a precursor protein called amyloid β precursor protein (AβPP). Amyloid β precursor protein gene is located on chromosome 21 (21q11-22) (St George-Hyslop et al., 1987;

Tanzi et al., 1987). The normal function of AβPP is unknown, but it is shown to be involved in several or more broad physiological functions in neurons. The mutations in

AβPP results in a great increase of the Aβ protein production, which may cause the extracellular protein aggregation (Selkoe 1997, 2001). The AβPP mutant mice have an overproduction of Aβ protein and demonstrate senile plaque formation and synaptic deficits without NFTs pathology, which indicates a key pathological role of the mutation in AβPP protein (Games et al., 1995; Hsiao et al., 1996).

Presenilin 1 and 2

21 The majority (~70%) of early-onset familial AD cases are associated with

mutations in two genes, presenilin 1 and presenilin 2, which are located on chromosome

14 and 1 respectively (Rogaev et al., 1995). Over 50 different pathogenic mutations in

presenilin 1 gene and 3 mutations in presenilin 2 gene have been described (Checler,

1999). There is considerable homology between the gene products of presenilin 1 and

presenilin 2, which are transmembrane proteins of 463 and 448 amino acids respectively,

both have six to nine hydrophobic membrance-spaning domains (Rogaev et al., 1995).

The physiological functions of these two proteins are unknown, SEL-12, one of the

homolog proteins in Caenorhabditis elegans involved in the Notch receptor pathway

(Sherrington et al., 1995), indicates a potential role for preseniline involving in Notch pathway. Nonetheless, other possible roles include ion channel transport, protein processing, or cellular trafficking functions (Czech et al., 2000). In AD, it is thought mutations in these proteins contribute to the disease by affecting the processing of AβPP, which leads an increase of the Aβ burden in both AD patients and transgenic mice

(Citron et al., 1997; Scheuner et al., 1996).

D Other Factors

CARDIOVASCULAR DISEASE AND RELATED VASCULAR RISK

FACTORS Hyperlipidemia, hypertension, diabetes, and related factors usually leading to the occurrence of heart disease or strokes have been identified as putative antecedents to

Alzheimer’s disease (Breteler, 2000).

DOWN SYNDROME Adults with Down’s syndrome develop the neuropathological changes of Alzheimer’s disease by age 40, but not all patients become demented. The risk of Alzheimer’s disease associated with a family history of Down’s

22 syndrome is increased two to three fold.

ALCOHOL Individuals who drank wine in moderate amounts daily were less

likely to develop Alzheimer’s disease than both heavier drinkers and abstainers

(Orgogozo et al. 1997). The risk reduction associated with alcohol is possibly related to its antioxidant properties or its effects on lipid metabolism.

EDUCATION AND EARLY LIFE EXPERIENCE Several studies showed that

the risk of Alzheimer’s disease among poorly educated individuals or individuals in poor

living condition is significantly higher than those among well-educated persons (Stern et

al., 1994; Hall et al., 2000).

MENTAL AND LEISURE ACTIVITY Mental and physical exercises are

positive factors with a lower risk of AD (Lindsay et al., 2002).

SMOKING Smokers had a two- to four-fold increase in risk of Alzheimer’s

disease, particularly those individuals without an APOE-4 allele (Merchant et al., 1999;

Ott et al., 1998).

TRAUMATIC HEAD INJURY There is an increase of the risk of Alzheimer’s

disease related with traumatic head injury (Plassman et al., 2000).

ANTI-INFLAMMATORY AGENTS Alzheimer’s disease was found to be less

frequent among individuals who used anti-inflammatory agents (in t’ Veld et al., 2001;

Weggen et al., 2001).

ANTIOXIDANTS That antioxidants have been proposed as a protective factor

for poor memory and dementia are based on their oxidative stress-reducing effect in vitro.

Studies show that healthy individuals consuming dietary or supplemental antioxidants

demonstrate inconsistent and small difference (Engelhart et al., 2002; Morris et al., 2002).

23 Therefore, if at an earlier stage and long-term usage of antioxidant need to be further

explored.

HORMONE REPLACEMENT The use of estrogen by postmenopausal women

has been associated with a decreased risk of Alzheimer’s disease (Baldereschi et al., 1998;

Waring et al., 1999). Women using hormone replacement had about a 50% reduction in the disease risk.

1.1.5 Research Relevance: Oxidative Stress and AD Pathogenesis

There is profound evidence showing that the oxidative damage is the very

earliest neuronal and pathological changes of AD (Perry et al., 1998; Nunomura et al.,

2000, 2001; Pratico et al., 2001). In addition, many AD risk factors are related with

production of free radicals in their hypothetical models (Figure 1.1). Indeed, this early

role is borne out by clinical management of oxidative stress, which appears to reduce the

incidence and severity of AD (Sano et al., 1997; Stewart et al., 1997). Many markers of

oxidative damage are present in susceptible neurons even in the absence of

neurofibrillary pathology (Nunomura et al., 1999). A recent time course study in Aβ

transgenic mice confirmed these findings: lipid peroxidation is significantly increased in

7–8-month-old Tg2576 mice, proceeding any apparent Aβ deposition or increase of Aβ

levels (Pratico et al., 2001).

Oxidative stress may play a role in the development of neuritic abnormalities

since paired helical filaments (PHF) are more often found in neurites with membrane

abnormalities indicative of extensive lipid peroxidation (Praprotnik et al., 1996).

24

Figure 1.1 Oxidative stress is a prominent central event in AD pathogenesis.

Many AD risk factors are related to oxidative stress and most of the AD risk factors will directly cause production of free radicals.

25 Moreover, crosslinking of proteins by oxidative processes may lead to the

resistance of the lesion removal in AD even though they are extensively ubiquitinated

(Cras et al., 1995). Moreover, this resistance of neurofibrillary tangles to proteolysis

might play an important role in the progression of AD (Smith et al., 1998). Lipid

peroxidation is marked by significantly increased levels of thiobarbituric acid reactive

substances (TBARS), malondialdehyde (MDA), 4-hydroxy-2-transnonenal (HNE), and

altered phospholipid composition (Sayre et al., 1997). Oxidation of sugars is marked by

increased glycation and glycoxidation (Vitek et al., 1994; Smith et al., 1994, 1995;

Castellani et al., 2001).

Tau and neurofilament proteins have already been found in NFTs with

predominant oxidative modifications. It is of particular interest to study the role of oxidative modification of these proteins in neurons, or more specific, to examine whether it is directly related to pathogenesis of AD.

26 1.2 Introduction of Neurofilament Proteins

The function of neurofilaments, the major component in large myelinated neurons, has not been well understood even though they were discovered as structures

over 100 years ago. Recent studies have suggested that neurofilaments are closely related

to many neurodegenerative diseases, such as amyotrophic lateral sclerosis, Parkinson

disease, Alzheimer disease, and diabetes. Using in vitro assays, cultures, and transgenic

mice, these studies provided new insights into neurofilament function. The function of

each subunit, the relationship of neuroflaments with other cytoskeletal elements, and the

clinical significance, are topic of increasing attentions.

1.2.1 General Information

Neurofilaments (NF) are intermediate filaments of neurons that are considered to

add rigidity, tensile strength, and possibly intracellular transport guidance to axons and

dendrites (UMLS, 2004). Exclusively expressed in neurons, NFs are one of the

members of the cytoskeleton proteins that act together to form and maintain the cell

shape and facilitate the transport of particles and organelles within the cytoplasm.

Based on the differences in diameter and protein components, cytoskeletal

polymers are classified into three groups: microtubule (MT) (~24 nm), microfilament

(MF) (~6-8 nm), and intermediate filament (IF) (~10 nm). MTs, which are composed of

tubulin, are responsible for maintaining cell shape, organelle and vesicle movement, and

the formation of the spindle fibers during mitosis. MFs, which are composed

predominantly of actin, are responsible for cellular movement, muscle contraction,

mechanical strength, and cytokinesis. IFs are composed of different proteins and are

prominent in cells that withstand mechanical stress. Moreover, IFs are the most insoluble

27 components of the cells. While the polymer compositions of MTs and MFs in all tissues are similar, IF polymers differ in different tissues and cells. Based on the homology in molecular structure, five types of IFs have been identified and NFs belongs to the type IV

IFs (Table 1).

1.2.2 History

As early as the 1830s, neuronal networks had already been described (Valentin,

1836; Purkinje, 1838). The discovery of the silver staining method in the late nineteenth century resulted in a clear vision of NFs (Apathy, 1897; Cajal, 1903), which also led to the characterization of neurofibrillary tangles and senile plaques by Alois Alzheimer

(Alzheimer, 1906). With the development of electron microscopy in 1931 (Max Knoll and Ernst Ruska in Germany), the molecular structures of NFs were further defined as filaments ~10 nm in diameter and present exclusively in neuronal cells. For a long time, it has been known that NFs are involved in several neuronal diseases. In the past decade,

NFs have been linked to other human diseases, which will be discussed here. With the development of specific antibodies, transgenic animal models, and molecular genetic methodologies, the study of NFs has advanced to the molecular level.

1.2.3 Structure

Together with peripherin, α-internexin, and nestin, NFs belong to type IV IFs and share common sequence structures (Table 1). A central α-helical rod domain of about

310 amino acids is flanked by a globular N-terminal region and non α-helical carboxy-terminal sidearm domains. The central rod domains including region 1a, 1b, and

2 contain highly conserved motifs and every seventh residue is hydrophobic, which

28

Table1.1 Intermediate filaments include five defined types and other undefined types. Their molecular weights are varied and they are found in different cell types.

Neurofilament subunits belong to type IV intermediate filaments.

29 facilitates the formation of α-helical coiled-coil parallel homodimers or heterodimers. A linker region aligns the hydrophobic residues. These properties are characteristics of the intermediate filaments and are essential for their proper assembly (Fig 1.2).

NFs are composed of three subunits, and these subunits are defined by their molecular weights: NF-L (light), NF-M (medium), and NF-H (heavy), which are 62 kDa,

97 kDa, and 125 kDa respectively, as predicted from the DNA sequences (Shaw, 1991).

These subunits exhibit higher molecular weights on SDS-PAGE: 68 kDa, 160 kDa and

205 kDa respectively, due to the enriched negatively charged amino acids (glutamic acid) in their sequences and post-translational modifications such as phosphorylation and glycosylation (Shaw, 1991; Dong et al., 1993). NFs constitute the most abundant structures in large myelinated neurons, and can account for 13% of total proteins and

54% of the Triton-insoluble proteins in some neurons (Morris and Lasek, 1984). More detailed description of all intermediate filaments structures is recorded in several reviews

(Shaw, 1991; Fuchs E. and Weber, 1994; Al-Chalabi and Miller, 2003).

1.2.4 Expression

The three NF subunits are expressed during distinct stages of vertebrate development, triggered by neuron differentiation (Shaw and Weber, 1982; Nixon and

Shea, 1992). Initially, NF-L is expressed at the beginning of neuronal differentiation and overlapped with the expression of α-internexin and peripherin (Carden et al., 1987;

Willard et al., 1983). NF-M is expressed shortly after with the emergence of neurite formation whereas NF-H is expressed later when axonal radial growth is required for nervous system maturation (Carden et al., 1987; Willard et al., 1983). The genes coding for human NFs include NEFL gene (NF-L protein), NEFM gene (NF-M protein),

30

Figure 1.2 Comparison of the structures of neurofilaments and other type IV

filaments. The neurofilament subunits and other type IV IFs all include α-helical rod domain and are varied in N- and C- termini. A unique character of NF-M and NF-H is that the C-termini of them have multiple KSP repeats that are heavily phosphorylated.

Human NFM has 13 KSP and human NFH has 43/44 KSP, and most of them are phosphorylated. In addition, here shows the posttranslational modifications including phosphorylation and glycosylation on NF subunits (Summarized from Shaw 1991; Dong et al., 1993; Fuchs and Weber 1994).

31 and NEFH gene (NF-H protein). NEFL and NEFM genes are closely located on

chromosome 8, 8p21 (Myers et al., 1987; Strausberg 2002); NEFH gene is at

chromosome 22, 22q12.2 (Lees et al., 1988). Expression of mRNAs of NF subunits is

consistent with their protein levels in cultured neuroblastoma cells and NF-L and NF-M

mRNAs are expressed several days before the appearance of NF-H mRNA in these cells

(Breen and Anderton, 1991) It appears that the NF-L and NF-M genes are coordinately

regulated while the NF-H gene is expressed independently later.

1.2.5 Assembly

NF assembly is not clearly understood yet, and it is directly related to the

functions of the three subunits and the filament. In vivo studies have demonstrated that

NFs are obligate heteropolymers requiring NF-L to form a proper polymer with either

NF-M or NF-H (Lee et al., 1993). NF formation is believed to start with the dimerization of NF-L with either NF-M or NF-H subunits. The highly conserved rod domains of the

NF subunits are coiled together in a head-to-tail fashion to form a dimer. Two coiled dimers overlap with each other in an anti-parallel, half-staggered manner forming the tetramers. Finally, eight tetramers are packed laterally and longitudinally together in a helical array, forming a ropelike 10 nm filament (Fuchs and Weber, 1994, Heins et al,

1994, Fuchs and Cleveland, 1998; Herrmann and Aebi, 2000). In the cross section of NF, there are ~32 molecules though the number may change in different stages or conditions

(Herrmann et al., 1999). Theoretically, before the synthesis of NF-H, NFs are exclusively composed of NF-L and NF-M (Carden et al., 1987). The C-termini of NF-M and NF-H are not in the coils but they form the side-arms of the NFs (Perrone Capano et al., 2001)

(Figure 1.3). (See related study in Chapter 2)

32

Figure 1.3 Schematic model of neurofilament assembly. The neurofilament assembly process includes, two NF subunits (NF-L and either NF-H or NF-M) form head-to-tail coiled-coil dimmers, anti-parallel half-staggered tetramers, protofilament, and 10nm neurofilament. There are around 32 molecules in the cross section and hyperphosphorylated side-arms formed by C-termini of NF-H and NF-M stick out of the stem of the filament.

33 NF-L is essential for the precise assembly of NFs (Ching and Liem, 1993; Lee et

al., 1993, Zhu et al., 1997). NF-M participates in the formation of cross-bridges,

stabilization of the filament network, and the longitudinal extension (Elder et al., 1998a,

1999a, b; Jacomy et al., 1999). NF-H also contributes to the formation of cross-bridges and may interact with microtubules/microfilaments and other cytoskeletal elements

(Elder et al., 1998b 1999b; Jacomy et al., 1999). However, recent studies showed NF-H side arms are not critical for the radial growth of the axons (Rao et al., 2002, Al-Chalabi et al., 2003). NF-M and NF-H by themselves do not form filaments in vivo (Ching and

Liem, 1993; Lee et al., 1993, Nakagawa et al., 1995, Zhu et al., 1997), while NF-L can form homopolymer in vitro (Geisler et al., 1981; Liem et al., 1982) and in cells transfected with NF-L (Carter et al., 1998), but not in mice (Jacomy et al., 1999).

Although NFs are obligate heteropolymers in vivo in rodent, this may not be the case in human (Carter et al., 1998).

In contrast to NF subunits, α-internexin and peripherin self-assemble into homopolymers and co-assemble with NF subunits (Cui et al., 1995; Beaulieu et al., 1999a, b). Many CNS neurons also contain varying levels of α-internexin, a type IV intermediate filament subunit (Fliegner and Liem, 1991) expressed abundantly in embryonic neurons but at lower levels postnatally (Kaplan et al., 1990; Fliegner et al., 1994). Although

α-internexin colocalizes with NFs and forms filaments with each of the triplet proteins in

vivo (Ching and Liem, 1993), the composition of the filaments is still not known in vivo.

Peripherin, a type III intermediate filament, is expressed in lower motor, autonomic, and sensory neurons and co-assembles with the NF subunits (Beaulieu et al., 1999a, b).

1.2.6 Transport

34 After their synthesis in the perikarya, neurofilament proteins (NFPs) are quickly translocated into the axons and assemble into filamentous structures. In radioisotopic labeling studies, neurofilament proteins move at a rate of 0.2–1 mm/day in the axon with the slow components whereas other organelles such as membrane vesicles can move at much faster rate of 200–400 mm/day (Lariviere and Julien 2004). (See more details in

Chapter 2)

1.2.7 Posttranslational Modifications

Two major posttranslational modifications, phosphorylation and glycosylation, are involved in the formation and functions of NFs.

A. Phosphorylation (See more details in Chapter 2) (Figure 1.4)

B. Glycosylation.

In addition to phosphorylation, NFs are also modified post-translationally by

O-linked N-acetylglucosamine (O-GlcNAc) on serine and threonine residues (Dong et al.,

1993, 1996), which is the simplest protein modification with sugars (Hart 1989). Initially discovered from murine lymphocytes (Torres et al., 1984), this modification has been found in all the compartments of eukaryotic cells (Holt et al., 1986). Similar to phosphorylation, the O-GlcNAc modification (O-GlcNAcylation) is highly abundant and dynamic (Hart et al., 1995; Hart et al., 1993; Haltiwanger et al., 1992; Chou et al., 1992).

NF-H, NF-M, and NF-L are modified by O-GlcNAc in vivo to stoichiometries of at least 0.3, 0.15, and 0.1 mol GlcNAc/mol of protein at both termini (Dong et al., 1996)

(Fig 1). Although the stoichiometry of O-GlcNAcylation on isolated NFPs is very low compared with phosphorylation, it seems likely that only a subset of NF subunits is modified by O-GlcNAc, and the stoichiometry in those subunits is probably much higher.

35

Figure 1.4 Regulatory functions of NF phosphorylation. N-terminal phosphorylation has both inhibitory and promotional effect for the C-terminal phosphorylation in NFs, and N-terminal phosphorylation can also control the NFs translocation and assembly in axons.

36 It is worthy of mentioning that this O-GlcNAcylation is most likely in a consensus of PX0-4(S/T) motif (X is usually a hydrophobic residue) similar to the motifs used by proline-directed kinases and phosphatases (X[S/T]PX or XP[S/T]X) (Hart et al.,

1995).

The proximity or competition of the O-GlcNAcylation sites to the phosphorylation sites in the head domain and the importance of the head domain in NF assembly (Gill et al., 1990; Wong et al., 1990; Chin et al., 1991) indicates that

O-GlcNAcylation may also play a role in NF assembly. This regulation could be exerted through either direct or indirect (e.g. by affecting phosphorylation) influence on the structure of the head domain. It is possible that O-GlcNAc is added to NFPs by

O-GlcNAc transferase (Haltiwanger et al., 1992) right after their synthesis and prior to their assembly into filaments and transport into the distal part of axons, and is removed by an N-acetyl-β-D-glucosaminidase subsequently (Dong et al., 1993).

By replacing phosphates with O-GlcNAc, the interactions between NFs may switch from a repulsive one to an associative one, leading to close packing of NFs, which happens in nodes of Ranvier. Therefore, it seems likely that organization of NFs is regulated by kinase/phosphatase (Nixon 1993; Xu et al., 1994) and O-GlcNAc transferase/N-acetyl-β-D-glucosaminidase (Haltiwanger et al., 1992; Dong et al., 1994).

The dynamic O-GlcNAcylation (Hart et al., 1993; Hart et al., 1995) and phosphorylation

(Nixon 1993) could therefore regulate proper NF assembly and dynamics, and the abnormalities in either of them could contribute to some of the motor neuron diseases in which NFs accumulate aberrantly (Xu et al., 1994; Cleveland et al., 1995).

1.2.8 Degradation

37 Many neurons extend their axons over great distances, up to 1 meter in human

sciatic nerves, to form synapses with appropriate receptor cells. To maintain the

physiological functions of the nerves, certain proteins need to have long lifetimes to span

over the axon. NFPs are among one of them. In the transport process NFP degradation is

not detected (See more details in Chapter 2).

1.2.9 Functions (See more details in Chapter 3)

There are several aspects of NF function. These include the basic function of neurofilament to support axonal structure, detrimental effects of NF accumulations, NF protective effects (Figure 1.5), and the possible function of KSP repeats.

1.2.10 Animal Models

Many animal models have contributed to the studies of NFs in establishing NF subunits functions. The first model is the Japanese quail discovered in 1991 (Quiverer)

(Yamasaki et al., 1991) with a spontaneous mutation in NF-L at amino acid 114, a nonsense mutant, which generates a truncated protein of only 113 amino acids of NF-L out of total 556 amino acids and is incapable of forming NFs (Ohara et al., 1993).

Homozygous mutants contain no axonal NFs and exhibit a mild generalized quivering. In

these animals, radial growth of myelinated axons is severely attenuated (Yamasaki et al.,

1991) with a consequent reduction in axonal conduction velocity (Sakaguchi et al., 1993).

The second model is the transgenic mouse reported in 1994, expressing a

NF-H/β-galactosidase fusion protein in which the C-terminus of NF-H was replaced by

β-galactosidase (Eyer and Petersen, 1994). NF inclusions in the perikarya of neurons, depletion of axonal NFs, and reduced calibers were found. Later on, Zhu et al (1997) have shown that mice lacking NFs due to a targeted disruption of the NF-L gene have

38

Figure 1.5 Hypothetical relationships among neurofilament proteins, oxidative stress, cell injury and tissue damage. Neurofilament especially NFH as the most abundant protein in neurons may work as the scavenger for oxidative stress, which protects the other critical factors from oxidative attack.

39 reduced axonal calibers and delayed maturation of regenerating myelinated axons.

To demonstrate the NF subunits functions, we will only compare the NF subunit knock-out and overexpressing mice in this paper (Table 1.2). To see more transgenic mice comparisons, other reviews are recommended (Julien 1999; Lariviere and Julien,

2004)

NF-L Knockout Mice The targeted disruption of the NF-L gene in mice confirmed the importance of NF-L in NF assembly (Zhu et al., 1997). The NF-L knock-out mice had no overt phenotype. In the absence of NF-L, the levels of NF-M and

NF-H subunits are dramatically decreased to ~5% of normal level and unable to assemble into 10 nm filaments. As a result, these mice have a scarcity of IF structures, severe axonal hypotrophy, and no large-sized myelinated axons (Zhu et al., 1997). Levels of other cytoskeletons like tubulin are increased possibly as a compensating effect. While the NF-L gene knock-out mice provide definite proof that NFs are a major determinant of axonal caliber and NF-L is required for NFs formation, the specific roles of NF-M and

NF-H subunits remain unclear. The reduced levels of NF-H and NF-M in NF-L-/- mice

are probably the result of an enhanced proteolytic turnover of unassembled or

disorganized NF proteins in the absence of NF-L (Zhu et al., 1997; Williamson et al.,

1998; Beaulieu et al., 1999a; Levavasseur et al., 1999).

NF-H Knockout Mice Several laboratories have recently reported the characterization of NF-H null mice (Rao et al., 1998; Zhu et al., 1998; Kriz et al., 2000).

Surprisingly, the absence of NF-H has little effect; the mice phenotype, the NF number, and the calibers of motor axons during development were all normal. With the C57Bl6 strain, Rao et al (1998) and Zhu et al (1998) reported only a slight reduction in the caliber

40 of myelinated axons from the ventral roots of NF-H knockout mice, maybe the small changes are due to the increase of NF-M protein level and phosphorylation levels. MTs increased two fold with a small decrease in actin level. In the mean time, the null mutant

NF-H in 129J strain mice described by Elder et al (1998) exhibited a more pronounced reduction in the caliber of axons. In addition, they reported little difference in NF-L and

NF-M levels. These discrepancies are most likely attributable to differences in the mouse genetic backgrounds. In any case, the combined results suggest that NF-H has a minor effect on the radial growth of axons and may be a key factor interacting with MT/mf.

NF-M Knockout Mice A clearer conclusion comes from the analysis of NF-M knockout mice. Two independent studies have shown that disruption of NF-M has a more severe effect than disruption of NF-H on the radial growth of large myelinated axons

(Elder et al., 1999; Jacomy et al., 1999). The absence of NF-M did not cause any phenotypic changes in mice but caused an increase in NF-H level and its phosphorylation.

However, the severe decreases in NF-L level (>50%) caused axonal NF content decrease more than 50%, resulting in axonal atrophy. Comparing NF-H-/- mice and NF-M-/- mice, it showed that NF-M is more critical in axonal radial growth.

NF-M/H Double Knockout Mice Moreover, double knockout mice lacking both

NF-M and NF-H subunits, showed hind-limb paralysis (Jacomy et al., 1999; Elder et al.,

1999). The absence of NF-M/H causes the sequestering of unassembled NF-L proteins in the neuronal perikarya, no NF formation in axons, increase of MT by 2 fold, and loss of about 24% of ventral root axons (Jacomy et al., 1999; Elder et al., 1999). The analysis of

NF-M-/-, NF-H-/- double knockout mice demonstrates that the high molecular weight subunits are required for the in vivo assembly of NFs and normal axonal growth.

41

Table 1.2 The summary and comparison of NF subunits knock-out and overexpression mouse models. “-” means not found in the original papers.

42 NF-L Overexpressing Mice Studies of transgenic mice have found that

overexpressing mouse NF-L to ~4 fold leads to a striking ALS-like pathology, in which most of the mice died before 3 weeks postnatally with only 1/3 to 2/3 of the body weight compared with normal mice (Xu et al., 1993). Also strikingly, the only two mice surviving over 3 weeks finally recovered slowly to normality after 9 months. The transgenic mice showed intensive NF aggregation in all the compartments of neurons with depleted rough endoplasmic reticulum (RER), neuron degeneration in both dorsal and ventral roots, phosphorylated NF-H in cell bodies, and severe muscle atrophy, which resembles the human ALS pathology. This study confirmed that overexpression of NF-L and NF aggregation in mice can directly cause abnormality and degeneration of motor neurons.

NF-M Overexpressing Mice The NF-M overexpressing transgenic mice appeared normal (Wong et al., 1995). With about 2 fold increase of total (WT plus mutant) NF-M level in the mice, accumulation of NFs was detected in cell bodies and axons. The axonal caliber of neurons decreased about 50% with unchanged NF-L level and 50% decrease of WT NF-H and WT NF-M levels. This caliber reduction is a result of growth retardation rather than a total growth inhibition, which was restored to the normal level with time, and it is correlated with reduction of WT NF-H and WT NF-M. No neuron degeneration was detected. Because the nearest inter-neurofilament spacing did not change, this reduction may be strictly correlated with NF numbers.

NF-H Overexpressing Mice Overexpressing 1.2 to 4.5 fold of WT NF-H in transgenic mice showed no overt phenotypes or neuron loss (Marszalek et al., 1996). The level of NF-M in axons were correlatively reduced with increased NF-H level, and NF-L

43 level decreased by 20%-40% in axons. Axonal caliber reduction was correlated with

NF-H overexpression level, more than 3 fold of which totally inhibited large axon

growth. This caliber reduction can also be considered as strictly correlated with NF-M

level or NF numbers.

In all these studies, only mice overexpressing >3 fold NF-L and NF-M/H double

knockout mice showed severe phenotypes. Taken together, these data clearly demonstrate

all the NF subunits functions: expression of NF-L is responsible for NF assembly, and

NF-M/H subunits are required for NF assembly in vivo; NF-M is more critical for axonal

growth than NF-H, and the latter may be working as a key factor for NF interaction with

other cytoskeletal polymers. The nearest inter-neurofilament spacing does not change

when NFs are present in the axons, but to achieve the long-ranged “cross bridges”, the

side chains of NF-H and other key factors are definitely required. Expression of NF-M

and NF-H are co-regulated in an inverse manner in motor axons, and they are competing

for the transport and assembly with NF-L. In large caliber axons, microtubule content

does not correlate with axonal diameter as closely as does NF content (Friede et al., 1970,

Hoffman et al., 1984), but microtubule or other IFs are more significant in maintaining the diameter of medium- and small-sized axons where they are the major cytoskeletal

components.

Noteworthy, studies other than those above also provide solid evidence for the

NF function. Recently, studies using NF-H truncated mutant mice (Rao et al., 2002)

demonstrate that neither C-terminal tail of NF-H nor its phosphorylation is important for

axonal caliber. Transgenic mice expressing human NF subunits (Tu et al., 1995, Gama

Sosa et al., 2003) or mutations in mouse NFs (Lee et al., 1994, Rao et al., 2002) all show

44 severe motor neuron degeneration, which indicate that non-WT NF forms may be the

major factors for neuronal degeneration. In addition, studies with SOD mutant mice that

utilize human NF subunits provide more important information (Julien et a., 1987, Cote

et al., 1993, Collard et al., 1995, Meier et al. 1999).

1.2.11 Neurofilaments and Neurodegenerative Diseases

Abnormal accumulation of NFs is detected in many human neurodegenerative

disorders including ALS, AD, dementia with Lewy bodies, Parkinson disease, and

diabetes, etc. Many alterations can potentially lead to accumulation of NFPs, including

deregulation of NFPs synthesis, defective axonal transport, abnormal phosphorylation, and proteolysis. Studies in transgenic mouse also showed that NFs affect the dynamics and function of other cytoskeletal elements, such as microtubules and actin filaments

(Zhu et al., 1998, Ahlijanian et al., 2000). It is widely believed that NF abnormalities in neurodegenerative disorders are the hallmark of neuronal dysfunction.

A. Amyotrophic Lateral Sclerosis (ALS)

ALS, a motor neuron disease and also called Lou Gehrig's disease, was first identified in 1869 by the noted French neurologist Jean-Martin Charcot. ALS affects as many as 20,000 Americans with 5,000 new cases occurring in the United States each year, typically with an onset at an age between 40 and 60. Men’s risk is about 1.5 times to that of women. ALS is usually fatal within 5 years after diagnosis. It is a progressive neurodegenerative disease that attacks motor neurons in the brain and the spinal cord.

The progressive degeneration of the motor neurons in ALS eventually leads to neuron death and loss of related muscle movements. Patients of ALS become partially or totally paralyzed, but most of the patients’ cognitive functions remain unaffected. There is no

45 cure but the Food and Drug Administration (FDA) recently approved riluzole, a drug that

has been shown to prolong the survival of ALS patients.

A key neuropathological hallmark of ALS is intraneuronal aggregates of NFs in

degenerating motor neurons (Carpenter 1968; Averback 1981; Delisle and Carpenter

1984; Manetto et al., 1988; Munoz et al., 1988; Leigh et al., 1989; Murayama et al.,

1992). Although the reason for the NF aggregates is unclear, transgenic mouse models which overexpress any of the NF subunits (NFL, NFM, and NFH) will provoke motor

neuropathy characterized by the presence of abnormal NF accumulations resembling those found in ALS (Cote et al., 1993; Julien et al., 1995; Wong et al., 1995; Xu et al.,

1993). Remarkably, the motor neuropathy in transgenic mice overexpressing human

NF-H was rescued by restoring a correct stoichiometry of NF-L to NF-H subunits with the co-expression of human NF-L transgene (Meier et al., 1999). It has been proposed that such alterations in NF homeostasis are directly relevant to the pathogenesis of ALS

(Williamson et al., 1998).

Interestingly, abnormal NF inclusions are often associated with decreases in levels of NF-L mRNA. NF-L mRNA is selectively reduced by up to 70% in degenerating

neurons of ALS and AD (McLachlan et al., 1988; Bergeron et al., 1994; Wong et al.,

2000; Menzies et al., 2002). Decrease of NF-L mRNA level occurs with decreases of

α-internexin or peripherin mRNA levels, suggesting an absolute alteration of the

stoichiometry of NF expression in ALS (Bergeron et al., 1994; Wong et al., 2000).

Also codon deletions or insertions in the KSP regions of NF-H including a large

deletion of five KSP repeats have been detected in a small number of sporadic cases of

ALS, (Figlewicz et al., 1994; Tomkins et al., 1998; Al- Chalabi et al., 1999).

46 Taken together, these observations have suggested a significant alternation in

NF expression in ALS.

B. Alzheimer’s Disease (AD)

AD is a central nervous system neurodegenerative disease. It was first described

in 1906 by German physician Dr. Alois Alzheimer. Although the disease was once

considered rare, it is now established as the leading cause of dementia. An estimated 4

million individuals in the United States have Alzheimer’s disease with an annual expense

of about $70-$100 billion, and as many as 350,000 individuals develop the disease each

year (Mayeux R., 2003). The risk of Alzheimer’s disease varied from 12% to 19% for

women over the age of 65 years and 6% to 10% for men (Seshadri et al. 1997). On

average, AD patients live about 8 years after they are diagnosed, although the disease can

last for as many as 20 years. The areas of the brain that control memory and thinking

skills are affected first, but as the disease progresses, cells die in other regions of the

brain. Eventually, the person with Alzheimer’s will need complete care. If the individual

has no other serious illness, the loss of brain function itself will cause death. There is no

cure and the only therapy known as cholinesterase inhibitors or anti-cholinesterase drugs including Reminyl (galantamine), Aricept (donepezil hydrochloride) and Exelon

(rivastigmine) provide symptomatic relieves but do little to slow the disease progression.

After the last two decades of intensive studies, more information has been yielded, but the cause of AD is still unknown. Many hypotheses have been proposed, such as amyloid-β cascade (Selkoe, 2000), tauopathy (Lee et al., 2001), inflammation

(McGeer et al., 2001; Weiner et al., 2002), oxidative stress (Markesbery, 1997; Christen,

2000; Picklo et al., 2002, Perry et al., 1998, 2002).

47 Cytoskeleton disruption is a prominent feature and a secondary event followed by

oxidative damage in Alzheimer disease (Smith et al., 1995; Nunomura et al., 2001;

Ahlijanian et al., 2000). NFT as the hallmark of AD is composed of abnormally modified

τ, NFs, and other cytoskeleton proteins. The cause for the aberrant biochemical processes that transform normal τ and NF into abnormal filaments is not understood.

In AD, inappropriate hyperphosphorylation of proteins such as τ, NF is prominent, which is likely due to perturbation in the balance between kinases and phosphatases activity (Gong et al. 2000). MAP kinase (Trojanowski et al., 1993), GSK-3

(Mandelkow et al., 1992), and CDK 5 (Lew et al., 1994; Maccioni et al., 2001) pathways are known to be involved in τ and NF phosphorylation in neurons. One of the earliest changes noted in AD is accumulation of hyperphosphorylated τ and NF in normal perikarya (Sternberger et al., 1985; Manetto et al., 1988; Sobue et al., 1990; Cleveland et al., 2001). This abnormal hyperphosphorylation may cause the protein aggregation in

NFT in AD neurons. The role of NFT and senile plaques in AD are still actively disputed.

C. Parkinson's Disease (PD)

PD is a progressive disorder of the central nervous system. The disease was originally described in 1817 by an English physician, James Parkinson, who called it

"Shaking Palsy." It is affecting more than 1.5 million people in the United States. Men and women are similarly affected. The frequency of the disease is considerably higher in the over-60 age group, and the average duration of illness is about 9 years (Hely et al.

1999).

PD is caused by the degeneration of the pigmented neurons in the substantia nigra of the brain, resulting in decreased dopamine availability. Clinically, the disease is

48 characterized by a decrease in spontaneous movements, gait difficulty, postural instability, rigidity and tremor. Only in the 1960's, however, pathological and biochemical changes in the brain of patients were identified, opening the way to the first effective medication for the disease. Administration of the drug levodopa, a dopamine precursor, has been the standard treatment for Parkinson's disease.

The major pathological change of PD is an accumulated protein inclusion called the Lewy body , composed of numerous proteins include the essential constituent of

α-synuclein protein, the three NF subunits (Galloway et al., 1992), ubiquitin and proteasome subunits (Trimmer et al., 2004). Electron microscopy and biochemical evidence indicates that the abnormally phosphorylated NFs form a non-membrane bounded compacted skein in the neuronal soma in the affected neurons.

In familial PD, mutations in the Parkin gene is the major cause, in which over

20 different mutations have been identified (Abbas et al. 1999, Leroy et al. 1998, Lucking et al. 2000, Lim et al. 2002).

More recently, a point mutation has been reported in the region of the NEFM gene coding for the rod domain 2B of NFM in an individual with Parkinson’s disease

(Lavedan et al., 2002). The base pair change results in substitution of serine for glycine at residue 336, and probably disrupts assembly. The patient developed Parkinson’s disease at the very young age of 16. Although this is an isolated report and three other unaffected family members also carried the mutation, it is possible that this mutation is responsible for the early-age onset of an aggressive form of PD. It is already known that mutations in the same region of neurofilament proteins might cause a peripheral motor nerve axonal loss in NEFL and central dopaminergic loss in NEFM. It has also been found that NF-L

49 mRNA decreased in PD correlating with the severity of the disease (Hill et al., 1993).

D. Charcot-Marie-Tooth Disease (CMT)

CMT is the most common inherited neurological disorder. It was discovered in

1886 by three physicians, Jean-Marie-Charcot, Pierre Marie, and Howard Henry Tooth.

Approximately 150,000 Americans are affected by the condition. CMT affects both

motor neurons and sensory neurons to the muscles, and the patients slowly lose normal

use of their feet/legs and hands/arms as nerves to the extremities degenerate. The loss of

nerve function in the extremities also leads to sensory loss. The ability to distinguish hot

and cold is diminished as well as the sense of touch. There are several types of CMT but

the simplest classification is into types 1, 2, and 3. Types 1 and 3 are due to demyelization, and type 2 is an axonal disease. CMT is generally inherited in an autosomal dominant pattern. At present there is no cure for CMT, although physical therapy and moderate activity are often recommended to maintain muscle strength and endurance.

Several families have now been identified in which heterozygosity for mutations of the NEFL gene on chromosome 8 are associated with CMT2 (Mersiyanova et al., 2000; De Jonghe et al., 2001; Georgiou et al., 2002; Jordanova et al., 2003).

The first of these changes is a substitution of proline at residue 8 with glutamine. The second change is a switch of leucine to a proline residue at position 333 in the highly conserved rod domain 2B.

These NEFL CMT2 mutations disrupt neurofilament assembly and axonal transport, which probably underlies the disease mechanism. A mouse model with a mutation in the same coil, leucine to proline at residue 394, develops severe selective

50 motor neuron death due to disruption of neurofilament assembly (Lee et al., 1994).

Georgiou et al. (2002) identified a novel NF-L missense mutation (Cys64Thr) that caused

the disease in a large Slovenian CMT2 family. This mutation results in a proline to serine

substitution at codon 22 (Pro22Ser) and the mutation showed complete co-segregation

with the dominantly inherited CMT2 phenotype in this family. A similar Pro22Thr was also detected in unrelated Japanese patients with CMT disease (Yoshihara et al., 2002).

Jordanova et al. (2003) found six pathogenic mis-sense mutations and one 3-bp inframe deletion in the NF-L gene in 323 patients with different CMT phenotypes.

Some types of CMT are caused by defects in proteins expressed in the Schwann cell, including the connexin 32 (CX32/GJB1) genes. Such mutations result in aberrant myelination and altered neurofilament phosphorylation. In such cases, therefore, abnormalities of neurofilament phosphorylation occur downstream of the primary pathological process but are still upstream of the end point.

E Diabetes

Diabetes is a disease in which the body does not produce or properly use insulin. The cause of diabetes is still a mystery, although both genetics and environmental factors such as obesity and lack of exercise appear to play roles. There are 18.2 million people in the United States, or 6.3% of the population, develop diabetes. It is estimated that at least 20.1 million Americans have pre-diabetes, in addition to the 18.2 million with diabetes. About 65% of deaths among people with diabetes are due to heart disease and stroke.

Diabetes is associated with a symmetrical distal axonal neuropathy. The neuropathy is predominantly in sensory neurons or dorsal root ganglia. There is an

51 increase in phosphorylation of NFs in lumbar dorsal root ganglia of rats with streptozocin-induced diabetes, but not in other neural cell types (Fernyhough et al.,

1999). This is probably the result of activation of c-Jun N-terminal kinase (JNK), which is a NFs kinase. Pathologically, neurites from sympathetic ganglia are swollen with disorganized aggregates of NFs and other proteins, including peripherin (Schmidt et al.,

1997). Many of these studies show that NFs abnormalities can be the primary cause of the disease and that NFs changes may not simply be a passive marker of pathological processes.

The neuropathy associated with diabetes includes well-documented impairment of axonal transport, a reduction in axon caliber and a reduced capacity for nerve regeneration. All of those aspects of nerve function rely on the integrity of the axonal cytoskeleton. NFPs mRNAs are selectively reduced in the diabetic rat and posttranslational modification of at least one of the NFPs is altered. There is some evidence that altered expression of isoforms of protein kinases may contribute to these changes. Tubulin and actin aberrant modifications are also found (McLean WG., 1997)

The list of diseases involved with NF abnormalities will likely increase and the common mechanism for all the protein aggregation diseases may soon be recognized

(Caughey and Lansbury, 2003).

1.2.12 Summary

In this review, we summarized as many as eleven aspects of NF, which may provide some help for the overall knowledge of the readers. Future studies will be focusing on functions, interactions, and pathology of neurofilament in normal axon development and neurodegenerative diseases, which will provide more knowledge and

52 value for NF. Hopefully, in the recent future, some effective therapeutic method for these neurodegenerative diseases will be available.

53 1.3 Introduction of Tau Protein

1.3.1 General Information

Neurons are cells with a very complex morphology that develop two types of cytoplasmic extensions, axons and dendrites. Neural transmission occurs through these

processes, and therefore, any abnormal changes in neuronal morphology may affect their

normal function and induce pathological events.

Microtubules are very dynamic structures, and in proliferating cells such as neuroblasts (neuron precursors), their probability of assembly is the same as that of depolymerization in all directions. This equilibrium results in the cell maintaining a

sphere-like morphology. However, during the differentiation of a neuroblast into a neuron

(Mitchison et al., 1988), the microtubules become stabilized in specific directions,

thereby generating the cytoplasmic extensions that will become the axon and the dendrites (Mitchison et al., 1988).

Many specific proteins serve to stabilize microtubules, which include the microtubule-associated proteins (or MAPs) MAP1A, MAP1B, MAP2, and tau. An

asymmetric distribution of MAPs (Matus, 1988) is observed in mature neurons (Craig and

Banker, 1994), and tau is preferentially localized in axons (Binder et al., 1985). Present mainly in the axon of a neuron, tau has been related to axonal microtubule function, both

alone and in synergy with other MAPs (Gonzalez-Billaul et al., 2002).

1.3.2 History

Tau protein was discovered almost simultaneously in the United States and

Europe as a protein that lowers the concentration at which tubulin polymerizes into microtubules in the brain (Cleveland et al., 1977a, b; Fellous et al., 1977). At the same

54 time, other high-molecular-weight MAPs were also found to influence the cycles of

microtubule assembly and disassembly in vitro (Shelanski et al., 1973).

1.3.3 Tau Gene and Tau Expression

The tau cDNA was first isolated from a mouse brain expression library (Lee et al., 1988), and subsequently, it was cloned from other species including goat (Nelson et al., 1996), chicken (Yoshida et al., 2002), bovine (Himmler et al., 1989), and human

(Goedert al., 1992). More recently, tau sequences have been described in a number of distinct species (Nelson et al., 1996).

The human tau gene is located on chromosome 17 (Neve et al., 1986), where it occupies over 100 kb and contains at least 16 exons (Andreadis et al., 1992). Right beside

GC-rich 5'-region, a single untranslated exon exists (exon-1) (Andreadis et al., 1996).

Upstream of this exon there are several DNA sequences that contain consensus binding

sites for promiscuous transcription factors such as AP2 or SP1. Tau is mainly expressed in neurons, and an interaction with a neural specific factor has been proposed (Sadot et al., 1996). Nevertheless, the neural specific expression of the protein could also be due to

the presence of possible silencer elements in non-neural cells (Kosik et al., 1997).

In the central nervous system (CNS), alternative splicing of exons 2, 3, and 10

results in the appearance of six tau isoforms. Because exon 10 encodes for one of the regions involved in the binding of tau to microtubules, alternative splicing of exon 10 produces tau isoforms, with either three (tau 3R without exon 10) or four (tau 4R with

exon 10) tubulin/microtubule binding regions. In chicken, an extra tubulin binding region

appears (tau 5R) (Yoshida et al., 2002). In the peripheral nervous system (PNS), there is a

55 high-molecular-weight tau isoform expressing the exon 4A, which yields a protein known

as big tau with an approximate size of 100 kDa (Goedert et al., 1992).

Different tau isoforms are characteristic during brain development. Isoforms lacking exon 10 are found at early developmental stages, or in specific cell types like granular cells of dentate gyrus (Goedert et al., 1989). The determinants of exon 10 splicing have been studied in detail (D’Souza et al., 2002), and this splicing seems to be regulated by the phosphorylation of splicing factors (Hartmann et al., 2001). Other tau isoforms that lack exon 2, or exons 2 and 3 (Himmler et al., 1989),have also been described. Indeed, exons 2, 3, and 10 are alternatively spliced and are adult brain specific.

Exons 2 and 3 are alternatively spliced cassettes; exon 2 can appear alone, but exon 3

never appears independently of exon 2 (Andreadis et al., 1995). Furthermore, in humans there is little (if any) expression of exons 6 and 8.

1.3.4 Tau Protein Structure (See more details in Chapter 4)

1.3.5 Posttranslational Modifications of Tau

Several post-translational modifications have been described for tau protein

including phosphorylation, glycation, glycosylation, oxidation, cross-linking, ubiquitinylation, deamidation, and nitration. The mostly studied one is phosphorylation.

A. Tau Phosphorylation

In the 1980s, tau was found as a phosphoprotein (Ihara et al., 1986), and all these

studies focused on the serine/threonine phosphorylation of the tau protein. Recently, one

study has focused on phosphorylation on its tyrosine residues (Williamson et al., 2002).

(See more details in Chapter 4)

B. Tau Glycation

56 Proteins with a relatively long half-life can be modified at lysine residues by nonenzymatic reactions involving the condensation of a sugar aldehyde or ketone group

with the ε-NH2-groups of the lysines (Maillard et al., 1912). Advanced glycation end products can be formed through irreversible changes of these products that can result in the cross-linking of the modified proteins (Eble et al., 1983). It has already been shown

that Tau isolated from PHF is glycated (Ledesma et al., 1994), and this glycation might be involved in the aggregation of PHF into more complex aggregates (neurofibrillary tangles) (Ledesma et al., 1995). Moreover, glycated tau could generate oxygen free

radicals that are capable of disturbing neuronal function (Smith et al., 1995).

C. Tau Glycosylation

Both N- or O-glycosylation of tau has been reported, with N-glycosylation

occurring in hyperphosphorylated tau (Wang et al., 1996) whereas O-glycosylated

occurring in unmodified tau (Arnold et al., 1996). This relationship between

phosphorylation and O-GlcNAc glycosylation of tau proteins may play a role in the nuclear localization of tau. The transport of tau into the cell nucleus was reduced with a decrease in the incorporation of O-GlcNAc into tau (Lefebvre et al., 2003). The effects of

N-glycosylation on the phosphorylation of tau have also been studied (Liu et al.,

2002a,b,c).

D. Tau Ubiquitinylation

Ubiquitin is a small protein with 76-amino acid and associates with proteins to be degraded in an ATP-dependent manner (Hershko et al., 1998). Ubiquitinylated tau has already been found in aberrant aggregates such as inclusion bodies found in Pick's or

Parkinson's diseases (Mayer et al., 1989) and in some types of paired helical filaments

57 (PHF) found in AD (Morishima et al., 1993). This indicates that these abnormal tau

aggregates areproduces through unsuccessful or unfinished cellular degradation process.

E. Tau Oxidation

The presence of one or two cysteines in the tau isoforms lacking or containing exon 10 has raised the possibility of the formation of dimmers of tau through protein crosslinking via disulfide bond formation. The oxidation of tau could promote the formation of disulfide bond and result in its aberrant aggregation (Schweers et al., 1995).

Recently, in relation to tau oxidation, the formation of dityrosine crosslinkings (Giunta et al., 2002) and tyrosine nitration has been described in a tau molecule (Mailliot et al.,

2002).

F. Other Modifications of Tau

Tau truncation has been defined as the cleavage at the glutamic acid residue 391

(Wischik et al., 1997). This modification could facilitate aberrant tau aggregation

(Wischik et al., 1997). The deamidation of tau at asparagine (or glutamine) residues has

also been described (Watanabe et al., 2000) and it could play a role in tau aggregation

(Montejo de et al., 1986).

It has been proposed that tau may become cross-linked through the enzymatic reaction modulated by transglutaminase (Miller et al., 1995). Moreover, modifications involving a conformational change of tau protein, and that could involve proline cis-trans

isomerization, could also occur (Wedemeyer et al., 2002).

1.3.6 Tau Degradation

In vitro studies have already showed that tau could be degraded by different

proteases, such as lysosomal calpain (Johnson et al., 1989). However,

58 phosphorylated tau is more resistant to proteolysis by calpain degradation than unphosphorylated tau (Wang et al., 1995). Ubiquitination is not necessary for tau degradation through the ATP-dependent proteasome pathway (26S proteasome). There is evidence that tau can be degraded by the 20S proteasome without ubiquitination (David et al., 2002) and monoubiquitinylated tau (or with the addition of fewer than 5 moieties) can

be subjected to lysosomal degradation (Murphey et al., 2002). It is unclear whether polyubiquitinylated tau could be degraded by the ATP-dependent proteasome pathway.

While proteins like β-catenin need to be phosphorylated by GSK3 to be degraded, it looks

like tau protein is the opposite case. In addition, tau cleavage by caspases has been reported recently (Rohn et al., 2002).

1.3.7 Tau Functions

Weingarten et al. (1975) did the pioneer work clearly showing that tau facilitates

tubulin assembly. Later on it was found that tau both stabilizes polymerized microtubules

and nucleates microtubules (Drubin et al., 1986; Bre et al., 1990; Drechsel et al., 1992)

and that tau could also suppress microtubule dynamics (Bre et al., 1990; Panda et al.,

1995).

By depleting tau using antisense oligonucleotides, it was seen that tau is involved

in neurite outgrowth (Caceres et al., 1990, 1991). Additionally, in nonneuronal cells, the expression of tau induced the formation of long cytoplasmic extensions (Knops et al.,

1991) and resulted in microtubule stabilization and bundling (Lee et al., 1992; Knowles et al., 1994). Since the tau binding site on the tubulin molecule overlaps with that for other proteins like the molecular motor kinesin, tau could also influence axonal transport processes (Ebneth et al., 1998, Terwel et al., 2002).

59 The tau-deficient mouse produced by gene targeting has been seen to be quite normal (Harada et al., 1994), and the only identified differencescompared with wild-type are a decrease in numbers of microtubule in small caliber axons (Harada et al., 1994),

muscle weakness, and some behavioral deficits (Ikegami et al., 2000). This may be explained by the compensation for the lack of tau by other proteins, such as MAP1A, which has been found to increase in mice lacking tau (Harada et al., 1994). Consistently,

defects in axonal elongation were found in mice lacking both MAP1B and tau (Takei et

al., 2000).

Little is known about the function of tau as a membrane-associated protein

(Arrasate et al., 1997, Brandt et al., 1995a) and its nuclear localization (Brady et al.,

1995, Greenwood et al., 1995). As tau also binds to many different proteins with diverse

functions, it may therefore play different roles in the wide variety of processes in which

those proteins are involved.

1.3.8 Tau Polymerization

Oxidation could play a role in tau aggregation. Oxidation of cysteine to produce

disulfide cross-linking favors tau self-assembly in tau 3R molecules, where a single

cysteine is present, but not in tau 4R molecules, where the presence of two cysteines may

permit the formation of intramolecular disulfide bonds (Barghorn et al., 2002). In vitro

studies have shown that tau can be assembled through oxidation processes (Fenton's reaction) in the presence of iron (Troncoso 1993). Also, fatty acids like arachidonic acid can induce tau polymerization in vitro (Abraha et al., 2000, Wilson et al., 1997). This

type of polymerization could be related to the possible interaction of tau with plasma membrane components (Arrasate et al., 2000, Brandt et al., 1995). Another lipid

60 peroxidation molecule-HNE has been shown facilitating tau aggregation in vitro and in cells (Perez et al., 2000, 2002). (See more details in Chapter 4)

1.3.9 Animal Models

Transgenic animals expressing genomic human tau, as well as mice carrying cDNAs encoding either the largest or the smallest CNS isoform of human tau have been generated. In addition, transgenic mice carrying tau cDNA with the missense mutations found in FTDP-17 patients have also been obtained. Finally, an additional strategy to

generate animal models of tauopathies has been to overexpress the kinases responsible for tau hyperphosphorylation.

Mice have been generated that overexpress the genomic sequence of human tau, containing the coding sequence, intronic regions, and the regulatory regions of the gene

(Duff et al., 2000). The human tau is distributed in neurites and at synapses but is absent from cell bodies, and no neuropathological lesions were reported in mice of up to 8 months of age.

The polymers assembled from tau only contain the four-repeat isoforms among

majority of the FTDP-17 patients, (Heutink et al., 1989), and in several FTDP-17 families, the only tau mutations found have been those that affect the splicing of exon 10, increasing the ratio of four-repeat with respect to three-repeat isoforms (Heutink et al.,

1989). Taking these observations into account, it was suggested that the four-repeat forms of tau may favor fiber formation, and thus mice have been generated to test this hypothesis that carry cDNAs encoding either the largest or the smallest CNS isoform of human tau (Gotz et al., 1995). On the other hand, several transgenic animals

overexpressing the shortest isoform have been obtained but using different transgene

61 promoters. With the murine 3-hydroxy-methyl-glutaryl CoA reductase promoter (Brion et

al., 1999), hyperphosphorylated somatodendritic transgenic tau was detected although

NFTs did not appear to form in these animals. The level of expression of the same tau

isoform was increased by using the murine PrP promoter (Ishihara et al., 1999).

Transgenic lines with high levels of overexpression were not viable while lines with less

than 10-fold overexpression of tau protein developed inclusions in cortical and brain stem

neurons. These inclusions were most abundant in spinal cord neurons and were correlated with axon degeneration, diminished microtubules, reduced axonal transport in ventral roots, spinal cord gliosis, and motor weakness. NFT-like inclusions (detected by histochemistry using dyes such as Congo red and Thioflavin S) were detected in the same

transgenic mice at 18–20 mo of age (Ishihara et al., 2001). In addition, filaments were isolated from detergent-insoluble tau fractions (Ishihara et al., 2001).

The first transgenic mice carring cDNAs encoding the largest human brain tau

isoform and showing low levels (10%) of overexpression (Gotz et al., 1995) has been studied. Transgenic human tau protein was present in nerve cell bodies, axons, and dendrites. Tau was phosphorylated at sites that are hyperphosphorylated in PHFs, although filaments were not detected. The murine Thy 1.2 promoter has been used by two groups to drive the expression of the cDNA encoding the longest human brain tau isoform, although on different genetic backgrounds. In the first case, axonal degeneration

in the brain and spinal cord was detected in the mice (Spittaels et al., 1999). Axonal

dilation with accumulation of neurofilaments, mitochondria, and vesicles was also observed. The axonopathy and the accompanying dysfunction of the animals’ sensory

motor capacities were transgene-dosage dependent. The mice generated by the second

62 group (Probst et al., 2000) contained numerous abnormal, tau-immunoreactive nerve cell bodies and dendrites. In addition, large numbers of pathologically enlarged axons containing neurofilament and tau-immunoreactive spheroids were present, especially in

the spinal cord.

The third approach to generate animal models of tauopathies is based on that

mutations in tau are associated with FTDP-17 (Clark et al., 1998, Goedert et al., 1998,

Grover et al., 1999). Recently two groups have reported the generation of transgenic mice

expressing mutant human tau containing the P301L mutation (Gotz J et al., 2001, Lewis

et al., 2000), which is located in the tubulin-binding domain and reduces the affinity of tau for microtubules. These animals developed NFTs mainly in the spinal cord, and

similar to the previous transgenic models developing some neuropathological symptoms

encouragingly reminiscent of the human disease. In the second transgenic model published, short filaments of tau could be isolated from the brains of the transgenic mice

(Gotz J et al., 2001).

More recently, another transgenic mouse model containing the P301L mutation

has been characterized (Sahara et al., 2002), as well as one containing the P301S mutation

(Allen et al., 2002). In the latter model, abundant tau filaments within a PHF structures were observed. Interestingly, in the transgenic mouse expressing the tau P301L mutation, tau filaments were only observed in old female mice but not in their male counterparts.

This could be due to a decrease in the amount of tau in male mice (Bu et al., 2002).

Filaments were also found in transgenic mice expressing the mutation R406W of human

tau (Tatebayashi et al., 2002), a mutation that decreases the phosphorylation of tau at the

site recognized by Ab PHF-1 (Perez et al., 2000). Finally, it has been shown that

63 filaments are formed in transgenic mice expressing mutant (V337M) human tau

(Tanemura et al., 2002).

In summary, a significant body of data demonstrates that a large excess of normal

or mutated human tau can produce some of the cellular changes observed in tauopathies, but this may not be sufficient for the formation of the mature neurofibrillary aggregates

observed in the human diseases. Nevertheless, some transgenic lines do develop NFT-like

structures as well as short PHF-like filaments. What these approaches seem to indicate is

that, first, tau is linked to neurodegeneration, second, that neurons with long axons, such as those present in spinal cord, seem to be especially susceptible to alterations in tau. We have recently produced a transgenic mouse expressing three FTDP-17 missense mutations in tau: G272V, P301L, and R406W (Lim et al., 2001). Ultrastructural analysis of mutant tau-positive neurons revealed a pre-tangle appearance; filaments of tau were found as well as increased numbers of lysosomes displaying aberrant morphology similar to those found in AD (Lim et al., 2001).

1.3.10 Tau and Neurodegenerative Diseases (Tauopathies)

Tauopathies are considered as a group of disorders resulting from abnormal tau

phosphorylation, abnormal levels of tau, abnormal tau splicing, or mutations in the tau

gene. In some tauopathies, like AD or Down's syndrome, the tau pathology is associated with other cerebral changes.

A. AD

AD is the most common and the best-studied tauopathy. The disease results in widespread atrophy in the brain that begins in the temporal and parietal lobes. It leads to

problems in short-term memory, a function associated to the temporal lobe, visual and

64 spatial dysfunction, and poor performance of over-learned tasks related to the functions of

the parietal lobe (Wilhelmsen et al., 1999). Tau aggregates from AD are composed of the

six CNS tau isoforms in their phosphorylated form. The six tau isoforms are as follows:

1) those containing exons 2, 3, and 10, plus all the constitutive exons (Baker et al., 1999);

2) those having exons 2 and 3; 3) those containing exons 2 and 10; 4) those having only

exon 2; 5) those with only exon 10; and 6) only containing the constitutive exons. This combination of hyperphosphorylated tau isoforms results in the appearance of three major electrophoretic bands with a mobility corresponding to that of proteins with a relative molecular weight of 68,000, 64,000, and 60,000 (Delacourte et al., 1999; Greenberg et al., 1992; Lee et al., 2001). Furthermore, AD brains contain higher quantities of tau compared with unaffected controls (Khatoon et al., 1992, 1994).

The best-known tauopathy is AD, in which two main pathological structures are formed in the brains of patients: senile plaques (composed of the β-amyloid peptide), and neurofibrillary tangles (NFTs). These NFTs are made up of by paired PHF comprising of hyperphosphorylated tau (Grundke-Iqbal et al., 1986, Ihara et al., 1986). The number of

NFT has been correlated with the degree of dementia in this disease (Arriagada et al.,

1992), and thus there is great interest to study how these structures are formed. The formation of PHF from tau molecules may involve different steps and could include tau phosphorylation (although this may not be essential), a conformational change in the

protein, and finally polymerization. If indeed phosphorylation does facilitate tau assembly

into PHF, it will be of interest to know which kinases (and phosphatases) contribute to the phosphorylation of tau molecules to be reached.

B. Corticobasal Degeneration

65 Corticobasal degeneration is a disorder characterized by cognitive disturbances

like aphasia and apraxia, moderate dementia, and motor disturbances such as rigidity,

limb dystonia, and tremor. Pathological analyses have indicated frontoparietal atrophy

and glial and neuronal tau inclusions. Tau is also present in hyperphosphorylated form,

but only the 68,000- and 64,000-molecular weight forms can be detected by

electrophoresis.

C. Downs' Syndrome

Downs' syndrome is due to the trisomy of chromosome 21 that results in the

defective growth and maturation of the brain, producing a cognitive impairment and

dementia at 50 years of age. In this disorder, tau is also hyperphosphorylated yielding a

pattern similar to that of AD.

D. Frontotemporal Dementia With Parkinsonism Linked to Chromosome 17

In this disorder, the patient displays frontotemporal atrophy, with neural loss, gliosis, and cortical spongiform changes in the lobes. Those result in behavioral changes, language deficit, and hyperorality. Tau inclusions were observed in neuron and glia cells, and two types of hyperphosphorylated tau form were detected. In some cases, the pattern is similar to that of AD, with the 68,000-, 64,000-, and 60,000-molecular weight forms being detected, while in other cases the pattern is like CBD and only the 68,000- and

64,000-molecular weight forms were found. This is due to intronic mutations that result in the forced expression of exon 10 (Grover et al., 1999; Hutton et al., 1998; Spillantiniet al., 1998; Varani et al., 1999). On the other hand, mutations resulting in the deletion of lysine-280 lead to reduced splicing of exon 10 (Rizzu et al., 1999), while the G342V

mutation may affect the splicing of exons 2 and 3 (Lippa et al., 2000). Some factors

66 involved in the alternative splicing of tau have been already identified (D’Souza et al.,

20002), and more than 25 mutations in tau have been identified in exons 1, 9, 10, 11, 12,

and 13. These mutations may effect RNA splicing or the protein levels of tau. These tau

(missense) mutations preferentially occur in the microtubule binding region (exon 10), decreasing binding of the mutated protein to microtubules (Hasegawa et al., 1996, Hong et al., 1998) or affecting the binding of tau to other proteins that bind to that region of tau

(Goedert et al., 2000). Furthermore, tau self-assembly was increased in most of these

mutated tau proteins (Arrasate et al., 1999).

E. Pick's Disease

Pick's disease is a dementia that produces disturbances in language and behavior

and is associated with frontal lobe atrophy. It provokes changes in the character of the

patient and in their relationships with others, as well as depression. This disorder is characterized by the presence of cytoplasmic tau inclusions in neurons of the frontal lobe, known as Pick bodies. The granular cells of the dentate gyrus are also affected. The appearance of inclusions is in agreement with the appearance of 64,000- and

60,000-molecular weight hyperphosphorylated tau, indicating the absence of exon 10 expression. This exon is not expressed in tau present in the granular cells of the dentate gyrus (Goedert et al., 1989), suggesting that the degeneration of selective neuronal

populations in different tauopathies reflects the physiological pattern of tau isoforms expressed.

F. Postencephalic Parkinsonism

This disorder appears to be the consequence of an encephalic parkinsonism found

in patients that survived the Spanish influenza pandemia (Geddes et al., 1993). Different

67 brain regions are affected, but tau inclusions are mainly found in the hippocampus and the

putamen. The electrophoretic pattern for hyperphosphorylated tau is similar to that of AD.

G. Progressive Supranuclear Palsy

Progressive supranuclear palsy (PSP), also known as the Steele, Richardson, and

Olszewski disorder, is characterized by supranuclear gaze palsy, prominent postural instability, and dementia in the later stages. Tau inclusions have been found in neuronal and glial cells, with both astrocytes (tufted astrocytes) and oligodendrocytes (coiled bodies) being affected. Recently, as in FTDP-17, missense mutations have been found in

PSP patients with a monogenic (familial) origin (Delisle et al., 1999), and in fact to date,

mutations in Tau have been identified in PSP (J. G. De Yebenes, personal communication), CBD (Bugiani et al., 1999), and Pick's disease (Murrell et al., 1999).

Some specific tau polymorphisms may be a risk factor for PSP (De Yebenes et al., 1999), and these polymorphisms are based on the different numbers of repeats of a dinucleotide repeat (TG) present in the intron between exons 9 and 10 (see above). Individuals containing 11 dinucleotide repeats (tau allele A0) have been seen to have a greater risk of developing PSP.

H. Niemann-Pick Type C Disease

Niemann-Pick type C disease is an autosomal recessive lysosomal lipid storage disorder, mainly caused by mutations within the NPC-1 gene (Vanier et al., 1996). The clinical features are motor disturbances (due to cerebellar dysfunction) and dementia

(Vanier et al., 1991). The pattern of hyperphosphorylated tau is similar to that of AD.

I. Other Tauopathies

68 Other tauopathies involving hyperphosphorylated tau include parkinsonism with dementia, myotonic dystrophy, prion diseases with tangles, Blint disease (an ophthalmologic disorder), dementia pugilistica, dementia with tangles only or amyotrophic lateral sclerosis, and parkinsonism-dementia complex of Guam. These have been well-documented in the reviews of Buee et al. (2000) and in that of Ingram and

Spillantini (Ingram et al., 2002).

1.3.11 Summary

Tau is a microtubule-associated protein that is not essential for mammalian development, probably due to functional redundancy and the presence of other microtubule-associated proteins. However, in pathological situations tau may be

hyperphosphorylated and assembled in an aberrant way. Because of these modifications,

neural toxicity are augmented, resulting in the appearance of neurological disorders, mainly dementias, which are collectively known as tauopathies. Different types of neurons are damaged in different tauopathies, and the elucidation of the mechanisms

underlying the basis for this specificity will probably be the goal of several laboratories

studying in neurodegeneration.

69 1.4 Introduction of Oxidative Modification

1.4.1 Free Radical Theory of Aging

Aging is the seemingly inevitable decline in physiologic function that occurs over time. For all living organisms, the ultimate ending of aging is the same: death. At least four major theories of aging have been proposed to explain most or all of the cause of biological aging:

A. The free radical theory of aging (FRTA) (Harman, 1956);

B. The mitochondrial theory of aging (Harman,1972);

C. The membrane hypothesis of aging (Zs-Nagy, 1978);

D. The cross-link theory of aging (Kleinsek, 2002).

Dr. Denham Harman, the establisher of the free radical theory of aging, has defined aging as the increased probability of death with the increase of the age of an organism increases, and the accumulation of diverse adverse physiologic changes. The

FRTA was first proposed by Dr. Harman almost 50 years ago in November 1954

(Harman, 1992), and his groundbreaking original paper on the FRTA was published in

1956 (Harman, 1956). Free radicals, the highly reactive, small oxygen-containing molecules, play key roles in the mitochondrial, membrane and cross-link theories, as well as in the FRTA.

1.4.2 Free Radicals

Free radical is simply a molecule carrying an impaired electron. All free radicals are extremely reactive and ready to acquire an electron in any way possible. In the process of acquiring an electron, the free radical will attach itself to another molecule, thereby modifying it biochemically (Bradford and Allen, 1997). However, as free radicals

70 acquired an electron from the other molecules, they convert these molecules into free radicals, or break down or alter their chemical structure. Thus, free radicals are capable of damaging virtually any biomolecule, including proteins, sugars, fatty acids and nucleic acids (Leibovitz and Siegel, 1980). Harman points out that free radical damage occurs to long-lived biomolecules such as collagen, elastin and DNA; mucopolysaccharides; lipids that make up the membranes of cells and organelles such as mitochondria and lysosomes; components of blood vessel walls; and proteins and lipids that combine and accumulate as large pigment- lipofuscin (Harman, 1984).

The main radicals are superoxide radical, hydroxyl radical, hydroperoxyl radical, alkoxyl radical (AR), peroxyl radical (PR) and nitric oxide radical (NOR) (Jacob, 1995).

Other molecules that are technically not free radicals, but act much like them, are singlet oxygen, hydrogen peroxide (H2O2), and hypochlorous acid (HOCl) (Jacob, 1995).

Collectively, the free radicals and non-free radical mimics are called oxidants or reactive oxygen species.

Free radicals are extremely short-lived because of their extreme reactivity. The longest half-life is 1-10 seconds for nitric oxide radical; the shortest (and most deleterious) is only one nanosecond (10 -9 sec) for hydroxyl radical (Jacob, 1995). A broad range of the chief diseases of aging, including cancer, heart/artery disease, essential hypertension, Alzheimer’s disease, aging immune deficiency, cataracts, diabetes,

Parkinson’s disease, arthritis and inflammatory disease, as well as aging itself, are now believed to be caused in whole or part through free radical damage (Harman, 1984;

Jacob, 1995; Ames et al., 1993; Pierrefiche and Laborit, 1995).

1.4.3 The Primary Sources of Free Radicals

71 There are four primary sources of oxidants formed within living organisms. The major source of free radicals and oxidants is the mitochondrial generation of ATP energy using oxygen. A small percentage (2-3% or less) of oxygen in mitochondria is inadvertently converted to superoxide radical, which can in turn generate hydrogen peroxide, hydroxyl radical, and all other free radicals (Beckman and Ames, 1998, Ames et al., 1993; Pierrefiche and Laborit, 1995). A second source of oxidants, especially hydrogen peroxide, is the peroxisomes, organelles that degrade fatty acids (Beckman and

Ames, 1998; Ames et al., 1993). A third source of oxidants is cytochrome P450 .

These enzymes help cells, especially in the lungs and liver, detoxify a broad range of potentially toxic food, drug and environmental pollutant molecules. Superoxide radical is a by-product of many of these detoxification reactions (Beckman and Ames, 1998; Ames et al., 1993).

Finally, white blood cells (phagocytes) attack germs with a mixture of oxidants including superoxide radical, hydrogen peroxide, nitric oxide radical, HOCl, and hydroxyl radical (Beckman and Ames, 1998; Leibovitz and Siegel, 1980; Ames et al.,1993). This may create serious free radical problems, especially in those suffering a chronic immune-activation condition, such as AIDS, chronic candidiasis, protozoal infections, chronic fatigue syndrome, etc. (Beckman and Ames, 1998 ; Ames et al.,

1993). In addition, various biomolecules including hydroquinones, flavins, catecholamines, thiols, pterins and hemoglobin, may spontaneously auto-oxidize and produce superoxide radical (Leibovitz and Siegel, 1980).

From outside the body, polluted urban air, cigarette smoke, iron and copper salts, some phenolic compounds found in many plant foods, and various drugs may all

72 contribute free radicals or provoke free radical reactions (Leibovitz and Siegel, 1980;

Ames et al.,1993).

1.4.4 Physiological Damages Caused by Free Radicals

Free radicals and ROS lead to several damaging effects as they can attack lipids, protein/

enzymes, carbohydrates, and DNA in cells and tissues. They induce undesirable

oxidation, causing membrane damage, protein modification, DNA damage, and cell death

induced by DNA fragmentation and lipid peroxidation. This oxidative damage/stress,

associated with ROS is believed to be involved not only in the toxicity of xenobiotics but

also in the pathophysiological role in aging of skin and several diseases like heart disease

(atherosclerosis), cataract, cognitive dysfunction, cancer (neoplastic diseases), diabetic

retinopathy, critical illness such as sepsis and adult/acute respiratory distress syndrome,

shock, chronic inflammatory diseases, organ dysfunction, deep injuries, production of

nitric oxide by the vascular endotheliums, vascular damage caused by ischaemia

reperfusion (Halliwell et al, 1984). Iron changes have been detected in multiple sclerosis,

spastic paraplegia, and amyotrophic lateral sclerosis, which reinforces the belief that iron

accumulation is a secondary change associated with neurodegeneration in these diseases,

although it could also be related to gliosis (glia might produce free radicals) in the

diseased area, or the changes in the integrity of the blood brain barrier caused by altered vascularisation of tissue or by inflammatory events (Gutteridge et al., 1986).

1.4.5 Antioxidant Systems

Endogenous Antioxidants

Biological systems have evolved with endogenous defense mechanisms to help

protect against cell damage induced by free radicals. Glutathione peroxidase, catalase,

73 and superoxide dismutases (SOD) are antioxidant enzymes, which metabolize toxic

oxidative intermediates. They require micronutrient as cofactors such as selenium, iron,

copper, , and manganese for optimum catalytic activity and effective antioxidant

defense mechanisms (Halliwell et al., 1992; 2001). SOD, catalase, and glutathione

peroxidase are three primary enzymes, involved in direct elimination of active oxygen

species (hydroxyl radical, superoxide radical, and hydrogen peroxide). In addition,

glutathione reductase, glucose-6-phosphate dehydrogenase, and cytosolic GST are

secondary enzymes, which help in the detoxification of ROS by decreasing peroxide

levels or maintaining a steady supply of metabolic intermediates like glutathione and

NADPH necessary for optimum functioning of the primary antioxidant enzymes

(Vendemialeet al., 1999; Singh et al., 2003).

Glutathione, ascorbic acid, alpha-tocopherol, betacarotene, bilirubin, selenium,

NADPH, butylhydroxyanisole (BHA), mannitol, benzoate, histidine peptide, the

iron-bonding transferrin, dihydrolipoic acid, reduced CoQ10, melatonin, , and

plasma protein thiol, etc., as a whole play a homoeostatic or protective role against ROS produced during normal cellular metabolism and after active oxidation insult. Glutathione is the most significant component that directly quenches ROS such as lipid peroxides

(like HNE) and plays major role in xenobiotic metabolism. When an individual is exposed to high levels of xenobiotics, more glutathione is utilized for conjugation making it less available to serve as an antioxidant. It also maintains ascorbate (vitamin C) and alpha-tocopherol (vitamin E), in their reduced form, which also exert an antioxidant effect by quenching free radicals (Meister et al., 1994; Sies et al., 1995; Anderson et al.,

1996).

74 Exogenous Antioxidants: Contribution From the Diet

The most widely studied dietary antioxidants are vitamin C, vitamin E, and

beta-carotene. Vitamin C is considered the most important water-soluble antioxidant in

extracellular fluids, as it is capable of neutralizing ROS in the aqueous phase before lipid

peroxidation is initiated. Vitamin E is a major lipid-soluble antioxidant, and is the most

effective chain-breaking antioxidant within the cell membrane where it protects

membrane fatty acids from lipid peroxidation. Beta-carotene and other carotenoids also

provide antioxidant protection to lipid rich tissues. Fruits and vegetables are the major

sources of vitamin C and carotenoids, while whole grains, i.e., cereals and high quality

vegetable oils are the major sources of vitamin E (Block et al., 1992; Halliwell et al.,

1994; Jacobet al., 1995).

A number of other dietary antioxidants exist beyond the traditional vitamins

collectively known as phytonutrients or phytochemicals, which are being increasingly

appreciated for their antioxidant activity. One example is flavonoids, which are a group

of polyphenolic compounds. These are widely found in plants as glucosylated derivatives.

They are responsible for the different brilliant shades such as blue, scarlet, and orange.

They are found in leaves, flowers, fruits, seeds, nuts, grains, spices, different medicinal

plants, and beverages such as wine, tea, and beer (Harborne et al., 1994; Weisburger et al., 1997; Pietta et al., 2000; Gale et al., 2001).

Flavonoids exhibit several biological effects such as antitumoral, anti-ischaemic, anti-allergic, anti-hepatotoxic, antiulcerative, and anti-inflammatory activities. These are also known to inhibit the activities of several enzymes, including lipoxygenase, cyclooxygenase, monooxygenase, oxidase, glutathione-S-transferase,

75 mitochondrial succino-oxidase, NADH oxidase, phospholipase A2, and protein kinases.

Many of the biological activities of flavonoids are attributed to their antioxidant properties and free radical scavenging capabilities. The antioxidant activities of flavonoids vary considerably depending upon the different backbone structures and functional groups. A number of flavonoids efficiently chelate trace metals, which play an important role in oxygen metabolism. Free iron is a potential enhancer of ROS formation as it leads to reduction of H2O2 and generation of the highly aggressive hydroxyl radical.

Free copper mediates LDL oxidation and contributes to oxidative damage due to lipid peroxidation (Shu et al., 1998). Due to the inefficiency of our endogenous defense systems as well as the existence of some physiopathological situations, such as cigarette smoke, air pollutants, UV radiation, inflammation, ischaemia/reperfusion, etc., ROS can be produced in excess, and increasing amounts of dietary antioxidants will be needed for diminishing the cumulative effect of oxidative damage over an individual’s life span (Sun et al., 2002).

1.4.6 Oxidative Stress in The Nervous System

The oxidative stress is a shift towards the pro-oxidant in the pro-oxidant/antioxidant balance that can occur as a result of an increase in oxidative metabolism. Its increase at the cellular level can come as a consequence of several factors, including exposure to alcohol, aging, medications, trauma, infections, toxins, radiation, strenuous physical activity, and poor diet. Defense against all of these processes is dependent upon the adequacy of various antioxidants that are derived either directly or indirectly from the diet (Gotz et al., 1994).

The nervous system, including the brain, spinal cord, and peripheral nerves, is

76 rich in both unsaturated fatty acids and iron. The high lipid content of nervous tissue, coupled with its high aerobic metabolic activity, makes it particularly susceptible to oxidative damage. The high level of iron may be essential, particularly during brain development, but its presence also means that injury to brain cells may cause the release of iron ions, which leads to oxidative stress via the iron-catalyzed formation of ROS

(Bauer et al., 1999; Andorn et al., 1990). In addition, those brain regions that are rich in catecholamines are exceptionally vulnerable to free radical generation. The catecholamine adrenaline, noradrenaline, and dopamine can spontaneously break down

(auto-oxidise), or can be metabolized to radicals by the endogenous enzymes such as

MAO (monoamine oxidases) to form free radicals. One such region of the brain is the substantia nigra (SN), where a connection has been established between antioxidant depletion (including GSH) and tissue degeneration (Perry et al., 2002). A number of in vitro studies have shown that antioxidants, both endogenous and dietary, can protect nervous tissue from damage by oxidative stress (Contestabile et al., 2001). Uric acid, an endogenous antioxidant, was found to prevent neuronal damage in rats, both in vitro and in vivo, from the metabolic stresses of ischemia, oxidative stress as well as exposure to the excitatory amino acid glutamate and the toxic compound, cyanide. Vitamin E was found to prevent cell death (apoptosis) in rat neurons subjected to hypoxia followed by oxygen reperfusion (Yu et al., 1998). In the same study, it was shown that vitamin E prevented neuronal damage from reactive nitrogen species. Both vitamin E and beta-carotene were found to protect rat neurons against oxidative stress from exposure to ethanol (Copp et al., 1999). In an experimental model of diabetes-caused neurovascular dysfunction, beta-carotene was found to protect cells most effectively, followed by

77 vitamin E and vitamin C (Mitchell et al., 1999).

Most in vivo and clinical studies of the effects of lipid-soluble antioxidant

supplementation on neurological diseases have focused on vitamin E. A report in 1991

demonstrated that the rate at which Parkinson’s disease progressed to the point when the

patient required treatment with levodopa was slowed by 2.5 years in patients given large

doses of vitamin C and synthetic vitamin E (Prasad et al., 1999). Although one study

reported that high doses of vitamin E resulted in elevated plasma levels but failed to

increase vitamin E levels in cerebrospinal fluid (CSF), a later report demonstrated that

high doses of vitamin E did result in elevation of CSF vitamin E levels, and possibly

brain vitamin E levels (Pappert et al., 1996). Recently it was shown that the protein

responsible for the uptake of vitamin E is in fact present in brain cells of patients

suffering from vitamin E deficiency or diseases associated with oxidative stress (Jama et

al., 1996). In a Dutch study, it was found that the risk for Parkinson’s disease was lower for subjects who had higher dietary intakes of antioxidants, particularly vitamin E. The same group reported that a low dietary intake of beta-carotene was associated with impaired cognitive function in a group of persons aged 55-95; no such association was observed for either vitamins C or E (Hellenbrand et al., 1996). In an Austrian study, serum concentration of vitamin E was found to be significantly associated with cognitive function in adults aged 50 - 75 years measured by a standardized test (Schmidt et al.,

1998). In another study, it was found that patients suffering from Parkinson’s disease had consumed less of the small-molecule antioxidants betacarotene and vitamin C than did non-sufferers of the disease, implying that dietary antioxidants do play a protective role in this disease (Fahn et al., 1991). About 20% of FALS cases are associated with a

78 mutation in the gene for copper/ zinc superoxide dismutase, an important antioxidant

enzyme, and in vitro experiments demonstrated that expression of the mutant enzyme in neuronal cells caused cell death, which could be prevented by antioxidant small molecules such as glutathione and vitamin E (Ferrante et al., 1997).

There is substantial evidence that oxidative stress is a causative or at least ancillary factor in the pathogenesis of major neurodegenerative diseases, including

Parkinson’s disease, Alzheimer’s disease, and amyotrophic lateral sclerosis (ALS, “Lou

Gehrig’s disease”) as well as in cases of stroke, trauma, and seizures (Ghadge et al.,

1997). Decreased levels of antioxidant enzyme activity have been reported in patients with Parkinson’s disease (Spina et al., 1989). Evidence of increase in lipid peroxidation and oxidation of DNA and proteins has indeed been seen in the substantia nigra of patients affected with Parkinson’s disease. Similar increase in markers of oxidative stress has also been seen in Alzheimer’s disease, Huntington’s disease, and in both familial

ALS (FALS) and sporadic ALS (SALS) patients (Saggu et al., 1989). It has also been reported that there are elevated lipid peroxide levels in the cerebrospinal fluid of patients maintained on neuroleptics and exhibiting symptoms of TD (Berger et al., 1997).

Weisburger et al. (1997) succeeded in decreasing the severity of TD using high doses of vitamin E, and called for further trials with combinations of various known antioxidants.

Schizophrenia (SCZ) is also believed to have a component of free-radical overload. Lipid peroxides have been found elevated in their blood and increased pentane gas, a marker for lipid peroxidation, in the breath of schizophrenics as compared with normal volunteers and with patients having other psychiatric illness (Phillips et al., 1993). The enzyme SOD was found increased, possibly as an adaptive response to free radical

79 overload in the SCZ patients (Abdalla et al., 1986). Studies of antioxidant treatment in schizophrenia have been very few; two recent studies that examined only vitamin C yielded conflicting results. As GSH peroxidase level was also found to be reduced in

SCZ patients, future trials with antioxidants in schizophrenia should include selenium and

GSH precursor nutrients (Levine et al., 1985).

1.4.7 HNE and HNE Modification in AD

In the past decade HNE (trans-4-hydroxy-2-nonenal, C9H16O2; Fig. 1.6) has been the most well-studied product of endogenous lipid peroxidation occurring in oxidative stress (Esterbauer et al., 1991). The x-6-family (linoleic and arachidonic acids) of polyunsaturated fatty acids may produce HNE as a result of free radical attack. HNE is a highly reactive electrophile and can form adducts with several nucleophilic groups of biomacromolecules. (See more details in Chapter 2, 3, 4)

1.4.8 Summary

Neurodegenerative disorders remain an important source of morbidity and suffering of the humankind. The role of free-radical-mediated oxidative injury in acute insults to the nervous system including stroke or trauma, as well as in chronic neurodegenerative disorders, is just being recognized. As we know, oxygen is an essential molecule for the survival of majority of living organisms. Oxidative stress is the harmful condition that occurs when there is an excess of free radicals and/or a decrease in antioxidant levels. There is evidence suggesting that the increase in energy metabolism via aerobic pathways enhances the intracellular concentration of free oxygen radicals, which in turn enhance the rate of the autocatalytic process of lipid peroxidation, inducing damage to brain structures, especially when physiological defenses become insufficient

80 or depleted. Antioxidants combat oxidative stress by neutralizing excess free radicals and

stopping them from setting off the chain reactions that contribute to various diseases and

premature aging. The evidence to date for oxidative stress in PD, TD, SCZ, AD and other

neurodegenerative diseases is strongly convincing. Clinical studies show that a number of

events associated with AD are capable of stimulating production of free radicals and depletion of antioxidant levels. Patients with Parkinson’s also have reduced glutathione levels and free radical damage is found in the form of increased lipid peroxidation and oxidation of DNA bases.

Tackling of the role of free radicals bring up a novel therapeutic target in those diseases. However, only when the mechanisms and involvement of free radicals in the pathogenesis of neurodegeneration as well as neurological or CNS diseases are understood, will antioxidant therapy be designed effectively and specifically targeted.

As pointed out, whether oxidative stress is eventually proven to be primary or secondary during the etiologic progression, the therapeutic rewards of antioxidants are likely to be substantial. The essential consideration is to deliver the proper scavenger to the affected site within the time frame of or prior to maximal tissue damage. Clearly, strategies aimed at limiting free radical production during oxidative stress and damage may slow the progression of neurodegenerative diseases. How can one combat the production of free radicals? Patients can replete their cellular and body stores with the body’s most important antioxidant,

81

Figure 1.6 The structure of HNE and its adducts with amino acids. A The structure of HNE and the formation of its Michael adducts with amino acids lysine histindine and cysteine. B The formation of 2 2-pentylpyrrole and fluorescent crosslink lysine adducts formed by HNE.

82 glutathione and its synergistic partner antioxidants, maintain their antioxidant defense

system, thus help to prevent or delay the progression of free radical related damages.

There are several natural and synthetic compounds already selected and extensively

studied. Preclinical experimental studies have proven that they can reduce acute

neurological disorders by preventing lipid peroxidation and diminishing free radical

generation. In fact, they appear to be highly beneficial in experimental models. Clinical

trials are therefore warranted with dietary GSH precursors, administered in combination

with other antioxidants, antioxidant cofactors, and non-antioxidant brain-trophic nutrients

such as phosphatidylserine.

1.5 Research Relevance

In the following chapters, we examined HNE modification of neurofilament

protein and Tau protein. In chapter 2, we found that neurofilament is a scavenger for

HNE and this adduction is a phosphorylation-regulated and sequence-specific modification may preserve very important function for neuron survival. In chapter 3, we further explore the protective function of neurofilament protein in cell culture. In chapter

4, we found that HNE not only is involved in NFT formation, but it also plays a major role in the formation and epitope production of NFT.

These works add significant new evidence for the critical role of Tau and neurofilament proteins in NFT formation in AD. In addition, these studies also strongly support the fact that HNE is a major player in oxidative modification and AD pathogenesis.

83

Chapter 2

Oxidative Modification of Neurofilament

Proteins in Alzheimer Disease

84 Introduction

HNE and HNE Modification in AD

In the past decade HNE (trans-4-hydroxy-2-nonenal, C9H16O2; see Figure 1.6 A in Chapter 1) has been the most well studied product of endogenous lipid peroxidation occurring in oxidative stress (Esterbauer et al., 1991). The ε-6-family of polyunsaturated fatty acids (linoleic acid and arachidonic acids) may produce HNE as a result of free radical attack. HNE is a highly reactive electrophile and can form adducts on several nucleophilic groups on biomacromolecules. HNE reacts with cysteine, histidine, and lysine residues of proteins to form Michael adducts that can be stabilized by cyclization to hemiacetals (see Figure 1.6 A in Chapter 1). Michael adducts of cysteine and histidine are irreversible and the lysine Michael adducts are formed reversibly (Nadkarni and

Sayre, 1995). HNE also forms Schiff base adducts with lysine, leading to a

2-pentylpyrrole (Sayre et al., 1993, 1996) (see Figure 1.6 B in Chapter 1) and other advanced lipoxidation end-product (ALE) adducts through condensative and oxidative evolution of initially- formed adducts (Sayre et al., 2001). The bifunctionality of HNE permits it to form a Schiff base Michael adduct crosslink between two protein-based lysines. Recently, a stable fluorescent HNE-derived lysine–lysine crosslink structure has been identified that contributes to the fluorescence seen in aged and/or diseased tissues

(Xu et al., 1999) (see Figure 1.6 B in Chapter 1). HNE-modified proteins may display altered biological functions due to conformational changes, hydrophobicity changes, and crosslink resulting from the adduction. An array of antibodies to specific HNE adducts are available to visualize the HNE adducts.

In last decade, HNE modification has been used intensively as a marker for the

85 studies related with lipid peroxidation. HNE modification has already been reported to be

significantly increased in AD brain comparing to controls (Sayre et al, 1997). Therefore, whether and how HNE is involving in the AD pathogenesis is an interesting topic that needs to be further explored.

Neurofilament Properties Related with This Study

Assembly

NF assembly is not clearly understood yet, though in vivo studies have demonstrated that NFs are obligate heteropolymers requiring NF-L to form a proper polymer with either NF-M or NF-H (Ching et al., 1991; Lee et al., 1993). NF formation is

believed to start with the dimerization of NF-L with either NF-M or NF-H subunits. The

highly conserved rod domains of the NF subunits are coiled together in a head-to-tail

fashion to form a dimer. Two coiled dimers overlap with each other in an anti-parallel,

half-staggered manner forming the tetramers. Finally, eight tetramers are packed laterally

and longitudinally together in a helical array, forming a ropelike 10 nm filament (see

Figure1.3 in Chapter 1) (Hein et al., 1994; Fuchs et al., 1998; Herrmann et al., 2000) In

the cross section of NF, there are ~32 molecules though the number may change in

different stages or conditions (Herrmann et al., 1999). The C-termini of NF-M and NF-H

are not in the coils but they form the side-arms of the NFs (Perrone et al., 2001).

These assembly characteristics of neurofilament make the C-termini of NF-H and

NF-M exposed in the cytoplasm and available for the post-assembly modifications, such

as oxidative modifications. However, this may be difficult or impossible for the

N-termini of all three subunits.

Transport

86 After their synthesis in the perikarya, neurofilament proteins (NFPs) are quickly translocated into the axons and assembled into filamentous structures. In radioisotopic labeling studies, neurofilament proteins move at a rate of 0.2–1 mm/day in the axon with the slow components whereas other organelles such as membrane vesicles can move at much faster rate of 200–400 mm/day (Okabe et al., 1990).

The mechanism by which NF proteins are transported has been a topic of debate for many years due to the lack of techniques in determining the form in which these proteins are transported. Polymer hypothesis suggests that NF subunits are assembled and transported mainly as polymeric structures (Roy et al., 2000; Wang et al., 2000), whereas the subunit hypothesis argues that the NF proteins are transported down the axon as individual subunits or small oligomers (Prahlad et al., 2000; Yabe et al., 1999; Hsieh et al., 1994; de Waegh et al., 1992). The photobleaching experiments with fluorescent-tagged proteins support the subunit hypothesis, in which the bleached axonal segment remained stationary with gradual recovery of its fluorescence without any obvious directionality (Cole et al., 1994). Recently, using GFP-tagged NF subunits and live cell imaging, two research groups have found a fast transport of NF polymers down the axon, at a rate up to 2 mm/s (Jaffe et al., 1998; Sternberger and Sternberger, 1983), which was interrupted by prolonged pauses and, resulted in a net slow velocity.

Therefore, the current understanding of NF transport is that they are transported bi-directionally in the axon along microtubules through common motors like dynein and kinesin. The average slow rate of NFs movement is because NF structures spend most of their time (>99%) pausing in the axon. The pauses might be the result of the lack of the transient interactions between motors and NF structures. Consistent with this model the

87 kinesin motor was found to co-localize with the motile NF pool in the squid giant axon

(Prahladet al., 2000). In addition, phosphorylation of NFPs (mainly on NF-H) resulted in

their dissociation from the kinesin motor, thereby decreasing their transport speed (Yabe

et al., 1999).

In the human sciatic nerves (~1m long), neurofilament may take 1-3 years to

reach the synapses. In this over long transport process, how the neurofilaments maintain

their physiological function is an interesting question.

Degradation

Many neurons extend their axons over great distances, up to 1 meter in human

sciatic nerves, to form synapses with appropriate receptor cells. To maintain the

physiological functions of the nerves, certain proteins need to have long lifetimes to span

the axon. NFPs are among them. During the transport process, degradation of NFPs is not

detected. Pulse-labeling experiments indicate that the cytoskeletal proteins are

metabolically stable during their passage through the axon, and essentially all of these proteins complete the trip to the axon terminus (Hoffman and Lasek 1975; Lasek and

Black 1977). So it is believed that the degradation of NFPs may only happen at synapses

(Lasek and Hoffman 1976; Roots, 1983), where dephosphorylation of NFs by the protein phosphatase 2A (PP2A) precedes NF degradation. This also indicates phosphorylation has a protective function for NFPs from degradation (Gong et al., 2003). At the axonal terminus, the half residence time of NF is around 2 days (Paggi et al., 1987). The disintegration of NFs during pathological conditions is accounted for by the presence of calcium-activated protease (calpains) in the axoplasm (Pant and Gainer, 1980; Buki et al.,

1999). Calpains can degrade many different proteins including important axonal

88 cytoskeletal proteins such as spectrin, microtubule-associated protein-1, tau, tubulin and

NFs, as well as several protein constituents of myelin (e.g. myelin basic protein, myelin

associated glycoprotein, proteolipid protein) (Kampfl et al., 1997). The breakdown and turnover of NFs could be accounted for by the activation of this protease in nerve termini.

It is possible that the recycling of NFs degradation products represent an important feedback mechanism regulating NFs production. The success or failure of axonal regeneration may be determined by the intrinsic neuronal factors that control the production of the axonal cytoskeleton (Buki et al., 1999).

In addition, less than 1% of total NFs synthesis was also detected in axons (Grant and Pant, 2000; Sotelo-Silveira et al., 2000), and the role of this small amount of newly synthesized NFs is unknown.

Summarizing the neurofilament transportation and degradation process, there is a very intriguing question is that how these proteins keep their integrity in such a long time and how can they deal with the intensive modifications in the cytoplasm. In this study, we provide some new findings for these questions.

Phosphorylation of Neurofilament Proteins

NFPs phosphorylation is topographically regulated within neurons. Recent studies show that phosphorylation of the NF subunits play a critical role in the regulation of the filament translocation, formation, and functions. It is also involved in the pathogenesis of some related neurodegenerative diseases.

Almost all (>99%) of the assembled NFs in myelinated internodal regions are known to be stoichiometrically phosphorylated in the Lys-Ser-Pro (KSP) repeat domains

(Hsieh et al., 1994). In contrast, the KSP repeats of NFPs in cell bodies, dendrites, and

89 nodes of Ranvier are less phosphorylated. The unphosphorylated NFs only account for

~1% of NFs in the neuron as judged by NF density and the relative volumes of the

respective compartments (de Waegh et al., 1992; Cole et al., 1994).

Although other Ser/Thr motifs are also phosphorylated, most phosphorylation

sites are in KSP motifs of the tail domains of NF-M and NF-H. It is consistent with the

studies done by Jaffe et al (1998) showing that 33 out of 38 phosphorylation sites in the

C-terminus of NFH are in KSP motifs.

Normally NF-M and NF-H undergo the post-translational hyperphosphorylation

upon entry into the axons. In vivo phosphorylation of NFPs in axons is a slow process

(Komiya et al., 1986). The explanation of this includes the slow transport of NF proteins

into the axon, delayed activation of local kinases, or more possibly the inaccessibility of

phosphate acceptor sites because of a “closed” conformation of the assembled polymer.

The topologically sequential phosphorylation may be due to the sequential opening of the

next accessible phosphorylation site after the pre-phosphorylation of the other sites (see

Figure 1.4 in Chapter 1).

NFPs phosphorylation plays multiple roles, such as axonal diameter, NF axonal

localization and assembly, C-terminal phosphorylation of NFPs, and NF transport.

The major function of the phosphorylation of NFPs is believed to be the regulation of

axonal radial growth. Extensive phosphorylation of KSP repeats in the tail domain of

NF-M and NF-H occurs primarily in axons (Sternberger and Sternberger 1983; Lee et al.,

1988; Nixon et al., 1992). This results in sidearm formation, increased

inter-neurofilament spacing, radial growth of axons, and increased conduction velocity

(de Waegh et al, 1992; Nakagawa et al., 1995; Yin et al., 1998). Phosphorylation of KSP

90 motifs triggered by a Schwann cell signal has been proposed as a key determinant of local control of NF accumulation, inter-filament spacing and radial growth of myelinated

axons (de Waegh et al., 1992; Nixon et al., 1994; Sanchez et al., 1996; 2000; Yin et al.,

1998). Recent studies using NF-H null mice, or mice transfected with a NF-H tailless mutant, suggest that neither NF-H nor its phosphorylated tail is essential for determining

neuronal caliber (Rao et al., 1998; Rao et al., 2003). These studies, together with those

using NF-M knock-out mice (Elder et al., 1999a,b; Kriz et al., 2000), indicate that NF-M

plays a more profound role in regulating the axonal diameter. Moreover, alternation of

NF-M and NF-H tail domain phosphorylation is associated with the pathology seen in

neurodegenerative disorders such as ALS and AD, in which tail domain phosphorylation and NF accumulation occur abnormally in perikarya (Sternberger et al., 1985; Manetto et al., 1988; Sobue et al., 1990; Cleveland et al., 2001).

It has been proposed that the heavily phosphorylated NF-H subunit detaches from

the transport carrier and resides in the axon for months, whereas less phosphorylated

subunits are transported at normal slow axonal transport rates and have a short residence time in the axon (Lewis et al., 1987; Nixon et al., 1991). Consistent with this hypothesis,

the studies of complete deletion of NF-H, NF-M, or both increase the rate of transport of the remaining neurofilament subunits in mouse sciatic nerves (Zhu et al., 1998)

Moreover, in cultured optic nerve axons, hypophosphorylated NF subunits have been interpreted to undergo axonal transport more rapidly than subunits more extensively

phosphorylated at their tail domains [90], and the C-terminal phosphorylation of the

NF-H subunit correlates with decreased NF axonal transport velocity (Jung et al., 2001;

Yabe et al., 2001). But Rao et al. (2002) constructed NF-H tailless mice by embryonic

91 stem cell mediated gene knock in approach and showed the NF-H tail is neither important for the axonal diameter or the transport rate. Ackerley et al. (2003) argued with these results by analyzing movement of GFP-tagged phosphorylation mutants of NF-H in neurons, which support a role for NF-H side arm phosphorylation as a regulator of NF transport. Most of the studies support phosphorylation of NF side arms regulating NF transport rates (Ackerley et al., 2000; Shea and Pant et al., 2003).

Research Relevance

As one of the most striking and earliest changes noted with the development of

Alzheimer’s disease (AD), the levels of oxidatively modified macromolecules are found increased in the brain of AD patients. These including oxidative modification of protein

(Good et al., 1996; Smith et al., 1996, 1997, 1998), nucleic acid (Nunomura et al., 1999), carbohydrate (Dong et al., 1993; Smith et al., 1994), and the adducts of lipid peroxidation product, trans-4-hydroxy-2-nonenal (HNE) (Montine et al., 1996; Sayre et al., 1997).

HNE adducts increased in neuronal cell bodies of AD cases and also presented in axons in the white matter of infants and aged controls (Smith et al., 1995; Sayre et al., 1997), which leads the question which specific axonal proteins are physiological targets of HNE modification.

In our previous studies, neurofilament heavy subunit (NFH), and to a less extent, neurofilament medium subunit (NFM), were identified to be the major targets of HNE adduction in axons (Wataya et al., 2002). This is consistent with the high lysine content in NFH and NFM, but unexpected results from studies of sciatic nerves of different animals and human brains showed that the levels of NFH-HNE adducts kept constant

92 along the axons and during the lifetime (Wataya et al., 2002). It is our great interest to find out what is (are) the mechanism(s) controlling the NFH-HNE level.

As a very unique character, NFH and NFM are almost the only proteins have multiple KSP repeats in their C-termini and more importantly these sites are also the targets of phosphorylation (Hsieh et al., 1990; de Waegh et al., 1992; Cole et al., 1994

Jaffe et al., 1998) and HNE modification (Wataya et al., 2002). In addition, we found that dephosphorylation can reduce the HNE antibody recognition, which let us to hypothesize that phosphorylation may control at least partially the level of NFH-HNE.

In this study, for the first time with in vitro analysis, we demonstrated that phosphorylation in the KSP repeats regulates the level of NFH-HNE, one type of oxidative modification on protein. It directly shows that NFH, as a cytoskeleton protein, can be a scavenger for at least one of the free radicals. With the studies of the neurofilament enriched sample, C-terminus of NFH, synthetic peptides, we show that

KSP repeats in NFH (maybe NFM) are the key motifs which used by neurons to regulate the level of HNE adducts through phosphorylation and dephosphorylation on serine residues in NFH. This is the first study that shows oxidative modification on protein can be a regulated cellular process through phosphorylation in neurons.

93 Materials and Methods

Antibodies:

The following antibodies were used throughout this study: (1) HNE-Michael antibody, rabbit antiserum to the products of HNE adduction to keyhole limpet hemocyanin

involving mainly Michael adducts of cysteine, histidine, and lysine (Uchida K et al.,

1993); (2) mouse monoclonal antibody to phosphorylated NFH (SMI-34; Sternberger

Monoclonals), (3) mouse monoclonal antibody to nonphosphorylated NFH (SMI-32;

Sternberger Monoclonals). The specificity of HNE-Michael antibody is confirmed with

these three methods. a) The genesis of the antibody is using HNE-KLH as immunogen. b)

This antibody can recognize HNE-Lys, HNE-His, HNE-Cys compounds and the

immunoreactivity is not diminished by immunoabsorption with the other two compounds.

c) Mass spectrometry data indicates that non-phosphorylated KSP peptides do not form

detectable HNE adducts. Specificity of the various antibodies was verified in all cases

by immunoabsorption for 16 h at 4 °C with the competing antigens or to adducts prepared from the reaction of HNE with lysine, cysteine, or histidine at a 1:1 concentration of

0.8 mM. Reaction of the latter two amino acids with HNE gives the corresponding

Michael adduct nearly stoichiometrically, whereas the reaction of HNE with lysine leads to a time-dependent mixture of mainly the reversibly formed Michael adduct but also

several additional adducts derived from initial Schiff base formation at the C1 carbonyl group of HNE (Nadkarni and Sayre ,1995). TG-4 recognizes phospho-serine 235 of Tau

(Vincent et al., 1996; Weaver et al., 2000) and MC15 recognizes phospho-Tau (Vincent

et al., 1996).

Enriched human neurofilament protein sample:

94 Enriched neurofilament protein fractions were obtained from frozen human tissue using the method of Shecket and Lasek (Shecket G et al., 1980).

Synthetic peptides:

The 20 amino acid-Peptides containing KSP repeats from both human and mouse

NFH and a random sequence were generous gift from Dr. Mervyn Monteiro (University

of Maryland, MA). Phospho/nonphospho-AKSPV and phospho/nonphospho-KSP were

synthesized from company (GeneMed Inc, CA). Phospho/nonphospho-KS were

synthesized in Dr. Sayre’s laboratory in CWRU. All the peptides are diluted to 2 mM in

PBS, TBS or water for the further studies.

HNE/dephosphorylation treatment and immunoblots:

HNE was synthesized (Nadkarni and Sayre, 1995) and initially dissolved in ethanol. Further dilutions were made in PBS, TBS, or water. NFH protein and peptide samples were incubated with 2mM HNE 37 ℃, 16 hrs.

For protein, ~20 µg/well of NFH protein were separated on 10% SDS PAGE and electrotransferred to Immobilon membrane (Millipore).

For peptides, 2 µl (2mM) peptide samples were applied on each dot on Immobilon membrane (Millipore), air dry, and rinse the membrane with methanol before usage.

Some samples were dephosphorylated (1 unit alkaline phosphatase type III

(Sigma)/50 µg protein or peptides in 0.1 M Tris, pH 8.0, with 0.01 M

phenylmethylsulfonyl fluoride for 16 h at room temperature or with 25% hydrofluoric

acid for 16 h at 4 °C after HNE treatment).

After all the pre-treatment, Membranes were blocked with 10% milk 1 hr, washed

with Tris-buffered saline-Tween 5X5 min, incubated with primary antibody 16 hrs 4 ℃,

95 after five washes of 5 min in Tris-buffered saline-Tween, peroxidase-labeled secondary antibodies were applied for 1 h at room temperature. After rinsing again as described before, the blots were developed together for the same length of time using ECL reagent

(Santa Cruz Biotechnology) and imaged by using image acquisition and analysis software

(Ultra-Violet Products Ltd.). The effective concentration of HNE and time for the

reaction were already determined in a previous study of tau protein (Takada et al., 2000).

CNBr cleavage:

cyanogen bromide (CNBr) cleaves the C-terminal of Methionine on human NFH

and gives a 135 Ka C-terminal fragment of NFH. It was performed as the method

described in literatures (Richard C et al., 1985; Elhanany E et al., 1994). 0.5 milligrams

of human NFH was incubated in the dark, at room temperature, and under argon in 50 ul of 70% formic acid containing 1 mg of CNBr (Sigma) for 3 hours. The reaction was stopped by dilution to 0.25 ml with water and dried by Speed Vacuum. The dry pellet was suspended in 0.25 ml of water and immediately separated by gel filtration with full-length NFH sample.

Chemical stabilization of HNE adducts:

KS Peptides were incubated with 2mM or 10mM HNE, pH 7.4, 37 0C in different

experiments. In the different time points, 100µl incubated samples was taken out and

quenched by 40µl, 100mM NaBH4 for 2 hours, then 20µl saturated NH4Cl was added for

1 hour. The final mixture was diluted 5 times and injected to HPLC for further analysis.

KSP repeats search in protein databases:

The website we used is http://us.expasy.org/tools/scanprosite/. We searched in

Swiss-Prot and TrEMBL protein databases for human proteins that have more than 10

96 KSP repeats. Only three proteins, NFH, NFM and human ovarian cancer related tumor marker CA125, were found .

Analysis of HNE modified AKSPV by mass spectrometry:

Both Matrix-assisted Laser Desorption/Ionization mass spectrometry (MALDI) and high performance liquid chromatography-electrospray ionization-mass spectrometry/mass spectrometry (HPLC-ESI-MS/MS) techniques were applied in the experiments (Landry F. et al., 2000, Aebersold R, 2003). In MALDI-MS, 2 ul samples incubated as above were mixed with 2 ul matrix (a-cyanine, 3% TFA). And 2 ul mixed samples were put onto MS plate and dry in air. Program for small peptides were selected for the optimum detection. MS scan was performed at both positive and negative mode, at reflection mode, attention 60, 150 shots.

In HPLC-ESI-MS/MS, peptide samples were incubated as above and separated by

HPLC on a microbore column (C18aq, 5 BioResources) at 50 mL/min with a 0% acetonitrile containing 0.1% formic acid in first 10 minutes then gradient of 0 to 95% acetonitrile containing 0.1% formic acid for 60 min. These elutes were monitored by the

UV absorbance at 214 nm, and were analyzed by on-line ESI-MS. ESI-MS was performed with an ion trap LCQ mass spectrometer (LCQDUO, Finnigan). The mass spectrometer was equipped with an ESI needle with an ion spray voltage set at 3000 V and with nitrogen as the sheath gas. The capillary temperature and sheath gas pressure were set at 160 °C and 80 psi, respectively. MS scan (m/z 80-2000) was performed in the positive ion acquisition mode to detect intact molecular ions. The peptide AKSPV was detected as [M + H] + at m/z 501. The peptide phospho-AKSPV was detected as [M + H]

+ at m/z 581. The peptide HNE modified phospho-AKSPV was detected as [M + H] + at

97 m/z 737. The identification of the peptide AKSPV (m/z 501), the peptide AKSPV (m/z

581), and the peptide HNE modified phospho-AKSPV (m/z 737) were further confirmed

by collision-induced dissociation using the tandem MS/MS scan.

Data Analysis:

All immunoblots were processed following the same time and dilution parameters

to allow direct comparisons between various samples. Blots from all samples were

stained concurrently for each antibody and for exactly the same time of development. The

intensity of immunostained NFH bands was quantitated as reflectivity with an optically enhanced densitometry scanner (QS30; pdi, Inc.). Statistical analysis was performed with analysis of variance Fisher's post hoc protected least significant difference test.

98 Results

Enriched neurofilament protein samples were purified from normal human brain homogenize and electrophoresed then transferred on the Immobilon membranes. The membranes were incubated with 2mM HNE for 16 hours at 37 ℃. HNE adducts were detected as a major band around 205 kDa and dephosphorylation with alkaline phosphatase (AP) or hydrofluoric acid (HF) can reduce the adduct (Figure 2.1). When replace AP with BSA for the incubation, there is no reduction was detected (Figure 2.1).

Tau protein, also a lysine-rich and hyperphosphorylated protein, formed HNE adducts but did not show reduction of the adducts after dephosphorylation with either AP or HF (Figure 2.2).

Since the phosphorylation and probably HNE modification mostly happened in the C-terminus of NF-H, a 135 kDa C-terminal fragment of NF-H from amino acid 394 was obtained by cyanogen bromide (CNBr) cleavage which cleaves on the C-terminal side of Methionine at 9 more sites in the N-terminus of NF-H (Figure 2.3A). Treatment with HNE showed that HNE adducts were readily formed on the full length NF-H and the

C-terminus of NF-H at the similar level (Figure 2.3B) and dephosphorylation will reduce both of them (data not shown).

Because the phosphorylation sites on NF-H are mainly in the KSP repeats in the

C-terminus (Hsieh et al., 1990; de Waegh et al., 1992; Cole et al., 1994 Jaffe et al., 1998) and lysine residues are the major amino acid modified by HNE (Wataya et al.,

99

Figure 2.1 Phosphorylation state of NFH regulates level of HNE-NFH adduct.

NFH formed HNE-Michael adduct after HNE treatment and the adduct are reduced by dephosphorylation of NFH. Replacing AP by BSA, the sample did not show any effect.

AP: alkaline phosphatase.

100

Figure 2.2 Tau protein did not show the similar effect. Tau protein is a lysine rich and hyperphosphorylated protein. Tau formed HNE-Michael adduct after HNE treatment and the adduct is not reduced by dephosphorylation. HF: hydrofluoric acid, AP: alkaline phosphatases.

101

Figure 2.3 C-terminus of NFH is very reactive with HNE. A. CNBr cleavages the

C-terminal of Methanine and gives a 135 kDa C-terminal fragment of NFH that includes all the

KSP repeats. B Comparing with the full length NFH, the C-terminal fragment is also very reactive with HNE treatment.

102 2002), we hypothesized that KSP repeats maybe the most important motifs in NF-H that form the HNE adducts. To demonstrate this question, several steps were taken sequentially. First, three 20 amino acid peptides were obtained. Two of them are consensus sequences from human NF-H and mouse NF-H with three KSP repeats without phosphate groups, one of them is a random peptide without lysine. These peptides were treated with HNE. Because these peptides did not have phosphate groups, they did not show a strong modification after HNE treatment (Figure 2.4). Comparing with them, the

5 amino acid synthetic peptides—phospho-AKSPV, the most repeated KSP sequence in

NFH, showed strong HNE modifications after HNE treatment (Figure 2.5).

Nonphospho-AKSPV peptide did not form or form undetectable level of HNE adducts, whereas the phospho-AKSPV forms higher level of HNE adducts and these adducts are reduced by dephosphorylation (Figure 2.5). These data give us the most direct evidence showing that KSP repeat may play the key role in regulating HNE modification on NF-H.

To determine the minimal sequence requirement for this regulation, KSP peptides with and without phosphate groups were also synthesized. After HNE treatment and dephosphorylation, phospho-KSP peptide showed significant increase and decrease in

HNE adduct level and still preserves the regulatory properties. In the meaning time nonphospho-KSP failed to show significant changes in HNE adduct level (Figure 2.6).

Two shorter peptides, the KS peptides with or without phosphate group also were synthesized. both of these two formed very low level of HNE adduct and did not

103

Figure 2.4 Nonphospho-20 AA KSP peptides are not able to be intensively modified by HNE. 20 AA peptides are concensus sequences from human and mouse

NFH protein including 3 KSP repeats. These peptides can not be modified by HNE.

104

Figure 2.5 Phospho-AKSPV peptide is modified by HNE. Synthetic AKSPV peptides with and without phosphate group showed significant difference in forming

HNE adduct. Dephosphorylation phospho-AKSPV-HNE sample reduce the level of HNE adduct.

105

Figure 2.6 Synthetic K-S-P peptides with and without phosphation show different reactivity with HNE. Phospho-KSP peptide shows HNE-Michael adduct and this adduct is reduced by dephosphorylation. Nonphospho-KSP peptide does not show the intensive formation of HNE adduct.

106 show difference in level of HNE adduct at all (Figure 2.7). These results indicates that

flanking sequence like proline is required for preserving the phosphorylation dependent

regulation of HNE modification on NFH.

To confirm these results, the matrix assisted laser desorption/ionization (MALDI)

technique and high performance liquid chromatography-electrospray ionization-mass

spectrometry/mass spectrometry (HPLC-ESI-MS/MS) were used to study most of these

KSP peptides. Phospho-AKSPV and phospho-KSP all form higher and detectable level of

HNE-Michael adduct comparing to nonphospho-peptides and all the product peaks were

confirmed by fragmentation of the peptide peaks with MS/MS (Figure 2.8A,B,C). Almost

no detectable level of adducts were found with HNE treated nonphospho-peptides (Figure

2.8D). These data further confirm the conclusion that phosphor-peptide is easier to form

HNE adduct than non-phospho-peptides and dephosphorylation will remove not only the

phosphate groups but also HNE modification from these phosphorylated synthetic KSP peptides. The HNE adducts are not removed in control samples replacing AP with BSA

(data not shown).

HNE-Michael antibody is a polyclonal antibody that was used to detect HNE adducts in immunostainning analyses, which is generated from HNE treated Keyhole

Limpet Hemocyanin (KLH) as immunogen (Uchida et al., 1993). The specificity of the antibody was confirmed by using HNE-lysine, HNE-histidine, HNE-cysteine compounds

(Sweda et al., 1997). This polyclonal antibody recognized all three compounds with

different epitopes because the immunoreactivity of this antibody to

107

Figure 2.7 Synthetic K-S peptides with and without phosphorylation did not show significant difference in the formation of HNE adduct. Both these K-S peptides form very low level of HNE adduct and phosphor-KS peptide did not show higher level of HNE adduct formation.

108

Figure 2.8 HPLC-ESI-MS/MS confirms the removal of HNE from the HNE adduct after dephosphorylation of the sample. A. Phospho-AKSPV is detected in molecular weight around 581, and the peptide is confirmed by fragmentation. B

Phospho-AKSPV forms HNE-Michael adduct at molecular weight around 737, and this peak is confirmed by fragmentation. C HNE is removed from the Phospho-AKSPV-HNE adduct after dephosphorylation with alkaline phosphatases. Only nonphospho-AKSPV is detected as final product. D Nonphospho-AKSPV is detected in molecular weight around

501, and this peak is confirmed by fragmentation of the peak Nonphospho-AKSPV does not form HNE adduct.

109 each compound was not blocked by pre-absorption with the other two compounds (data not shown).

The phosphorylation states of neurofilament proteins are monitored by the antibody that recognizing phosphorylated neurofilament (SMI-34) and nonphosphorylated neurofilament (SMI-32).

The phosphorylation states of Tau proteins are monitored by the antibodies that recognizing phosphorylated Tau (TG-4 and MC15). TG-4 recognizes phosphoserine 235

(Vincent et al., 1996; Weaver et al., 2000) and MC15 recognizes phospho-Tau (Vincent et al., 1996).

Experiments were performed both in PBS, TBS (pH 7.4), or Millipored water, there were no significant difference in the formation and reduction of HNE adducts.

110 Discussion

This is the first study shows that NFH-HNE level is a phosphorylation-dependent and sequence-specific modification that is controlled directly by protein phosphorylation/dephosphorylation in the key motif-KSP sequence, which are extensively and uniquely repeated in the C-termini of NFH/M. With the study of the full length NFH, C-terminus of NFH, and different synthetic KSP peptides with and without phosphate group, we finally confirm that the regulative mechanism is preserved in the

KSP motifs. This finding gives a good explanation to the question why NFH-HNE level keep constant in axons during lifetime in mice, rats and human. It also may provide more evidence for the long time mystery why evolutionarily the multiple KSP repeats reside in

NFH/M uniquely. In addition, this finding may provide an explanation for those two

NFH transgenic mice studies mentioned above (Couillard-Després et al., 1998; Beaulieu et al., 2003).

In our studies with the phospho/nonphospho-KS peptides, we have not found the similar properties like NFH/M, the phospho/nonphospho-AKSPV peptides, and phospho/nonphospho-KSP peptides. Therefore, KSP, the three amino acid motif, is clearly the minimal sequence that preserves the phosphorylation-dependent regulation for

NFH/M-HNE adducts.

Even the glutathione (GSH) is considered the major scavenger for HNE in vivo, it

may only account for 30- 50% of total HNE (Siems et al., 1997 a, b; Forman et al., 2003)

and there are up to 10% of total HNE adducts to proteins (Siems et al., 1997 a, b). HNE modifications can cause dysfunction of the macromolecules and eventually insults in cell death (Nakashima et al., 2003). Therefore, the neurons as the most vulnerable cells in the

111 body must have certain defense systems to protect the cells from death.

HNE as a lipid peroxidation product from peroxidation of ω 6 polyunsaturated fatty acids (like arachidonic acid and linoleic acid) was consider the most toxic compounds in cellular oxidative stress (Uchida et al., 1993; Liu et al., 2003). HNE can be existed in a relatively longer time and diffused cellularly to the closed organelles or extracellularly to other cells to adduct with all the macromolecules (Reviewed in Eckl P.

M., 2003; Nakashima et al., 2004; Alary et al., 2004). The concentration of natural induced HNE can be high enough to induce apoptosis (Shackelford et al., 2000; Haynes et al., 2001; Anuradha et al., 2001; Mark et al., 1997; Choudhary et al., 2002; Cheng et al.,

2001; Lee et al., 2004). So the neurons as the most vulnerable cells in the body must have the defense system to protect the cells from death. Even the GSH is considered the major scavenger for HNE in vivo, it may only account for 50% of HNE (Forman et al., 2003) and it still has the possibility to have free HNE time to time (Uchida et al., 1993). If the

HNE adducts are accumulated in the cells, this will causes dysfunction of the macromolecules and eventually insults in cell death.

HNE reacts with proteins can form HNE-Michael adduct, HNE-pyrrole adduct, and protein crosslink (Liu et al., 2003). HNE-Michael adducts account for ~80% of HNE adducts and only Lysine-HNE-Michael adduct is reversible (Montine et al., 1996; Sayre et al., 1997). In neurofilament protein, the lysine enriched protein, lysine is the major amino acid that adducted by HNE, which provides the possibility for neurons to remove the adducts through cellular regulation. In Tau protein, lysine residues are not likely the major amino acid adducted by HNE, because Tau contains certain amount of cysteine and histidine residues, which will form irreversible HNE-Michael adducts easier than lysine

112 residues, so even Tau is phosphorylated and has two KSP in the protein, it did not show

the same regulation (Table 2.1). Only peptides with KSP and phosphorylation on serine

residue preserved the phosphorylation regulated HNE adduction. In most conditions, the

non-phospho-KSP peptides and peptides with lysine did not even show detectable level

of HNE adduct compared with phospho-KSP peptides (Table 2.1), which also suggests

that phosphorylation somehow promotes HNE adduction.

It has already been shown that HNE can induce the activation of MAPK,

PI3K/AKT pathways, which are serine/threonine kinases (Liu et al., 1999; Dozza et al.,

2004). This is reasonable link to the phosphorylation level of cellular proteins. HNE

induce GSH biosynthesis (Forman et al., 2003) which account for the clearance of around

50% of the free HNE. In concerning of the rest of the free HNE, we give a possible

clearance mechanism in neuronal cells.

NFH, with the multiple lysine residues, is highly susceptible to HNE adduction.

The high phosphorylation level in C-terminus, which is resided mainly in KSP repeats

(Jaffe et al., 1998), also made this protein interesting. As phosphorylation has already

been proposed and proven to play roles in interfilamental distance, axonal caliber,

transportation rate, and filament assembly, phosphorylation and translocation controlling

(Nakagawa et al., 1995, de Waegh et al., 1992, Yin et al., 1998,Lewis and Nixon 1988,

Ching et al., 1999,Zheng et al., 2003,Gibb et al., 1998, reviewed in Liu et al., 2004).

Moreover, here we proposed and proved another important and interesting function of

phosphorylation on NFH, which is to regulate the level of oxidative modification on NFH

by HNE. This provides the evidences that KSP repeats involved in a more important

113 cellular protective system. Evolutionally, as the most abundant protein in the most

essential cells, NFH could be applied some more important functions.

Protein database searches for small sequence repeats

(http://us.expasy.org/tools/scanprosite/), revealed that only NFH, NFM and ovarian

cancer related tumor marker CA125 proteins contain more than 10 KSP repeats (Table

2.2). Human ovarian cancer related tumor marker CA125 has ~30 KSP in a 22,152 amino

acid sequence. This protein is a serum antigen used as a marker for the screening of the

ovarian cancer and other tumors (Bast et al., 1983; Kato et al., 2004). Therefore, as far as

we know, human NFH is the only protein has 43/44 KSP repeats concentrated within

around 300 amino acids protein sequence in human protein database.

As we tested around 10 other peptides (Table 2.3), no other sequence showed

similar regulation to that obtained by phospho-KSP motif, suggesting a dual requirement

of KSP repeats and phosphorylation.

In axons, the possibility of proteolytic removal of the HNE adducts is not likely to occur, because degradation of neurofilament is only detected in the area close to synapses

(Lasek and Hoffman 1976; Roots, 1983). Our preliminary studies with the hypertonic

infusion of protease inhibitor leupeptin treatment of exposed sciatic nerve, did not show altered level of NFH-HNE adducts. However, we do not know whether leupeptin is the

appropriate inhibitor for the putative proteolytic activity or whether the infusion was an effective means of delivery (Perry’s lab unpublished data).

In this study, we proposed another possibility that phosphorylation may regulate the level of NFH-HNE and may provide protection for neurons from oxidative insults. This is supported by the studies showed that expressing human NFH can rescue vulnerable

114 neurons from death in two transgenic mice studies (Couillard-Després et al., 1998;

Beaulieu et al., 2003). More consistently, neuronal cells treated with HNE showed time

dependent increase of both level of NFH and phosphorylation level of NFH, which

accommodates the oxidative insults by HNE (Liu and Perry unpublished data). In

addition, cells overexpressing NFH showed significant protection from HNE cytotoxicity

in different neuronal cell lines (Liu and Perry unpublished data). These are also supported

by the recent studies about HNE functions, which showed that there is a narrow range for

HNE to promote cell proliferation (Cheng et al., 1999). Thus it appears that cells need to deliberately adjust the HNE level from time to time (Awasthi et al., 2003).

In this study, we explored the new function of these KSP repeats in human NFH, except for the functions from serine phosphorylation, KSP repeats in human NFH also acts as a scavenger for reactive oxygen species. Based on these new results we proposed the hypothetical model (Figure 2.9). When HNE level is increased in axons, neurons will response by tending to increase the phosphorylation level of NFH especially in KSP repeats, which can adduct more HNE reversibly; and when HNE levels decrease, neurons might signal the opposite response to trigger dephosphorylation, and HNE is released from NFH, then HNE is detoxified in the neurons. This function has never been shown before and it may be vitally important for neurons’ survival.

Even though we provide evidenceof the importantce of KSP repeats using synthetic peptides in this study, we still consider that the secondary (and tertiary) structure in which the KSP sequence lies to be important because there are 1-5 amino acids between most KSP repeats in NF-H and NFM. The role of these neighbouring sequences needs further investigation.

115

Table 2.1 Comparison of all the proteins and peptides in these studies.

116

Table 2.2 Protein database search for KSP repeats (>10 KSP in homosapiens)

117

Table 2.3 All the control peptides used in these studies did not show high level of

HNE adduct.

118 In summary, we demonstrate that KSP repeats in human NFH are the key motifs in modulating HNE-adduct levels on NFH, through regulation by phosphorylation and dephosphorylation on serine residues. This also indicates that NFH-HNE adduction is a regulated event in neurons through activation of kinases and phosphatases, which is directly regulated by signal transduction in neurons. This will be a major discovery that opens the door to the studies of regulation of oxidative modifications in cells and we proved that phosphorylation is directly involved in these regulations in this study.

119

Figure 2.9 Hypothetical model for the phosphorylation-regulated HNE modification in KSP motif.

120

Chapter 3

Cellular Protective Function of

Neurofilament Heavy Subunit

121 Introduction

Neurofilament Function

Neurofilament proteins display several characteristics in neurons, such as control of axonal elongation, axonal caliber, and axonal transportation. They are subject to pathological aggregation in neurodegenerative diseases, and exhibit a protective function in ALS transgenic mice.

The Basic Function of Neurofilament is Supporting Axonal Structure

As one of the members of the cytoskeletal system, NFs work together with microtubules and microfilaments to enhance structural integrity, cell shape, and cell and organelle motility. The major function of NFs is to control the axonal caliber, which is

directly related with phosphorylation state. This is important since the speed of

conductivity of an impulse down the axon is proportional to its caliber. NFs are

particularly abundant in neurons with large diameter axons (>5µM) such as those of

periphery motor neurons controlling skeletal muscle, where fast impulse conduction

velocities are crucial for proper functioning.

Detrimental Effects of NF Accumulations

It is well known that accumulation of NFs is a general hallmark for several

neurodegenerative diseases. These include ALS, AD, Lewy bodies in Parkinson’s

disease, progressive supranuclear palsy, Charcot-Marie-Tooth disease, diabetic

neuropathy, and giant axonal neuropathy (Hirano et al., 1984;Shepherd et al., 2002; Mori

et al., 1996; Bomont et al., 2000; Watson et al., 1994; Schmidt 1996; Schmidt et al.,

1997;).

122 The detrimental effect of these accumulations is seen when the protein inclusions in axons mechanically block the transportation of particles through the axon, which will eventually lead to neuronal death (Xu et al., 1993; Grant and Pant, 2000; Lariviere and

Julien, 2004). Studies show that abnormal accumulations of NFs in the perikaryon of motor neurons could be induced by overexpression of any of the three NF subunits individually. Increasing the levels of NF-H or NF-M not only impairs NF transport into the axon, but also inhibits dendritic arborization (Xu et al., 1993; Grant and Pant, 2000;

Lariviere and Julien, 2004). Ma et al. (1999) overexpressed human NF-L in mice and reported the severe loss of neurons in the parietal cortex and ventrobasal thalamus with age. Xu et al. (1993) showed transgenic mice overexpressing ~4 fold of mouse NF-L cause neuron degeneration and neuron loss, which resemble the pathology of ALS.

Recently, Sanelli et al. (Sanelli et al., 2004) using cells and mice overexpressing human

NF-L showed that the co-localization of copper/zinc superoxide dismutase (SOD1) and neuronal nitric oxide synthase (nNOS) with NF aggregates may cause sequestration of nNOS in NF aggregates, which leads to enhanced NMDA-mediated calcium influx, a potential cause of neuronal death in amyotrophic lateral sclerosis (ALS). This hypothesis sheds light on the mechanism of NF aggregations in causing neuron death. All these studies emphasize the importance of subunit stoichiometry for correct NFPs assembly and transport.

Protective Effects

Other studies have shown NFPs have a protective effect for neuronal survival.

Couillard-Després et al.(1998) showed that crossing transgenic mice overexpressing human NF-H and mice expressing a mutant superoxide dismutase (SOD1G37R) related to

123 human ALS increase the lifespan of the progeny by up to 65%. Another example of NFH

protection is found in the analysis of doubly transgenic mice that overexpress both

peripherin and human NF-H transgenes in NFL knockout mice (Beaulieu and Julien,

2003). The NF-H overexpression completely rescued the peripherin-mediated

degeneration of motor neurons in vivo. The overexpression of human NF-H shifted the

intracellular localization of peripherin from the axonal to the perikaryal compartment of spinal motor neurons, which suggests that the protective effect of extra human NF-H proteins in this situation partly comes from the sequestration of peripherin into the perikaryon of motor neurons thereby abolishing the development of axonal IF inclusions that might block transport. These findings illustrate again the importance of IF protein stoichiometry in formation, localization, and potential toxicity of neuronal inclusion bodies. This is further supported by studies showing NFs were not required for pathogenesis induced by mutant SOD1 because in the absence of NF-L, with depletion of

NFs in axons, life span was extended by 15% in SOD1 mutant mice (Williamson et al.,

1998). All of these studies lend support to a protective effect during condition of depleted axonal NFs or accumulated NFPs in the cell perikaryon.

The disruption in one allele of the NFL gene in mutant SOD1 mice resulted in a

40% decrease in axonal NF content and caliber of motor axons yet had no effect on disease severity and life span of SOD1 mutant mice (Nguyen et al., 2000), which suggests that axonal depletion of NF alone was not responsible for slowing disease in mutant SOD1 mice. The authors’ hypothesis is that perikaryal NFs can alleviate the toxicity of mutant SOD1 by sequestering p25/Cdk5 complex and by acting as a phosphorylation sink for deregulated kinase activity, thereby reducing the potential

124 toxicity of hyperphosphorylation of tau and of other Cdk substrates like Rb (Nguyen et

al., 2001, 2003).

Another possible explanation of NFPs protective function may relate to the

normal function of these proteins. In ALS, AD, and control brains, axonal NF displays

the most abundant carbonyl modifications in gray matter of the central nervous system

[146,147]. The 205 kDa NF-H and, to a lesser extent, the 160 kDa NF-M, displayed the

majority of the adducts of the lipid peroxidation aldehyde product—hydroxynonenal

(HNE) in SOD1G85R mutant mice and 1-33 months normal aged mice (Wataya et al.,

2002). The level of HNE modification showed no difference with age in any cases. These

results also support the NF sequestration hypothesis that NF sequester toxic components,

either p25/Cdk5 complex or toxic product of oxidation or others, consequently lead to protection of the cells and maintenance of the physiological condition (Liu et al., 2003)

(see Figure 1.5 in Chapter 1).

Many questions remain, such as how to reconcile apparently conflicting functions of NFs. Resolution of NFs role of toxicity and protection requires an understanding of why some NF knockout mice and NF transgenic mice are completely normal even though they have perykaryal inclusions for years, and what the initial cause(s) of the degeneration of neurons is (are). In addition to investigating the perykaryal inclusions, increased effort needs to be put into determine the initial causes leading to inclusion formation.

Research Relevance

Neurofilaments as the enriched proteins in neuronal axons are required for axonal extension and axonal transportation. Normally it is heavily phosphorylated in most of the

125 serine residues among the KSP repeats resided in the C-terminus of NFH. In addition,

this phosphorylation is hypothesized to have multiple functions related to neurofilament

aggregation and axonal transportation, but it is never been linked to the oxidative

modifications.

NFH and NFM are almost the only human proteins to have multiple KSP repeats

in their C-termini, and more importantly, these sites are also the targets of

phosphorylation (Hsieh et al., 1990; de Waegh et al., 1992; Cole et al., 1994 Jaffe et al.,

1998) and HNE modification (Wataya et al., 2002). In addition, we found that

dephosphorylation could reduce the HNE antibody recognition, which led us to determine

whether phosphorylation controls, at least partially, the level of NFH-HNE adduction. In

chapter 2, we have already determined that NFH-HNE modification is a

phosphorylation-dependent and sequence-specific modification that can be regulated by

phosphorylation and dephosphorylation of NFH.

Recently two transgenic mice studies have shown that human NFH presented a

protective function. Couillard-Després et al., (1998) showed that expressing human NFH

in superoxide dismutase mutant mice (SOD1G37R) increases the lifespan by up to 65%,

indicating human NFH has a protective function for oxidative damage. Picklo et al.,

(2003) showed that overexpressing human NFH can rescue the peripherin-mediated

motor neurons death in NFL knock-out mice. The mechanism(s) is (are) undiscovered,

but these studies open the question why and how NFH has the protective function.

In this chapter, using neuronal and other cell models overexpressing NFH, we showed that overexpressing NFH protects cells from HNE toxicity and that HNE treated cells adjust themselves to the HNE toxicity by increasing their cellular NFH level and

126 phosphorylation level. These cell self-protective functions act through activation of

kinases including MAPK and P38 kinase pathway. this data directly shows that NFH

protein not only performs the cytoskeletal function but also protects neurons from

oxidative attack through activation of different kinase pathways. In this study, for the first time using cell culture models, we demonstrated that phosphorylation and increased levels of NFH could protect neuronal cells from HNE toxicity. This function is most likely through the activation of kinases induced by HNE treatment. It directly shows that

NFH, as a cytoskeleton protein, can be a scavenger for at least one of the free radicals.

127 Materials and Methods

Antibodies:

SMI-34, monoclonal antibody specific to phospho epitope on human NFH;

SMI-32, monoclonal antibody specific to nonphospho epitope on human NFH; NF200, monoclonal antibody specific to total NFH (Sternberg Monoclonal Inc); phospho-MAPK

, monoclonal antibody specific to phospho-ERK1/2 (Biolab); monoclonal antibody specific to Actin (Chemicon International, Inc.); HNE-Michael antibody, rabbit polyclonal antibody specific to HNE-Michael adduct (Alexis Biochemicals).

Chemicals and inhibitors:

HNE is synthesized in Dr. Lawrence M. Sayre’s laboratory as the method in the

literature (Nadkarni and Sayre, 1995).U0126 is the specific inhibitor for MAPK and

SB203580 is the specific inhibitor for p38 kinase (Calbiochem). 10 µM of inhibitors were

used in the cell culture 30 minutes before HNE treatment.

Mouse NFH cDNA:

Mouse NFH cDNA can is a generous gift from Dr. Don Cleveland (San Diego,

CA) (Lee et al., 1993). pMSV-NFH is a cDNA with full length NFH and MSV promoter was already cloned in pUC19 vector. Ampicillin was used for the cDNA amplification

process.

Cell culture:

Human M17 neuroblastoma cells and mouse N2A neuroblastoma cells are commonly used cell lines for neuronal cell studies, kindly provided by Dr. Robert B.

Petersen (Institute of Pathology, CWRU). Cells were cultured in Opti-MEM media (Life

Technologies) with 5% donor calf serum and 1% penicillin/streptomycin with fungizone

128 (Life Technologies). 10 ml 1X105 or 2 ml 1X105 cells/ml cells were put in 10 cm dish or

each well in the six-well plates. After overnight incubation, media was changed and the

cells were used for all the treatment.

M17 cells were differentiated by adding 10 µM retinoic acid in the media for 5-6

days before HNE treatment..

In HNE experiments, 50 µM HNE final concentration was applied in the cell

culture for different times (0-4 hours), After pipeting the cells off and centrifuging for 10

minutes at 4,000 rpm RT (IEC, CentraMP4R centrifuge), the cells were lysed by using 50

µl 1X cell lysis buffer (Cell Signaling). Protein concentration was determined by using

BCA protein assay (Pierce) and 20 µg total protein/lane was loaded for SDS PAGE.

In cell viability experiments, M17 cells were plated in six-well plates. After overnight incubation, both different concentrations of HNE and different incubation times were used.

In transfection experiments, cells were transfected with mouse NFH cDNA with

DOTAP liposomal transfection reagent (Roche) as the procedure suggested in the product sheet. After 24 hours transfection, final concentration of 50 µM HNE was put in the M17 cell culture medium and final concentration of 0.5 mM HNE was put in N2A cell culture medium for different time (0-4 hours, for >50% cell death in 4 hours). Cell death was determined using LDH assay (Invitrogen) and trypan blue staining.

In the MAPK kinase activity experiments, cells were first treated with 10 µM of the specific inhibitors for 30 minutes before adding final concentration of 50 µM HNE.

Cells were collected after a 4-hour treatment with HNE. Protein concentration was

129 determined by using BCA protein assay (Pierce) and 20 µg total protein/lane was loaded for electrophoresis.

Cell viability assay: LDH assay and Trypan blue staining:

LDH cell viability assay (Roche) is a colorimetric assay for the quantification of

cell death and cell lysis It is based on the measurement of lactate dehydrogenase (LDH) activity released from the cytosol of damaged cells into the supernatant. The experiments were performed in 96-well plate as the protocol.

Trypan Blue was diluted at 0.8 mM in PBS. The stock will be stable for 1 month

at room temperature. After all the pretreatment, cells were trypsinized and mix cells with

1:1 with trypan blue solution, then count cells on hemocytometer. Viable cells exclude trypan blue, while dead cells stain blue due to trypan blue uptake.

Protein Assay

Protein concentration was determined by using BCA protein assay (Pierce) and 20

µg total protein was loaded for electrophoresis.

Immunoblotting:

After all the pre-treatment, Membranes were blocked with 10% milk 1 hr, washed with Tris-buffered saline-Tween (50 mM Tris, 150 mM NaCl, 0.1% Tween20) 5X5 min, incubated with primary antibody 16 hrs 4 ℃, after five washes of 5 min in Tris-buffered saline-Tween, peroxidase-labeled secondary antibodies were applied for 1 h at room

temperature. After rinsing again as described before, the blots were developed together for the same length of time using enhanced chemiluminescence (ECL) reagent (Santa

Cruz Biotechnology) and imaged by using image acquisition and analysis software

(Ultra-Violet Products Ltd.).

130 Quantification of the intensity of the blots was performed by using image-analysis software (KS300, Zeiss).

Statistical Analysis:

All the experiments were repeated for at least three times. The data were analyzed using the students’ t-test.

131 Results

M17 cells are human neuroblastoma cells and differentiate under the treatment of retinoic acid forming long axons. 50 µM HNE treatment of M17 differentiated cells induced higher levels of NFH expression in 4 hours, including the both phosphorylated and non-phosphorylated forms of NFH (Fig 3.1). An ~5-fold increase of total NFH,

~10-fold increase of phosphorylated NFH, and ~3-fold increase of non-phosphorylated

NFH were detected using western blot analysis of cell lysate from M17 differentiated cells after 4-hour 50 µM HNE treatment.

NFH phosphorylation level was measured over a time course of HNE treatment. It was found that phosphorylation of NFH protein increased and peaked within 1 hour and

then decreased (Fig 3.2).

The activity of MAPK (including Erk1 and 2) was determined by using

anti-active-MAPK monoclonal antibody on western blot of the cell lysate. The result

showed that MAPK were activated in M17 differentiated cells following HNE treatment

(Fig 3.3). Inhibition of Erk1/2 by specific inhibitor U0126 showed total inhibition of

Erk1/2 activity (Fig 3.3). Total level of MAPK was re-probed by using anti-total MAPK

monoclonal antibody (Fig 3.3), and it showed same level of total MAPK.

M17 differentiated cells had higher level of NFH than non-differentiated M17

cells (Fig 3.4B). When M17 differentiated and undifferentiated cells were treated with

various concentrations of HNE for 4 hours, the differentiated cells were found to be more

resistant to HNE toxicity than the non-differentiated cells (Fig 3.4A). Since HNE can

132

Figure 3.1 HNE treatment induced higher level of NFH in M17 cells. 4 hour 50

µM HNE treatment induced around 5 fold in crease of total NFH level (NF 200). The included around 3 fold increase in nonphosphorylated NFH (SMI-32) and 10 fold increase in phosphorylated NFH (SMI-34).

133

Figure 3.2 HNE treatment induced higher phosphorylation level of NFH. 50

µM HNE induced higher phosphorylation level of NFH in different time, which peaked in samples of HNE treated M17 differentiated cells for 2 hour.

134

Figure 3.3 MAPK were activated in M17 cells after HNE treatment. 50 µM 4 hour

HNE treatment induced the activation of MAPK. Specific inhibitor U0126 totally inhibited MAPK activity.

135 induce increased protein and phosphorylation levels of NFH within 4 hours (Fig 4.1, Fig

4.2), one possibility is that increased NFH and its phosphorylation may play a role in the protection. However, retinoic acid can induce a broad range of gene expression in cells, so this is not an ideal cell model to investigate this point.

To further determine the relationship between increase of NFH /phospho-NFH and protection from HNE toxicity in neuronal cells, a mouse cDNA was obtained and transfected into cells to overexpress NFH. M17 differentiated cells were transiently transfected with the mouse NFH cDNA and exhibited a more than 5-fold increase of total

NFH and around 3-fold increase of phosphorylated NFH after 24 hours (Fig 3.5B). When the cells were treated with HNE, the transfected M17 cells showed significant protection from HNE cytotoxicity than the control M17 cells transfected with empty vector (Fig

3.5A).

This result was also confirmed in another mouse neuronal cell line—N2A cells.

After transfection, the N2A cells produced more than 10 fold of total NFH and more than

5 fold of phosphorylated NFH. NFH (Fig 3.6B). When treated with HNE, the transfected

N2A cells showed significant protection from HNE treatment compared to the control

N2A cells transfected with empty vector (Fig 3.6A).

Both types of cells overexpressing NFH did not show any morphological abnormalities even after 7 days under converse phase microscope (data not shown), which indicate the overexpression of mouse NFH even 10 fold than basal level may be not toxic for these cells in a short time.

136

Figure 3.4 M17 differentiated cells with higher level of NFH showed significant

protection form HNE cytotoxicity with LDH cytotoxicity assay. M17 cells were

differentiated by incubating with 10 µM retinoic acid. Differentiated M17 showed around

4 fold increases in NFH level. M17 differentiated cells showed more protection than M17 undifferentiated cells in different HNE concentration treatment in 2 hours. RA: retinoic acid. “*”=P<0.05.

137

Figure 3.5 Overexpressing NFH in M17 cells resulted in significant protection from HNE cytotoxicity using trypan blue staining. After transfection with mouse NFH,

M17 cells showed significant increases in both total NFH and phospho-NFH level.

Transfected M17 cells were more resistant to HNE toxicity than empty vector transfected

M17 cells.

138

Figure 3.6 N2A cells showed significant protection with overexpressing NFH using trypan blue staining. N2A cells are mouse neuroblasdoma cells. Transfected N2A

cells showed significant increases in both total NFH level and phospho-NFH level.

Overexpressing mouse NFH in N2A cells showed more protection than empty vector

transfected N2A cells.

139 Discussion

In this study, we found that HNE treatment induced increases in both total NFH and phosphorylated NFH protein levels. In addition, increased levels of NFH and phosphorylated NFH are coincident with highergreater protection from HNE cytotoxicity in cultured neuronal cell lines. When mouse NFH is overexpressed in two different neuronal cell lines, the transfected cells show significant protection from HNE toxicity as compared to control cells transfected with empty vector. Also further studies showed that

MAPK kinases were activated by HNE treatment in M17 cells, which already been shown MAPK can phosphorylate NFH. Combining all these results, this study directly demonstrated that NFH protein and its phosphorylation protect neuronal cells from oxidative stress.

Lipoxidation adducts are increased in neuronal cell bodies in AD cases and are present in axons in the white matter of infants and aged controls (Wataya et al., 2002).

These findings propose the questions of how these lipid peroxidation products and specific axonal proteins interact and whether neurons have compensatory mechanisms. In this study, we focused our studies on HNE adducts because they are the most readily detected and best characterized of the aldehyde-derived adducts. HNE is highly reactive and considered one of the most neurotoxic aldehydes produced in vivo (Sayre et al.,

1997). In addition, whereas HNE adducts are among the best characterized chemically, it is likely that HNE can be used as a model for the other reactive aldehydes produced by lipoxidation or glycoxidation because they have common features.

It has already been shown that HNE can induce the activation of MAPK,

PI3K/AKT pathways, which are serine/threonine kinases (Liu et al., 1999; Dozza et al.,

140 2004), and MAPK is a major kinase for neurofilament proteins phosphorylation (Liu et

al., 1999; Dozza et al., 2004). Therefore, NFH is probably phosphorylated by MAPK

after the kinases are activated by HNE treatment.

HNE also induce glutathione (GSH) biosynthesis, the major cellular molecule for

clearance of up to 50% free HNE (Forman et al., 2003). The remaining free HNE will at

least partially attack other biomolecules, such as enzymes in neurons, therefore, neurons

should have a compensatatory system for the extra free HNE.

It has already been reported that lower levels of HNE have physiological

functions in promoting cell proliferation (<10 µM), and higher level of HNE cause apoptotic pathway activation, and even higher level of HNE can directly cause necrosis of the cells (Mattson,2002). In this study, the cells are most likely dying through the apoptotic pathway.

Consistent with the cell culture studies, two mouse studies already demonstrated that overexpression NFH has a protective function (Couillard-Després et al., 1998; Picklo et al., 2002). The mechanism(s) is (are) unknown, but one hypothesis is that NFH may act as sink for oxidative stress. This is consistent with the property of the protein and also reasonable for neuronal survivable. NFH is a lysine-rich protein with the unique characteristic of multiple KSP repeats resided in the C-terminus of NFH. Phosphorylation is most prominent in the serine residues among the KSP repeats. In addition, the free amino groups from the side-chain of lysine provide a huge reactive pool for the reactive oxygen species (ROS), providing a possible molecular basis for a regulative mechanism for HNE modification.

141 Our recent studies have already proven this point (See chapter 2). In that study, we showed that NFH can act as a scavenger for HNE and this process is regulated by phosphorylation and dephosphorylation of NFH in the KSP motifs. This amazing regulative mechanism is arranged in the unique KSP repeats residing in the C-terminal of

NFH protein. NFH can adduct with more HNE molecules under phosphorylation or reverse the HNE adduct under dephosphorylation. This mechanism provides strong evidence for NFH acting as a scavenger in vivo.

In this study, only short-term protection of NFH was studied. In conditions of high concentrations of HNE, many cells are dying after four-hour HNE treatment presumably through apoptosis (Lee et al., 2004). Whether the overexpression of NFH can provide long-term protection in lower HNE concentration, more like the cellular condition, is another interesting question worthwhile to be determined in the future.

M17 cells are human neuroblastoma cell line, which can be differentiated by retinoic acid treatment and form long axons, enriched in neurofilament proteins. In addition, M17 is an efficient working cell line for transfection experiments that has already been used intensively in our lab. For further studies, primary neurons could be a good model for this type of studies with an efficient transfection method. N2A cells are also commonly used in transfection and neuronal cell studies.

In this study, mouse NFH was overexpressed in the neuronal cells, so whether the

NFH is in the free subunit form or assembled form is unknown. Since neurofilament has both free form and assembled form in vivo, it is worthwhile to ask which form is the functional form. Therefore, it will be helpful to study the NFH subunit and the filament separately.

142 In summary, neurofilaments as the most abundant proteins in the neurons, not only provide a cytoskeletal function but also provide further physiological function in adjusting the homeostasis of the free radicals and protect cells from oxidative attack.

Combined with the transgenic mice studies, it becomes clearer that NFH protective functions is at least partially an anti oxidative stress function. On the other side, this HNE modification on NFH can cause protein crosslinking or other changes promoting the NF aggregation in term of decades. The aggregation of neurofilament and other cellular proteins in neurons cause neuronal degeneration in patients.

143

Chapter 4

Oxidative Modification of Tau Protein

Is Involved in NFT Formation

144 Introduction

Tau Protein Structure

Tau is a hydrophilic protein that has been widely characterized in solution. It

appears as a random coiled protein by analysis of the circular dichroism spectra

(Cleveland et al., 1977). In addition, tau has been suggested to be a highly asymmetric

protein, compatible with the long rod structure observed by electron microscopy

(Hiokawa et al., 1988).

Several different tau polypeptides exist in the brain protein extract showed by gel electrophoresis. These polypeptides are generated by both alternative RNA splicing

(Goedert et al., 1989; Himmleret al., 1989a, b; Kosiket al., 1989) and posttranslational modifications including phosphorylation (Goedert et al., 1992). In NFT, there are six CNS

tau isoforms in their phosphorylated form. The six tau isoforms are as follows: 1) containing exons 2, 3, and 10, plus all the constitutive exons (Baker et al., 1999); 2) having exons 2 and 3; 3) containing exons 2 and 10; 4) having only exon 2; 5) with only

exon 10; and 6) only containing the constitutive exons. This combination of

hyperphosphorylated tau isoforms results in the appearance of three major electrophoretic bands with a mobility corresponding to that of proteins with a relative molecular weight of 68,000, 64,000, and 60,000 (Delacourte et al., 1999; Goedertet al., 1991, Greenberg et

al., 1992; Lee et al., 1991). Furthermore, AD brains contain higher quantities of tau than unaffected controls (Khatoon et al., 1992’ 1994).

Tau Phosphorylation

In the 1980s, tau was found as a phosphoprotein (Baudier et al., 1987,

Grundke-Iqbal et al., 1986, Iharaet al., 1986), and all these studies focused on the

145 serine/threonine phosphorylation of the tau protein. Recently, one study has focused on

its phosphorylation on tyrosine (Ihara et al., 2002).

The phosphorylation of tau is developmentally regulated; it is higher in fetal

neurons and decreases with age during the development. Hyperphosphorylation of tau is already been related to pathological situations (tauopathies) (Lovestone et al., 1997, Ihara et al., 1995). Interestingly, different tau isoforms has different phosphorylation sites

(Hernandez et al., 2001). This could be due to the different cellular localization or

subcellular compartmentalization of the different tau isoforms, or the fact that different kinases or phosphatases can modulate tau phosphorylation in a different way.

Hyperphosphorylated tau has been described in extracts of both normal and AD tissue

(Kopke et al., 1993).

There are 79 putative serine or threonine phosphorylation sites on the longest

CNS tau isoform, which contains 441 residues. And there are even more in PNS tau,

although its phosphorylation has not been widely studied. The kinases involved in the phosphorylation of Tau protein have been divided into two main groups: proline-directed

kinases and non-proline-directed kinases. The first group includes tau protein kinase I

(glycogen synthase kinase 3, GSK3), tau protein kinase II (cdk5), MAP kinase (p38),

JNK, and other stress kinases or cdc2. The second groups includes protein kinase A

(PKA), protein kinase C (PKC), calmodulin (CaM) kinase II, MARK kinases (Goedert et al.,1997, Hanger et al., 1992, Imahori et al., 1997, Lucas et al., 2001), or CKII that modifies residues close to acidic residues mainly in exons 2 and 3 (Correas et al., 1992).

In many cases, phosphorylation regulates the binding of tau to microtubules or to the

146 membrane (Brandt et al., 1995). Thus, phosphorylation appears to play a predominant role in regulating tau function.

One example is GSK3, which may play a major role in the phosphorylation of

Tau. Two types of GSK3 phosphorylation have been proposed: primed (following prior phosphorylation of the substrate by another kinase) or unprimed phosphorylation (Frame et al., 2001). Primed phosphorylation occurs at threonine-231 and affects microtubule

binding, while unprimed phosphorylation occurs at serine-396 or -404 and does not affect microtubule binding (Cho et al., 2002). These sites can be identified by the AT180 and

PHF-1 antibodies, respectively (Cho et al., 2002).

GSK3 phosphorylation of tau facilitates its aggregation into filamentous polymers

(Jackson et al., 2002). The β-amyloid peptide (Aβ) is also involved in augmenting the phosphorylation of tau by GSK3 (Yankner et al., 1996), probably by increasing the

enzymatic activity of the kinase (Alvarez et al., 1999), which only proves to be toxic for neurons in which tau protein is present (Rapoport et al., 2002).

The phosphorylation of tau may promote a conformational change that can be

functionally reversed in the presence of trimethylamine N-oxide (TMAO), a natural occurring osmolyte (Tseng et al., 1999). This conformational change can also be reversed

by the chaperon protein Pin-1 (Lu et al., 1999), a molecule that upon tau binding facilitates the posterior action of PP2A in dephosphorylating the protein (Zhou et al.

2000). These conformational changes could facilitate tau aggregation, and recently, it has been suggested that such conformational changes (or others) may result in an increase of

α-helices in the secondary structure of tau, since the content of α-helices is greater in tau from PHF (Jicha et al., 1999, Minoura et al., 2002, Sadqi et al., 2002). However, it has

147 also been suggested that filament formation is also partially dependent on the presence of certain β-sheet structures within tau (Von Bergen et al., 2000). It would be of interest to know whether the tau present in other aggregates such as Pick's bodies adopts a different conformation. Nevertheless, it does seem that a conformational change is undertaken

when tau polymerizes into PHF (Abraha et al., 2000, Ghoshal et al., 2002). Furthermore, it appears that this conformational change could involve the binding of the

amino-terminal region of the tau molecule to its microtubule binding region (Carmel et al., 1996).

Several phosphatases like protein phosphatase (PP) 1, PP2A, PP2B (calcineurin),

and PP2C (Geddes et al., 1993, Goedert et al., 1992, Szucs et al., 1994, Yamamoto et al.,

1995) have been implicated in reversing the phosphorylation of tau. However, only PP1,

PP2A, and PP2B (Gong et al., 1994) have been shown to dephosphorylate abnormally hyperphosphorylated tau. Although PP2C can dephosphorylate tau when it is

phosphorylated by PKA in vitro, it is not capable of dephosphorylating the abnormally hyperphosphorylated tau isolated from AD brain tissue (Gong et al., 1994).

Tau Polymerization

Montejo et al. (1986) described that purified tau protein can form fibrillar

polymers resembling the PHF found in the brain of AD patients and that deamidation

could facilitate this polymerization (Montejo et al., 1987, 1988) which was found in PHF

(Watanabe et al., 1999). In vitro, tau assembly has been further studied (Troncoso et al.,

1993, Wilson et al., 1997). The requirement of a high concentration of protein for tau to polymerize suggests that other factors could be needed to facilitate tau assembly in vivo.

The sulfoglycosaminoglycans (sGAGs), which are present along with tau in NFT, were

148 the first molecules tested and found to facilitate tau polymerization in vitro independently of the phosphorylation state of tau (Goedert et al., 1996, Perez et al., 1996). sGAGs are polyanions, and other polyanions such as the glutamic acid-rich region present at the carboxy-terminal region of tubulin can also facilitate aggregation, the third tubulin

binding motif of the tau molecule are required for this aggregation (Perez et al., 1996).

Oxidation could play a role in tau aggregation. Oxidation of cysteine to produce

disulfide cross-linking favors tau self-assembly in tau 3R molecules, where a single

cysteine is present, but not in tau 4R molecules, where the presence of two cysteines may

permit the formation of intramolecular disulfide bonds (Barghorn et al., 2002). In vitro

studies have shown that tau can be assembled through oxidation processes (Fenton's reaction) in the presence of iron (Troncoso et al., 1993). Also, fatty acids like arachidonic acid can induce tau polymerization in vitro (Abraha et al., 2000, Wilson et al., 1997). This type of polymerization could be related to the possible interaction of tau with plasma membrane components (Arrasate et al., 1997, 2000). Another lipid peroxidation molecule-HNE has been shown facilitating tau aggregation in vitro and in cells (Perez et al., 2000, 2002).

It has been shown that anionic micelles and vesicles induce tau fibrillation in vitro

(Chirita et al., 2003). It has been suggested that tau could play a role in the activation of phospholipase C (PLC) -γ to generate arachidonic acid through the hydrolysis of phosphoacetylcholine (Hwang et al., 1996). The arachidonic acid generated could

facilitate tau aggregation (Abraha et al., 2000), possibly due to the arachidonic acid micelles can act as polyanions since their negative charged carboxyl groups are exposed

at the surface. A compound produced from oxidation of arachidonic acid-HNE has been

149 shown to facilitate tau assembly, but only if tau is hyperphosphorylated (Perez et al.,

2000, 2002), which is in part due to GSK3 phosphorylation. On the other hand, it has been proposed that carnosine can quench the effect of HNE (Aldini et al., 2001), and antioxidants like N-5 butyl hydroxylamine could be used as neuroprotectors (Atamna et al., 2001). Finally, α-synuclein has also been seen to facilitate tau assembly (Giasson et

al., 2003).

A Phosphorylation Before or After Aggregation

A long time argument is whether tau phosphorylation occurs before or after PHF assembly. The antibody 12E8 recognizes a phospho-serine (S262) located in one of the

tubulin binding motifs present in tau molecule. The binding of this antibody to tau is

augmented in AD extracts; however, this antibody does not react with assembled PHF

(Hernandez et al., 2002). This suggests that tau phosphorylation occurs before its

assembly, and this is supported by other studies (Gordon-Krajcer et al., 2000). Indeed, strong evidence exists that the phosphorylation of tau promotes its self-assembly (Alonso

et al., 2001).

B. Toxic Effects of Tau

In neurodegenerative diseases, the cytotoxicity mediated by tau could be due to its

hyperphosphorylation or to the formation of aberrant aggregates. In the first case, it has been observed that expression of pseudo-hyperphosphorylated tau promotes toxicity

associated with the induction of apoptotic cell death (Fath et al., 2002). This result is in agreement with observations in GSK3 transgenic mice (Hernandez et al., 2002).

There is an inverse correlation between the number of extracellular tangles and the number of surviving neurons in AD, suggesting that neurons developing

150 neurofibrillary lesions could degenerate (Goedert et al., 2002). On the other hand, the

appearance of extracellular NFT occurs in those regions that contain neurons with intracellular NFT. This suggests that intracellular inclusions precede cell death and that extracellular NFT form as a result of cell lysis, perhaps due to the binding of tau to extracellular matrix components like sGAGs (Perez et al., 1996). Thus, the presence of

tau aggregates could be postulated as toxic and result from the fact that tau aggregates are

sticky structures. This tau form could bind to proteins needed for normal metabolism and

therefore deprive the cell of these proteins, like ferritin (Hernandez et al., 2001) or Pin-1

(Lu et al., 1999). It has also been proposed that abnormally hyperphosphorylated

cytosolic tau might sequester other MAPs like MAP1B or MAP2. Such sequestering may result in the destabilization and disassembly of microtubules, and in the appearance of

morphological changes that could promote the disruption of synaptic contacts (Iqbal et

al., 1994).

C. PHF Morphology

It has been proposed that PHF morphology may depend on the proportion of tau

isoforms containing three or four tubulin binding motifs (Goedert et al., 1996), on the

association with sGAGs (Arrasate et al., 1997), or on the presence of specific residues present in tau molecule (DeTure et al., 2002). Since the pioneer paper of Kidd (Kidd et

al., 1963), different techniques have been applied to characterize the structure of PHF, including X-ray diffraction (Wisniewski et al., 1976), electron microscopy,

high-resolution transmission electron microscopy (TEM), atomic force microscopy, or cryoelectron microscopy (Ruben et al., 1991, Wischik et al., 1985). The results suggest that the structure of PHF is compatible with a helical ribbon made up of two parallel

151 strands. The formation of NFT from PHF appears to be facilitated by glycation (Ruben et

al., 1995), oxidation (Perez et al., 2000, 2004).

The Pathological Relevance Between Tau Protein and AD

The best-known tauopathy is AD, in which two main pathological structures form in the brains of patients: senile plaques (composed of the β-amyloid peptide), and neurofibrillary tangles (NFT). These NFT are made up of paired PHF comprised of

hyperphosphorylated tau (Grundke-Iqbal et al., 1986). The number of NFT has been

correlated with the degree of dementia in this disease (Arriagada et al., 1992), and thus

there is great interest to know how these structures form.

NFT are composed of bundles of paired helical filaments (PHF), the major component of which is the microtubule-associated protein τ (Grundke-Iqbal and Iqbal,

1989; Lee et al., 1991). In PHF, τ is abnormally hyperphosphorylated (Grundke-Iqbal et al., 1986; Trojanowski and Lee, 1995; Lee et al., 2001), ubiquitinated (Iqbal and

Grundke-Iqbal, 1997; Mori et al., 1987; Perry et al., 1987), oxidized (Mattson et al.,

1995; Takeda et al., 2000a), proteolytically processed and aggregated into filaments (Lee et al., 2001; Lovestone and McLoughlin, 2002; Perez et al., 2002) when compared to that in normal neurons. Hyperphosphorylation of τ renders it unable to bind to microtubules and therefore unable to promote or maintain microtubule assembly (Baudier and Cole,

1987). Moreover, hyperphosphorylation makes τ more resistant to proteolytic degradation, which may play a key role in neurofibrillary degeneration in AD patients

(Iqbal et al., 1998; Eidenmuller et al., 2000). Hyperphosphorylation of τ most likely reflects a combination of increased kinase and decreased phosphatase activities (Gong et

152 al., 2000). Many kinases such as GSK3, p38, ERK, and PKC can phosphorylate τ, although which kinase is critical in producing hyperphosphorylated τ in AD is unknown.

In previous studies, we found a clear mechanistic relationship between τ

phosphorylation and oxidative stress, a key pathogenic modulator of disease (Perry et al.,

1998; Smith et al., 2002a; Gomez-Ramos et al., 2003). Data from the literature shows

that oxidative stress activates several kinases such as ERK, p38, and JNK, which have

been shown to be activated in AD (Zhu et al., 2000, 2001, 2002) and are capable of

phosphorylating τ (Reynolds et al., 2000). Additionally, once phosphorylated, τ and

other cytoskeletal proteins are vulnerable to modification by carbonyl products of

oxidative stress (Perez et al., 2000, 2002; Wataya et al., 2002) and consequent

aggregation into fibrils (Perez et al., 2000).

In previous studies, we found a clear mechanistic relationship between τ

phosphorylation and oxidative stress, a key pathogenic modulator of disease (Perry et al.,

1998; Smith et al., 2002a; Gomez-Ramos et al., 2003). Data from the literature shows

that oxidative stress activates several kinases such as ERK, p38, and JNK, which have

been shown to be activated in AD (Zhu et al., 2000, 2001, 2002) and are capable of

phosphorylating τ (Reynolds et al., 2000). Additionally, once phosphorylated, τ and

other cytoskeletal proteins are vulnerable to modification by carbonyl products of

oxidative stress (Perez et al., 2000, 2002; Wataya et al., 2002) and consequent

aggregation into fibrils (Perez et al., 2000).

In vitro analysis of the assembly of τ into fibrillar polymers has indicated that

recombinant, unmodified, or weakly phosphorylated τ (Montejo et al., 1986; Crowther et

al., 1992; Wille et al., 1992) is able to polymerize, while hyperphosphorylated τ has a

153 much lower capacity for polymerization (Schneider et al., 1999), which indicates that

there must be an alternative mechanism to facilitate the assembly of carbonyl-modified

phosphorylated τ into fibrillar polymers in vivo. Anionic compounds like sulfated

glycosaminoglycans (sGAG), RNA, and heparin can facilitate τ polymerization (Siedlak

et al., 1991; Goedert et al., 1996; Kampers et al., 1996; Perez et al., 1996). Proteolysis

(Wischik et al., 1988), glycation, or direct oxidation (Grundke-Iqbal et al., 1986;

Troncoso et al., 1993; Yan et al., 1994; Schweers et al., 1995; Ledesma et al., 1998) have

also been implicated in τ aggregation. Fatty acids such as arachidonic acid (Wilson et al., 1995, 1997; Abraha et al., 2000; Gamblin et al., 2000) and products of lipid oxidation

(Uchida et al., 1993; Perez et al., 2000) have also been demonstrated to promote τ aggregation. The relative role of each of these contributors in vivo is unknown. We previously showed that phosphorylated τ, but not other τ forms, polymerize both in vitro

(Perez et al., 2000) and in cells (Perez et al., 2002) by modifications induced by the lipid peroxidation product, 4-hydroxynonenal (HNE), a naturally occurring product of lipid peroxidation (Uchida et al., 1993) that forms protein adducts which are increased in AD

(Montine et al., 1997, 2002; Sayre et al., 1997). These data are consistent with other studies showing that τ conformational change, as detected with the well-studied monoclonal antibody Alz50, can be induced by HNE (Takeda et al., 2000a). To investigate whether HNE-modified τ forms the major epitopes in NFT, we examined seven distinct monoclonal antibodies raised against NFT. The recognition of all seven antibodies is significantly enhanced by HNE treatment, which may indicate that HNE is a major effector for NFT formation.

154 MATERIAL AND METHODS

Antibodies:

Mouse monoclonal antibodies raised against NFT: PHF-1 recognizing phosphoserine 396/404 (Otvos et al., 1994); TG-4 recognizing phosphoserine 235

(Vincent et al., 1996; Weaver et al., 2000); MC15 recognizing phospho-τ (Vincent et al.,

1996); TG-5 recognizing 220-240 of htau40 (Jicha et al., 1997a); MC-1 and Alz50

(Carmel et al., 1996; Jicha et al., 1997a), both recognizing an intramolecular conformation epitope of residues 7-9 and 312-342; TG-3 recognizing conformational epitope involving phosphothreonine 231 (Jicha et al., 1997b). Mouse monoclonal antibodies raised against normal τ: τ-1 recognizing serine 199/202 without phosphates

(Papasozomenos et al., 1987) and 5E2, a τ antibody independent of the phosphorylation state (Joachim et al., 1987; Kosik et al., 1989). Monoclonal antibodies raised against neurofilaments: SMI-32 and SMI-34 (Sternberger Monoclonals, Inc). Polyclonal antisera to ubiquitin (Perry et al., 1989).

Immunocytochemistry:

Hippocampal tissue from AD cases (ages 65,82,90, all with 4 hour postmortem

intervals) as well as age-matched controls (ages 65,72,84) was fixed in methacarn

(methanol:chloroform:acetic acid, 6:3:1), dehydrated and embedded in paraffin.

Methacarn was used since it does not artifactually create an “oxidative damage-like state”

in the proteins caused by carbonyl adduction, in contrast to formalin or

glutaraldehyde-based fixatives. Sections (6µm) were placed on silane-coated slides.

After rehydration through xylene, graded ethanols, water, and treatment with 3% H2O2, during 30 min to reduce endogenous peroxidase activity, sections were blocked with 10%

155 normal goat serum (Sigma) in Tris-buffered saline (TBS) for 30 min. The sections were

stained with the peroxidase anti-peroxidase method (Sternberger, 1986) using

3,3’-diaminobenzidine (DAB) and 0.015% H2O2 as cosubstrates.

Adjacent serial sections were used to directly compare pathological structures

recognized by different antibodies. The number of NFT immunostained by the

antibodies was determined by image analysis (KS300, Zeiss). The number of NFT and

density of immunolabeled structures were quantified using a computer-assisted imaging

of the same five consecutive fields, defined by landmarks, in the CA1 and CA2 regions in

adjacent serial sections.

Preparation of τ:

Human τ was prepared from cortex of cases showing no neuropathological or clinical neurological signs of AD and with a postmortem interval less than 8 hr. τ protein was isolated by a modification of the procedure previously described by Lindwall and

Cole (1984). Tissue was homogenized in five volumes of 0.1M PIPES, pH 6.8, 2mM

EDTA with a Polytron® homogenizer (Brinkman Instruments), then centrifuged at

12,000g for 20 min. The supernatant was brought to 0.75M NaCl and 2%

β-mercaptoethanol (BME), boiled for 4 min and rapidly cooled on ice. The supernatant was centrifuged again at 10,000g for 30 min at 4°C, brought to 2.5 % perchloric acid and left on ice for 15 min followed by centrifugation at 12,000g for 20 min. τ was precipitated from the supernatant with 20% trichloroacetic acid (TCA), and the pellet was washed with ethanol repeatedly. The purified τ was resuspended in phosphate-buffered saline (PBS) or TBS.

156 PHF-τ, which is highly phosphorylated and stable to postmortem dephosphorylation, was prepared as previously described (Greenberg et al., 1992) from

AD cases. Gray matter was homogenized in 10 vol “H buffer” (10mM Tris, 1mM

EGTA, 0.8 M NaCl and 10% sucrose) by using the Polytron® and centrifuged at 27,200g for 20 min at 4˚C. The supernatant was saved, and the pellet was rehomogenized in 10 vol “H buffer”, spun as described above, and combined with the previous supernatant.

The final supernatant was brought to 1% sarcosine and 1% BME and incubated for 2 hr at

RT. After, the supernatant was centrifuged at 22,000g for 1 hr at RT. The pellet was resuspended in 3-4 ml of H buffer, layered over a discontinuous sucrose gradient of 35% and 50% sucrose in 10mM Tris, 1mM EGTA and 0.1% β-mercaptoethanol, and spun for

2 hr at 40,000g in a Beckman ultracentrifuge (SW50 rotor). PHF-τ, layered at the

35/50% interface, was precipitated by trichloroacetic acid and resuspended in PBS.

Recombinant τ (full-length) was prepared as previously described (Perez et al.,

1998).

The phosphorylation state of τ was reduced by treating human τ samples with

20U/ml of alkaline phosphatase type III (Sigma) for 16 hr at 37°C in 0.1M Tris-HCl, pH

8.0, in the presence of 0.01 M phenylmethylsulfonyl fluoride (PMSF) to reduce proteolysis or, alternatively, τ was dephosphorylated by treatment in 50% hydrofluoric acid for 16 hr at 4°C (Smith et al., 1996; Takeda et al., 2000a).

Immunodots:

Samples were treated with 1 mM HNE (Takeda et al., 2000a) for 2 hr before application of protein to the membrane. Dot blots were prepared by applying 4 µg of τ

protein per dot directly onto Immobilon (Millipore) membrane and air drying. The

157 efficacy of the assay we use was validated in a previous study where the Alz50 epitope was examined (Takeda et al., 2000a). In the Takeda study, we observed that

HNE-induced modifications are time- and molecular weight polymer-dependent, i.e., τ is continuously modified by HNE and the Alz50 epitope decreases with the formation of high molecular weight polymers. The amount of protein used per dot coupled with the short antibody incubation times, allowed the assay to be performed to optimize the recognition of HNE-induced modifications. In the same study (Takeda et al., 2000a), using several protein concentrations and Coomassie blue staining, we observed that the assay proceeds in a linear range for the conditions used here. The membrane was then blocked with 10% nonfat dry milk for 15-30 min at RT and incubated for 1 hr with the primary antibodies. After five washes of 5 min each with 0.1% Tween in TBS, secondary horseradish peroxidase labeled antibodies were applied for 1 hr at RT. After extensive washing in Tween-TBS, dot blots were developed using enhanced chemiluminescence (ECL) (Amersham).

Quantitation of the intensity of binding was performed by using image-analysis software (KS300, Zeiss). Experiments were performed at least in triplicate and control dots of untreated proteins were included and used as background readings for each experiment.

Statistical Analysis:

The data were analyzed using the students’ t-test.

158 RESULTS

The specificity of seven unique NFT monoclonal antibodies are summarized

(Figure 4.1).

To compare relative recognition of each of the antibodies against different τ forms, we made comparative studies with recombinant τ, normal τ, and PHF-τ (Figure

4.2). Recombinant τ was poorly recognized by most of these antibodies, except TG-5, probably due to lack of phosphorylation and conformations (Fig 4.2A). Normal τ, which is phosphorylated in vivo, can be recognized by TG-5 and PHF-1 and, to a lesser extent, by TG-4 (Fig 4.2B). PHF-τ was strongly recognized by PHF-1, TG-4, MC15, and TG-5.

Surprisingly, three conformation-dependent antibodies, MC1, Alz50, and TG-3, showed little recognition (Fig 4.2C). This may be due to the disruption of specific τ conformations caused by the purification process, since these antibodies recognized NFT by immunocytochemistry staining of sections from the same AD brain samples (See Fig

4.4 in discussion). Furthermore, we observed that dephosphorylation reduced recognition by the phospho-dependent NFT antibodies (PHF-1, TG4, and MC15) (Figure

4.2B,C), especially in PHF-τ. TG-5 antibody is least affected in all the treatments, and the reason why TG-5 showed lower binding capacity to recombinant τ is unknown.

To determine whether the above described antibody recognition to τ was dependent on or promoted by oxidation-related modifications, we treated the three forms of τ, with or without prior dephosphorylation, with HNE. We found that HNE treatment of normal τ greatly enhanced the antibody recognition from 4 to 34 fold (Figure 4.3B).

Such carbonyl-enhancement was dependent on τ phosphorylation state since dephosphorylation with either phosphatase or hydrofluoric acid prior to HNE treatment

159 greatly reduced the enhancement by HNE. This effect was statistically significant for

all the antibodies tested except TG-5 and PHF-1. PHF-1 antibody is usually considered as

a strong NFT antibody despite of dilution, which may explain the less significant

reduction of PHF-1 after dephosphorylation (Reduction is also apparent). Confirming

the phosphate-dependent nature of the effect of HNE treatment, antibody recognition of

recombinant τ is not enhanced by HNE (Figure 4.3A). Further, HNE had no effect on

the recognition capacity of any of the antibodies against PHF-τ irrespective of the phosphorylation state (Figure 4.3C) suggesting the possibility that PHF-τ has already been modified by endogenous aldehydes to an extent that the conformational change recognized has been fully induced or PHF-τ has acquired some specific structures to prevent further HNE modifications.

In contrast to τ, antibodies against neurofilaments (SMI 32 and SMI 34), phosphorylated or not, presented a lower recognition capacity for neurofilament proteins after treated with HNE (data not shown). Also, treatment of purified ubiquitin with

HNE did not alter the recognition capacity of ubiquitin antibodies (data not shown).

These results exclude the possibility that HNE treatment has a universal enhancement effect for all antibodies.

τ-1 (dephosphorylated τ) and a monoclonal antibody specific to phospho-serine/threonine were used to examine the efficiency of the dephosphorylation process by alkaline phosphatases or hydrofluoric acid. In a previous study, dephosphorylation by HF lead to the complete removal of organic phosphate from τ [61].

5E2 (phosphate independent) was used to control for the maintenance of τ protein levels throughout the treatments (data not shown).

160

Figure 4.1 Summary of the specificity of seven NFT antibodies. (A) They belong

to three groups: PHF-1, TG-4, and MC15 recognize phospho-epitopes; TG-5 recognizes

primary sequence; and MC1, Alz-50, and TG-3 recognize conformational epitopes. (B) A

diagram shows the recognition sites of these antibodies on τ protein (MC15 is not included).

161

Figure 4.2 Recognition of various τ forms with and without dephosphorylation by antibodies to NFT. Three phospho-dependent antibodies, PHF-1, TG-4, and MC15 recognized PHF-τ and normal τ (weaker), dephosphorylation showed significant reduced recognition on PHF-τ but not on normal τ (even also reduced) (B,C). TG-5 is the least influenced antibody, because it recognizes primary sequence (weaker on recombinant τ).

Conformational dependent antibodies MC1, Alz-50, and TG-3 did not recognize all three forms of τ. Relative intensities were compared on the same blots. (*) = p < 0.01.

162

Figure 4.3 Enhancement (fold increase) of antibody recognition by HNE modification. HNE treatment of normal τ enhances antibody recognition capacity by

4-34 fold, an effect was reduced significantly by prior dephosphorylation (B).

Recombinant τ, poorly phosphorylated, shows no HNE enhancement (A). Similarly, hyperphosphorylated τ, PHF-τ, shows no HNE enhancement in the presence or absence of phosphate. Relative intensities were compared on the same blots and with self-comparison for each antibody. (**) = p < 0.05, (*) = p < 0.01 (C).

163

Figure 4.4 Immunocytochemistry on adjacent serial sections (landmark, senile plaque (large arrows)) with the antibodies to NFT. All these antibodies recognized NFT, dystrophic neuritis of senile plaques, and neuropil threads. Scale bar: 50µm.

164 DISCUSSION

In this study, we analyzed seven antibodies against PHF-τ that recognize τ

phospho-epitopes, τ primary sequence, or τ conformational changes. They share the property of enhancement of recognition capacity by HNE modifications suggesting that carbonyl-modified τ may be a major antigenic determinant of NFT. Furthermore, the apparent requirement of phosphorylation for the reaction of HNE with τ is consistent with previous cytological and biochemical studies showing that increased phosphorylation precedes ubiquitination of τ and NFT formation (Bancher et al., 1989, 1991; Augustinack et al., 2002). The data support the idea that phosphorylation of τ is required for its modification by carbonyls (Figures 4.2 and Figure 4.3).

τ in NFT (PHF-τ) is abnormally hyperphosphorylated (Grundke-Iqbal et al., 1986;

Trojanowski and Lee, 1995; Lee et al., 2001), ubiquitinated (Iqbal and Grundke-Iqbal,

1997; Mori et al 1987; Perry et al., 1987), proteolytically-processed, aggregated

(Lovestone and McLoughlin, 2002; Perez et al., 2002), and modified by products of oxidative stress (Mattson et al., 1995; Takeda et al., 2000a). However, how the τ isomers aggregate into filament is still an open question. In a prior study, our group showed that Alz50, a well known NFT epitope, is induced by HNE modification of normally phosphorylated τ (Takeda et al., 2000a). Additionally, HNE, a carbonyl formed from lipid peroxidation, can facilitate phosphorylated but not non-phosphorylated or less phosphorylated τ isomer, to polymerize into PHF-like filaments both in vitro

(Perez et al., 2000) and in cells (Perez et al., 2002).

In a previous study, Jicha and colleagues (1997a) showed that Alz50 and MC-1 epitopes can be produced by the immunoblotting process for normal τ. Our group

165 showed the same result (Takeda et al., 2000a), so, in this study, we used dotblots instead

of immunoblots.

HNE can modify the nucleophilic side chains of the amino acids cysteine,

histidine, and lysine, through formation of Michael adducts, lysine-derived pyrroles, and

crosslinks (Xu et al., 1999; Liu et al., 2003). These structures influence the properties of

proteins through different mechanisms, such as by crosslinking, changing hydrophobicity

or altering protein structure, which will also influence the related antigen reactions with

antibodies.

From this point, one possible explanation for the HNE enhancement of antibody

recognition is that NFT-like epitopes exist at a very low level in normal τ but that the

modifications induced by HNE can stabilize these epitopes in a specific way. This

HNE-modified phospho-τ form can preserve higher levels of these specific epitopes,

which also represent the major NFT epitopes. If this is true, HNE-induced

modifications of phospho-τ may represent the major cause of NFT formation.

Consistent with the hypothesis that τ might exist in an equilibrium of conformational states that can be perturbed, is that 2,2,2-trifluoroethanol, which can stabilize certain local conformations of protein (Lang et al., 1994) includingτ (Hoffmann et al., 1997), promotes formation of the TG-3 epitope in τ (Jicha et al., 1997b).

All the phosphorylation sites of PHF-τ can also be found in normal τ. The difference is that normal τ has around 3 moles of phosphate while PHF-τ has about 11 moles (Ksiezak-Reding et al., 1992). This may explain why the phospho-dependent

NFT antibody recognition capacity to normal τ is lower than to PHF-τ. (Fig 4.2 B,C).

And the same reason can be applied to the dephosphorylation process, which showed that

166 reduction of antibody recognition of PHF-1, TG-4, and MC15 in normal τ is less significant than in PHF-τ after dephosphorylation (Fig 4.2 B,C).

Recombinant τ did not show HNE-induced enhancement, consistent with phosphorylation being required for the formation of NFT epitopes (Fig 4.3A). PHF-τ did not show the HNE-induced enhancement, and the possibilities are that PHF-τ is already extensively modified by carbonyls or has acquired some specific structure to prevent further HNE modification (Fig 4.3C) ..In the prior study with Alz50, we found that the epitope is lost with high molecular weight polymer formation like that found in

NFT. In fact, isolated SDS-insoluble NFT are not recognized by Alz50 (unpublished observation). Interestingly, our purified PHF-τ was not recognized by three conformational dependent NFT antibodies (MC1, Alz50, and TG-3), but all the antibodies recognition were enhanced in normal τ after HNE treatment (Fig 3B), which strongly indicates that all these specific NFT epitopes can be produced directly from

HNE modifications of τ. These results are also supported by the data showing that these antibodies stained NFT in brain samples form the same cases of AD (Fig 4). PHF-τ and normal τ should become indistinguishable from recombinant τ after dephosphorylation and linearization, but dephosphorylation of PHF-τ or normal τ did not render them exactly the same as recombinant τ (Can be noticed in Fig 2 and 3), which suggests that additional changes exist in normal τ and PHF-τ forms compared with recombinant τ.

TG-5 antibody reactivity is the least affected, possibly due to its mapping primarily to the primary sequence of τ.

Our findings raise the very real possibility that the lipoxidation-induced modifications of τ protein are a regulated and vital process of the phosphorylation

167 mediated stress response of neurons prior to formation of NFT. Studies concerning the

link between τ phosphorylation modification and the antioxidant enzyme heme oxygenase-1 show that these processes co-exist, and in vitro transfection studies also show that they are interrelated processes (Takeda et al., 2000b). This evidence suggests that hyperphosphorylation and carbonyl adduction of τ may be a critical response to oxidative insults, which play important homeostatic functions. This idea is supported by data indicating that neurons with hyperphosphorylated τ exhibit a reduction in oxidative stress (Nunomura et al., 2001). Therefore, therapeutic strategies envisaging the reduction of τ hyperphosphorylation may be counterproductive since they potentially may interfere with a normal stress response associated with τ phosphorylation (Smith et al., 2002b).

168

Chapter 5

Discussion and Future Directions

169 In this dissertation, we studied two critical proteins involved in Alzheimer disease and other neurodegenerative diseases—neurofilament heavy subunit (NFH) and Tau protein. They both involve in NFT formation in AD and individually they involves in many other types of neurodegenerative diseases, like ALS, CML, diabetes. As oxidative stress is a major concern related with neurodegenerative diseases in our lab, we further studied one of the lipoxidation markers---HNE modification of these two proteins.

In chapter 1, detailed background introduction about AD, neurofilament protein, tau protein, and oxidative stress were presented. Based on the previous finding in our lab, new evidence are presented that are supportive for the involvement of oxidative stress and these two proteins in neurodegenerative dieseaese.

In chapter 2, the molecular mechanism that regulates NFH-HNE adduct is defined. Our previous study reported that levels of NFH-HNE adducts did not change with age or transport in neuronal axons of human, mice, rats. The question is that what the mechanism is, and the result demonstrated that NFH-HNE modification is a phosphorylation-dependent and sequence-specific modification controlled by neuronal signals. This study provided new evidence that support the NFH protective function reported in other two transgenic mice studies (Couillard-Després et al. 1998; Picklo et al.,

2002). In their studies, they had already concerned about the oxidative stress that may play a role in these transgenic mice models, but they did not provide any further evidences than a hypothesis. In this study, we used HNE as a representative for oxidative stress showed that NFH can act as a scavenger for oxidative stress and the consequential adducts are reversibly controlled at least partially by neurons.

170 As showed in chapter 2, NFH is a very special protein in the neuronal axons, it

not only provides many cytoskeletal functions but also provide protective function. As

we know, NFH involves in neurodegenerative diseases and other diseases, it is necessary

for clearer study with this protein. Because AD is a disease developed over decades and

its correlation with age is more than any other factors. Therefore, as the major aging

theory, free radical modifications should be a direct cause for what happened in AD. Here

we studied HNE modification of neurofilament and we clearly showed an amazing

regulatory mechanism for control of NFH-HNE level in neurons. Therefore, there are

some other possibilities that other similar or different regulations exist in neurons.

Further explore these mechanisms will help us to understand the neuronal regulations and

hopefully we can define the cause of neurodegenerative disease.

HNE as a lipid –peroxidation product is well studied in the last two decades, the

chemical structure of the HNE adduct give us the opportunity to further explore the

cellular changes of the modifications. As the major and only reversible HNE

adduct-lysine HNE adduct. There is a possibility for the neuronal removable of this

modification. Other free radicals like MDA, acrolin, should be studied more as HNE,

then these studies will be able to provide strong support for the cellular studies.

Moreover, these will benefit for deciphering the removal of this adduct in cells. In addition, these studies will help people to understand the free radical damage more clearly and it will facilitate the production of antioxidant drugs and elongate human being lifespan. To understand the oxidative stress better, it will fundamentally improve the understanding of aging process and many related disease. The major killer disease like cardiovascular diseases, cancer, and neurodegenerative diseases, all related with

171 oxidative stress. The oxidative modified biomolecules are the major pathophysiological changes and we need to know how to deal with them.

HNE modification reversibility provides the opportunity for us to understand how the cell can manipulate the oxidative modifications. So many oxidative modification are presented in cells and some will make real damage some may not some harmful.

As opposite to the HNE-Lys-Michael adduct, other HNE adducts are not chemically reversible. In case of formation of those modifications, two possibilities exist.

One is these adducts will accumulate and degraded by proteasome pathway. Another is cellular enzymatically removal of these adducts. How powerful the cells are, we do know yet.

In chapter 3, we studied the effect of HNE treatment in neuronal cells and the effect of overexpression of NFH in neuronal cells. The results showed that HNE can induce increased level of total NFH and phosphorylated NFH, which provide the evidence that NFH may play a major role in the axons against oxidative stress. The activation of different kinases by HNE treatment also showed by many studies (Lee et al.,

2004). We showed that HNE treatment induced the activation of MAPK, which has already been reported as a neurofilament kinase and caused the phosphorylation of neurofilament (Pant et al., 2001). Therefore, signal transduction pathway may have already involved in the neuronal control of HNE modification.

The physiological function of HNE is still need more studies and the clear difine the phyisologocal concentration and pathological concentration of HNE will give a further demonstration about HNE functions in cells. HNE modifications give us the opportunity to further explore the oxidative modifications in cellular level, whether it is

172 possible for us to understand more about the oxidation in cellular level and find a way to

utilize the knowledge we get is still a time consuming question..

Now neurofilament as the most abundant protein in axons performed to maintain

the homeostasis of HNE concentration, this is a example for other similar modifications

or mechanisms, so as regulations are the major cellular characteristics, we need to study

the things more carefully and thoughtfully. That how the cells accommodate most of the

oxidative attacks in the daily bases needs to be found out.

In chapter 4, we studied the HNE modification in the NFT formation and NFT

epitope formation. Here we proved that HNE promote the major epitope formation in

NFT. HNE modification has already been reported to facilitate tau aggregation forming

PHF-like filaments in vitro or in cells (Perez et al., 2000, 2002). No study has shown how

important the HNE modification in NFT formation. Through this study with seven

distinct NFT antibodies, we provide new evidence that HNE can produce most of the

epitopes of NFH, which indicates that HNE may play a major role in the formation of

NFT at least NFT epitopes.

Aggregation of tau protein is a topic that has been intensively studied, but the

conditions that promote tau aggregation is still ambiguous. As the hyperphosphorylated

tau has lower potential to aggaregate in vitro, there must be some actor facilitating tau

polymerized. Several compounds have been reported in facilitating tau aggregation. One

is HNE molecule. As HNE adduct level was found increased in neurofibrillary tangles

and HNE is produced in cells commonly, it is a close connection to propose that HNE may play a role or play a major role in NFT formation. As demonstrated in this studies, distinct NFT antibodies covered all kinds of epitopes in NFT showed universally

173 increased immunoreactivity after HNE treatment, which suggests a more important role

for HNE in NFT formation or at least in NFT epitopes formation.

All the functions of these protein aggregations in neurodegenerative diseases are unknown, and whether they are the cause or the consequences are still in argument. In the point of cell self-protection, it is impossible for a huge aggregation exist in cells. One possibility is that these aggregates are the results of incapable removal of the modifications of these proteins in cells, which leads to the formation of these protein aggregates. Another possibility is that the protein aggregates are actually part of the cellular protection mechanisms. The dilemma should be solved by the extensive research works as soon as possible.

Concerning with the research in this dissertation, the following experiments should be finished in the future. This will include the concerns below:

1. Study the similar peptide sequences that could have the phosphorylation-dependent regulation.

2. Exame the effect of interruption of phosphorylation and dephosphorylation process with different inhibitors, such as okdiatic acid in cell culture.

3. Find out the signaling pathways that can induce neuronal protection from oxidative insults, such as testing different chemicals and biomolecules in cell cultures.

4. Determine the HNE midifcaton in NFT formation and the effect of removing phosphorylation in these aggregates.

5. Mice models that overexpress NFH can be tested for the HNE toxicity in some careful ways.

6. Determine whether NFH aggregation will affect tau aggregation or vise

174 versa.

More cellular or animal studies will provide precise data for the NFH protection in neurons. Moreover, the neuronal situation is so special that need more studies and broad coverage of the knowledge we got. I hope that base on the understanding of these specific characteristics in neurons, some more valuable data can be produced in the recent future.

175 REFERENCES

Abbas N., Lucking C. B., Ricard S., Durr A., Bonifati V., De Michele G., et al. A wide variety of mutations in the parkin gene are responsible for autosomal recessive parkinsonism in Europe. French Parkinson's Disease Genetics Study Group and the European Consortium on Genetic Susceptibility in Parkinson's Disease. Hum. Mol. Genet. 8: 567-574,1999.

Abraha A, Ghoshal N, Gamblin TC, Cryns V, Berry RW, Kuret J, and Binder LI. C-terminal inhibition of tau assembly in vitro and in Alzheimer’s disease. J Cell Sci 113: 3737–3745, 2000.

Abdalla DSP, Monteiro HP, Oliveira JAC et al. Activities of superoxide dismutase and glutathione peroxidase in schizophrenic and manic-depressive patients. Clin Chem 32: 805-7. 1986

Ackerley S., Grierson A. J., Brownlees J., Thornhill P., Anderton B. H., Leigh P. N., et al. Glutamate slows axonal transport of neurofilaments in transfected neurons. J. Cell Biol. 150: 165-176 2000

Ackerley S., Thornhill P., Grierson A. J., Brownlees J., Anderton B. H., Leigh P. N., et al. Neurofilament heavy chain side arm phosphorylation regulates axonal transport of neurofilaments. J. Cell Biol. 161: 489-495 2003

Aebersold R, Mann M. Mass spectrometry-based proteomics. Nature. 422(6928):198-207. 2003

Ahlijanian M. K., Barrezueta N. X., Williams R. D., Jakowski A., Kowsz K. P., McCarthy S., et al. Hyperphosphorylated tau and neurofilament and cytoskeletal disruptions in mice overexpressing human p25, an activator of cdk5. Proc. Natl. Acad. Sci. USA 97: 2910-2915 2000

Al-Chalabi A., Andersen P. M., Nilsson P., Chioza B., Andersson J. L., Russ C., et al. Deletions of the heavy neurofilament subunit tail in amyotrophic lateral sclerosis. Hum. Mol. Genet. 8: 157-164 1999

Al-Chalabi A. and Miller C. C. Neurofilaments and neurological disease. Bioessays 25: 346-355 2003

176 Aldini G, Carini M, Beretta G, Bradamante S, and Facino RM. Carnosine is a quencher of 4-hydroxy-nonenal: through what mechanism of reaction? Biochem Biophys Res Commun 298: 699–706. 2002

Allen B, Ingram E, Takao M, Smith MJ, Jakes R, Virdee K, Yoshida H, Holzer M, Craxton M, Emson PC, Atzori C, Migheli A, Crowther RA, Ghetti B, Spillantini MG, and Goedert M. Abundant tau filaments and nonapoptotic neurodegeneration in transgenic mice expressing human P301S tau protein. J Neurosci 22: 9340–9351. 2002

Alonso A, Zaidi T, Novak M, Grundke-Iqbal I, and Iqbal K. Hyperphosphorylation induces self-assembly of tau into tangles of paired helical filaments/straight filaments. Proc Natl Acad Sci USA 98: 6923–6928. 2001

Alvarez G, Mun˜oz-Montan˜o JR, Satru´ stegui J, Avila J, Bogonez E, and Dı´az-Nido J. Lithium protects cultured neurons against beta-amyloid-induced neurodegeneration. FEBS Lett 453: 260–264. 1999

Alzheimer A, Stelzmann RA, Schnitzlein HN, Murtagh FR. An English translation of Alzheimer's 1907 paper, "Uber eine eigenartige Erkankung der Hirnrinde". Clin Anat.8(6):429-31. 1995

Alzheimer A. Neurolisches Zentralblatt 25:1134. 1906

Ames, B. et al “Oxidants, antioxidants, and the degenerative diseases of aging” Proc Natl Acad Sci USA 90: 7915-22. 1993

Anderson D. Antioxidant defenses against reactive oxygen species causing genetic and other damage. Mut Res 19: 350: 103-8. 1996

Anderson ME. Glutathione and glutathione delivery compounds. Advances Pharmacol 38:65-78. 1997

Andorn AC, Britton RS, Bacon BR. Evidence that lipid peroxidation and total iron are increased in Alzheimer’s brain. Neurobiol Aging 11: 316. 1990

Andreadis A, Broderick JA, and Kosik KS. Relative exon affinities and suboptimal splice site signals lead to non-equivalence of two cassette exons. Nucleic Acids Res 23: 3585–3593, 1995.

Andreadis A, Brown WM, and Kosik KS. Structure and novel exons of the human tau

177 gene. Biochemistry 31: 10626–10633, 1992.

Andreadis A, Wagner BK, Broderick JA, and Kosik KS. A tau promoter region without neuronal specificity. J Neurochem 66: 2257–2263, 1996.

Anuradha, C.D., Kanno, S. & Hirano, S. Oxidative damage to mitochondria is a preliminary step to caspase-3 activation in fluoride-induced apoptosis in HL-60 cells. Free Radic. Biol. Med. 31, 367–373. 2001

Apathy S. Nach welcher richtung hin soll die nervenlehre reformiert werden. Biol. Zentralal. 9: 527-648 1897

Arnold CS, Johnson GV, Cole RN, Dong DL, Lee M, and Hart GW. The microtubule-associated protein tau is extensively modi- fied with O-linked N-acetylglucosamine. J Biol Chem 271: 28741– 28744, 1996.

Arrasate M, Pe´rez M, Armas-Portela R, and Avila J. Polymerization of tau peptides into fibrillar structures. The effect of FTDP-17 mutations. FEBS Lett 446: 199–202, 1999.

Arrasate M, Perez M, and Avila J. Tau dephosphorylation at tau-1 site correlates with its association to cell membrane. Neurochem Res 25: 43–50, 2000.

Arrasate M, Pe´rez M, Valpuesta JM, and Avila J. Role of glycosaminoglycans in determining the helicity of paired helical filaments. Am J Pathol 151: 1115–1122, 1997.

Arriagada PV, Growdon JH, Hedley-Whyte ET, and Hyman BT. Neurofibrillary tangles but not senile plaques parallel duration and severity of Alzheimer’s disease. Neurology 42: 631–639, 1992.

Ashford JW. APOE genotype effects on Alzheimer's disease onset and epidemiology. J Mol Neurosci.;23(3):157-65. Review. 2004

Atamna H, Robinson C, Ingersoll R, Elliott H, and Ames BN. N-t-Butyl hydroxylamine is an antioxidant that reverses age-related changes in mitochondria in vivo and in vitro. FASEB J 15: 2196– 2204, 2001.

Augustinack, J. C.; Schneider, A.; Mandelkow, E.-M.; Hyman, B. T. Specific tau phosphorylation sites correlate with severity of neuronal cytopathology in Alzheimer’s disease. Acta Neuropathol. 103:26-35; 2002.

178 Averback P Unusual particles in motor neuron disease. Arch. Pathol. Lab. Med. 105: 490-493 1981

Baas P. and Brown A. Slow axonal transport:the polymer transport model. Trends Cell Biol. 7: 380-384 1997

Baas P. W. Microtubules and axonal growth. Curr. Opin. Cell Biol. 9: 29-36 1997

Baas PW, Pienkowski TP, Cimbalnik KA, Toyama K, Bakalis S, Ahmad FJ, and Kosik KS. Tau confers drug stability but not cold stability to microtubules in living cells. J Cell Sci 107: 135–143, 1994.

Baker M, Litvan I, Houlden H, Adamson J, Dickson D, Perez- Tur J, Hardy J, Lynch T, Bigio E, and Hutton M. Association of an extended haplotype in the tau gene with progressive supranuclear palsy. Hum Mol Genet 8: 711–715, 1999.

Bancher, C.; Brunner, C.; Lassmann, H.; Budka, H.; Jellinger, K.; Wiche, G.; Seitelberger, F.; Grundke-Iqbal, I.; Iqbal, K.; Wisniewski, H. M. Accumulation of abnormally phosphorylated tau precedes the formation of neurofibrillary tangles in Alzheimer's disease. Brain Res. 477:90-99; 1989.

Bancher, C.; Grundke-Iqbal, I.; Iqbal, K.; Fried, V. A.; Smith, H. T.; Wisniewski, H. M. Abnormal phosphorylation of tau precedes ubiquitination in neurofibrillary pathology of Alzheimer disease. Brain Res. 539:11-18; 1991.

Barghorn S and Mandelkow E. Toward a unified scheme for the aggregation of tau into Alzheimer paired helical filaments. Biochemistry 41: 14885–14896, 2002.

Baudier J, Lee SH, and Cole RD. Separation of the different microtubule-associated tau protein species from bovine brain and their mode II phosphorylation by Ca2_/phospholipid-dependent protein kinase C. J Biol Chem 262: 17584–17590, 1987.

Baudier, J.; Cole, R. D. The phosphorylation of the microtubule-associated tau proteins. In: Perry, G., ed. Advances in Behavioral Biology, volume 34, Alterations in the Neuronal Cytoskeleton in Alzheimer Disease. New York: Plenum Press: 25-32. 1987

Bauer V, Bauer F. Reactive oxygen species as mediators of tissue protection and injury. Gen Physiol Biophys; 18:7-14. 1999

179 Beaulieu J. M. and Julien J. P. Peripherin-mediated death of motor neurons rescued by overexpression of neurofilament NF-H proteins. J. Neurochem. 85: 248-256 2003

Beaulieu J. M., Nguyen M. D. and Julien J. P. Late onset death of motor neurons in mice overexpressing wild-type peripherin. J. Cell Biol. 147: 531-544 1999

Beaulieu J. M., Robertson J. and Julien J. P. Interactions between peripherin and neurofilaments in cultured cells: disruption of peripherin assembly by the NF-M and NF-H subunits. Biochem. Cell. Biol. 77: 41-45 1999

Beal MF. Energetic in the pathogenesis of neurodegenerative disease. Trends Neurosci; 23: 298- 303. 2000

Beckman, K. & Ames, B. “The free radical theory of aging matures” Physiol Rev 78: 548-81. 1998

Berger TM, Polidori MC, Dabbagh A et al. Antioxidant activity of vitamin C in iron -over loaded human plasma. J Biol Chem 272: 15656-60. 1997

Bergeron C., Beric-Maskarel K., Muntasser S., Weyer L., Somerville M. J. and Percy M. E. Neurofilament light and polyadenylated mRNA levels are decreased in amyotrophic lateral sclerosis motor neurons. J. Neuropathol. Exp. Neurol. 53: 221-230 1994

Binder LI, Frankfurter A, and Rebhun LI. The distribution of tau in the mammalian central nervous system. J Cell Biol 101: 1371–1378, 1985.

Block G, Patterson B, Subar A. Fruits, vegetables and cancer prevention: A review of epidemiological evidence. Nutr Cancer 18: 1-29. 1992

Bomont P., Cavalier L., Blondeau F., Ben Hamida C., Belal S., Tazir M., et al. The gene encoding gigaxonin, a new member of the cytoskeletal BTB/kelch repeat family, is mutated in giant axonal neuropathy. Nat. Genet. 26: 370-374 2000

Bordeaux area. Rev Neurol (Paris). Apr;153(3):185-92. 1997

Boveris A, Oshino N, Chance B. The cellular production of hydrogen peroxide. Biochem J 128: 617-30. 1972

Brady RM, Zinkowski RP, and Binder LI. Presence of tau in isolated nuclei from human brain. Neurobiol Aging 16: 479–486, 1995.

180 Brandt R, Leger J, and Lee G. Interaction of tau with the neural plasma membrane mediated by tau’s amino-terminal projection domain. J Cell Biol 131: 1327–1340, 1995.

Bradford R. & Allen H.. Oxidology. Chula Vista CA: R.W. Bradford Foundation, 1. Pp. 64-65. Pierrefiche, G. & Laborit, H. (1995) “Oxygen free radicals, melatonin and aging” Exp Gerontal 30: 213-27. 997

Breen K. C. and Anderton B. H. (1991) Temporal expression of neurofilament polypeptides in differentiating neuroblastoma cells. Neuroreport 2: 21-24

Breitner and Folstein. Familial nature of Alzheimer's disease. N Engl J Med. Jul 19;311(3):192. 1984

Bre MH and Karsenti E. Effects of brain microtubule-associated proteins on microtubule dynamics and the nucleating activity of centrosomes. Cell Motil Cytoskeleton 15: 88–98, 1990.

Breteler MM. Vascular risk factors for Alzheimer's disease: an epidemiologic perspective. Neurobiol Aging. 21(2):153-60. Review2000.

Brion JP, Tremp G, and Octave JN. Transgenic expression of the shortest human tau affects its compartmentalization and its phosphorylation as in the pretangle stage of Alzheimer’s disease. Am J Pathol 154: 255–270, 1999.

Brownlees J, Ackerley S, Grierson AJ, Jacobsen NJ, Shea K, Anderton BH, Leigh PN, Shaw CE, Miller CC. Charcot-Marie-Tooth disease neurofilament mutations disrupt neurofilament assembly and axonal transport Hum Mol Genet;11(23):2837-44, 2002.

Bruijn, L. I., Beal, M. F., Becher, M. W., Schulz, J. B., Wong, P. C., Price, D. L. & Cleveland, D. W. Proc. Natl. Acad. Sci. USA 94, 7606-7611 1997

Bu B, Klunemann H, Suzuki K, Li J, Bird T, Jin LW, and Vincent I. Niemann-pick disease type C yields possible clue for why cerebellar neurons do not form neurofibrillary tangles. Neurobiol Dis 11: 285–297, 2002.

Buee L, Bussiere T, Buee-Scherrer V, Delacourte A, and Hof PR. Tau protein isoforms, phosphorylation and role in neurodegenerative disorders. Brain Res Brain Res Rev 33: 95–130, 2000.

Bugiani O, Murrell JR, Giaccone G, Hasegawa M, Ghigo G, Tabaton M, Morbin M,

181 Primavera A, Carella F, Solaro C, Grisoli M, Savoiardo M, Spillantini MG, Tagliavini F, Goedert M, and Ghetti B. Frontotemporal dementia and corticobasal degeneration in a family with a P301S mutation in tau. J Neuropathol Exp Neurol 58: 667–677, 1999.

Buki A., Siman R., Trojanowski J. Q. and Povlishock J. T. The role of calpain-mediated spectrin proteolysis in traumatically induced axonal injury. J. Neuropathol. Exp. Neurol. 58: 365-375 1999.

Caceres A and Kosik KS. Inhibition of neurite polarity by tau antisense oligonucleotides in primary cerebellar neurons. Nature 343: 461–463, 1990.

Caceres A, Potrebic S, and Kosik KS. The effect of tau antisense oligonucleotides on neurite formation of cultured cerebellar macroneurons. J Neurosci 11: 1515–1523, 1991.

Cajal S. R. Embryogenesis of neurofibrils. Trab. Lab Investigaciones Biol. Univ. Madrid 2: 219-225 1903.

Carden M. J., Trojanowski J. Q., Schlaepfer W. W. and Lee V. M. Two-stage expression of neurofilament polypeptides during rat neurogenesis with early establishment of adult phosphorylation patterns. J. Neurosci. 7: 3489-3504 1987

Carmel G, Mager EM, Binder LI, and Kuret J. The structural basis of monoclonal antibody Alz50’s selectivity for Alzheimer’s disease pathology. J Biol Chem 271: 32789–32795, 1996.

Carpenter S. Proximal axonal enlargement in motor neuron disease. Neurology 18: 841-851 1968

Castellani RJ, Harris PLR, Sayre LM, Fujii J, Taniguchi N, Vitek MP, Founds H, Atwood CS, Perry G, Smith MA. Advanced glycation in neurofibrillary pathology of Alzheimer disease: Nε-(carboxymethyl) lysine and hexitol-lysine. Free Radic Biol Med 31:175-180, 2001.

Carter J., Gragerov A., Konvicka K., Elder G., Weinstein H. and Lazzarini R. A. Neurofilament (NF) assembly; divergent characteristics of human and rodent NF-L subunits. J. Biol. Chem. 273: 5101-5108 1998

Checler F. Presenilins: multifunctional proteins involved in Alzheimer's disease pathology. IUBMB Life. 48(1):33-9. Review. 1999

182 Chen J, Kanai Y, Cowan NJ, and Hirokawa N. Projection domains of MAP2 and tau determine spacings between microtubules in dendrites and axons. Nature 360: 674–677, 1992.

Cheng, J.Z., Singhal, S.S., Sharma, A., Saini, M., Yang, Y., Awasthi, S., Zimniak, P. & Awasthi, Y.C. Transfection of mGSTA4 in HL-60 cells protects against 4-hydroxynonenalinduced apoptosis by inhibiting JNK-mediated signaling. Arch. Biochem. Biophys. 392, 197–207. 2001

Chin S. S., Macioce P. and Liem R. K. (1991) Effects of truncated neurofilament proteins on the endogenous intermediate filaments in transfected fibroblasts. J. Cell Sci. 99: 335-350

Ching G. Y. and Liem R. K. (1993) Assembly of type IV neuronal intermediate filaments in nonneuronal cells in the absence of preexisting cytoplasmic intermediate filaments. J. Cell Biol. 122: 1323-1335

Ching G. Y. and Liem R. K. (1999) Analysis of the roles of the head domains of type IV rat neuronal intermediate filament proteins in filament assembly using domain-swapped chimeric proteins. J. Cell Sci. 112: 2233-2240

Chirita CN, Necula M, and Kuret J. Anionic micelles and vesicles induce tau fibrillization in vitro. J Biol Chem 278: 25644–25650, 2003.

Chou C. F., Smith A. J. and Omary M. B. (1992) Characterization and dynamics of O-linked glycosylation of human cytokeratin 8 and 18. J. Biol. Chem. 267: 3901-3906

Chou S. M., Wang H. S., Taniguchi A. and Bucala R. Advanced glycation endproducts in neurofilament conglomeration of motoneurons in familial and sporadic amyotrophic lateral sclerosis. Mol. Med. 4: 324-332 1998

Choudhary, S., Zhang, W., Zhou, F., Campbell, G.A., Chan, L.L., Thompson, E.B. & Ansari, N.H. Cellular lipid peroxidation end-products induce apoptosis in human lens epithelial cells. Free. Radic. Biol. Med. 32, 360–369. 2002

Cho JH and Johnson GV. Glycogen synthase kinase 3beta (GSK3 beta) phosphorylates tau at both primed and unprimed sites: differential impact on microtubule binding. J Biol Chem 278: 187–193, 2002.

Christen Y. Oxidative stress and Alzheimer disease.Am J Clin Nutr. 71(2):621S-629S. Review. 2000

183 Clark LN, Poorkaj P, Wszolek Z, Geschwind DH, Nasreddine ZS, Miller B, Li D, Payami H, Awert F, Markopoulou K, Andreadis A, DSouza I, Lee VMY, Reed L, Trojanowski JQ, Zhukareva V, Bird T, Schellenberg G, and Wilhelmsen KC. Pathogenic implications of mutations in the tau gene in pallidoponto- nigral degeneration and related neurodegenerative disorders linked to chromosome 17. Proc Natl Acad Sci USA 95: 13103–13107, 1998.

Cleveland DW, Monteiro MJ, Wong PC, Gill SR, Gearhart JD and Hoffman PN Involvement of neurofilaments in the radial growth of axons. J Cell Sci 15: 85-95,. 1991

Cleveland D. W. and Xu Z. Neurofilaments and motor neuron disease. In: Alzheimer's Disease: Lessons from Cell Biology, pp. 180-192, 1995

Cleveland DW, Hwo SY, and Kirschner MW. Physical and chemical properties of purified tau factor and the role of tau in microtubule assembly. J Mol Biol 116: 227–247, 1977a.

Cleveland DW, Hwo SY, and Kirschner MW. Purification of tau, a microtubule-associated protein that induces assembly of microtubu les from purified tubulin. J Mol Biol 116: 207–225, 1977b.

Cleveland D. W. and Rothstein J. D. From Charcot to Lou Gehrig: deciphering selective motor neuron death in ALS. Nat. Rev. Neurosci. 2: 806-819 2001

Cole J. S., Messing A., Trojanowski J. Q. and Lee. V. M. Modulation of axon diameter and neurofilaments by hypomyelinating Schwann cells in transgenic mice. J. Neurosci. 14: 6956-6966 1994

Collard J. F. and Julien J. P. A simple test to monitor the motor dysfunction in a transgenic mouse model of amyotrophic lateral sclerosis. J. Psychiatry Neurosci. 20: 80-86 1995

Collins F. S., et al. Generation and initial analysis of more than 15,000 full-length human and mouse cDNA sequences. Proc. Natl. Acad. Sci. USA 99: 16899-16903 2002

Contestabile A. Oxidative stresses in neurodegeneration: mechanisms and therapeutic perspectives. Curr Top Med Chem 1 (6): 553-68. 2001

Copp RP, Wisniewski T, Hentatin F et al. Localisation of alphatocopherol transfer protein in the brains of patients with ataxia with vitamin E deficiency and other oxidative stress related neurodegenerative disorders. Brain Res 822 (1-2): 80-7. 1999

184 Correas I, Dı´az-Nido J, and Avila J. Microtubule-associated protein tau is phosphorylated by protein kinase C on its tubulin binding domain. J Biol Chem 267: 15721–15728, 1992.

Correas I, Padilla R, and Avila J. The tubulin-binding sequence of brain microtubule-associated proteins, tau and MAP-2, is also involved in actin binding. Biochem J 269: 61–64, 1990.

Cote F., Collard J. F. and Julien J. P. (1993) Progressive neuronopathy in transgenic mice expressing the human 47. Collot M, Louvard D, and Singer SJ. Lysosomes are associated with microtubules and not with intermediate filaments in cultured fibroblasts. Proc Natl Acad Sci USA 81: 788–792, 1984.

Couillard-Despres S., Zhu Q., Wong P. C., Price D. L., Cleveland D. W. and Julien J. P. Protective effect of neurofilament heavy gene overexpression in motor neuron disease induced by mutant superoxide dismutase. Proc. Natl. Acad. Sci. USA 95: 9626-9630 1998

Craig AM and Banker G. Neuronal polarity. Annu Rev Neurosci 17: 267–310, 1994.

Cras P, Smith MA, Richey PL, Siedlak SL, Mulvihill P, Perry G. Extracellular neurofibrillary tangles reflect neuronal loss and provide further evidence of extensive protein cross-linking in Alzheimer disease. Acta Neuropathol (Berl). 89(4):291-5. 1995

Crowther, R. A.; Olesen, O. F.; Jakes, R.; Goedert, M. The microtubule binding repeats of tau protein assemble into filaments like those found in Alzheimer's disease. FEBS Lett. 309:199-202, 1992.

Cui C., Stambrook P. J. and Parysek L. M. Peripherin assembles into homopolymers in SW13 cells. J. Cell Sci. 108: 3279-3284 1995

David DC, Layfield R, Serpell L, Narain Y, Goedert M, and Spillantini MG. Proteasomal degradation of tau protein. J Neurochem 83: 176–185, 2002.

De Jonghe P., Mersivanova I., Nelis E., Del Favero J., Martin J. J., Van Broeckhoven C., et al. Further evidence that neurofilament light chain gene mutations can cause Charcot-Marie-Tooth disease type 2E. Ann. Neurol. 49: 245-249 2001

Delacourte A, David JP, Sergeant N, Buee L, Wattez A, Vermersch P, Ghozali F, Fallet-Bianco C, Pasquier F, Lebert F, Petit H, and Di Menza C. The biochemical pathway of neurofi- brillary degeneration in aging and Alzheimer’s disease. Neurology

185 52: 1158–1165, 1999.

Delisle M. B. and Carpenter S. Neurofibrillary axonal swellings and amyotrophic lateral sclerosis. J. Neurol. Sci. 63: 241-250 1984

Delisle MB, Murrell JR, Richardson R, Trofatter JA, Rascol O, Soulages X, Mohr M, Calvas P, and Ghetti B. A mutation at codon 279 (N279K) in exon 10 of the Tau gene causes a tauopathy with dementia and supranuclear palsy. Acta Neuropathol 98: 62–77, 1999.

58. DeTure MA, Di Noto L, and Purich DL. In vitro assembly of Alzheimer-like filaments. How a small cluster of charged residues in Tau and MAP2 controls filament morphology. J Biol Chem 277: 34755–34759, 2002.

De Waegh S. M., Lee V. M. and Brady S. T. Local modulation of neurofilament phosphorylation, axonal caliber, and slow axonal transport by myelinating Schwann cells. Cell 68: 451-463 1992.

De Waegh S. and Brady S. T. Altered slow axonal transport and regeneration in a myelin-deficient mutant mouse: the trembler as an in vivo model for Schwann cell-axon interactions. J. Neurosci. 10: 1855-1865 1990.

De Yebenes JG, Sarasa JL, Daniel SE, and Lees AJ. Familial progressive supranuclear palsy. Description of a pedigree and review of the literature. Brain 118: 1095–1103, 1995.

Di Maria E, Tabaton M, Vigo T, Abbruzzese G, Bellone E, Donati C, Frasson E, Marchese R, Montagna P, Munoz DG, Pramstaller PP, Zanusso G, Ajmar F, and Mandich P. Corticobasal degeneration shares a common genetic background with progressive supranuclear palsy. Ann Neurol 47: 374–377, 2000.

Dong D. L., Xu Z. S., Hart G. W. and Cleveland D. W. Cytoplasmic O-GlcNAc modification of the head domain and the KSP repeat motif of the neurofilament protein neurofilament-H. J. Biol. Chem. 271: 20845-20852 1996.

Dong D. L. and Hart G. W. Purification and characterization of an O-GlcNAc selective N-acetyl-beta-D-glucosaminidase from rat spleen cytosol. J. Biol. Chem. 269: 19321-19330 1994.

Dong D. L., Xu Z. S., Chevrier M. R., Cotter R. J., Cleveland D. W. and Hart G.W. Glycosylation of mammalian neurofilaments. Localization of multiple O-linked

186 N-acetylglucosamine moieties on neurofilament polypeptides L and M. J. Biol. Chem. 268: 16679-16687 1993.

Dotti CG, Banker GA, and Binder LI. The expression and distribution of the microtubule-associated proteins tau and microtubule- associated protein 2 in hippocampal neurons in the rat in situ and in cell culture. Neuroscience 23: 121–130, 1987.

Dou F, Netzer WJ, Tanemura K, Li F, Hartl FU, Takashima A, Gouras GK, Greengard P, and Xu H. Chaperones increase association of tau protein with microtubules. Proc Natl Acad Sci USA 100: 721–726, 2003.

Drechsel DN, Hyman AA, Cobb MH, and Kirschner MW. Modulation of the dynamic instability of tubulin assembly by the microtubule- associated protein tau. Mol Biol Cell 3: 1141–1154, 1992.

Drewes G, Lichtenberg KB, Doring F, Mandelkow EM, Biernat J, Goris J, Doree M, and Mandelkow E. Mitogen activated protein (MAP) kinase transforms tau protein into an Alzheimer-like state. EMBO J 11: 2131–2138, 1992.

Drubin DG and Kirschner MW. Tau protein function in living cells. J Cell Biol 103: 2739–2746, 1986.

D’Souza I and Schellenberg GD. Determinants of 4-repeat tau expression. Coordination between enhancing and inhibitory splicing sequences for exon 10 inclusion. J Biol Chem 275: 17700–17709, 2000.

D’Souza I and Schellenberg GD. tau Exon 10 expression involves a bipartite intron 10 regulatory sequence and weak 5 and 3 splice sites. J Biol Chem 277: 6587–26599, 2002.

Dudek SM and Johnson GV. Transglutaminase catalyzes the formation of sodium dodecyl sulfate-insoluble, Alz-50-reactive polymers of tau. J Neurochem 61: 1159–1162, 1993.

Duff K, Knight H, Refolo LM, Sanders S, Yu X, Picciano M, Malester B, Hutton M, Adamson J, Goedert M, Burki K, and Davies P. Characterization of pathology in transgenic mice overexpressing human genomic and cDNA tau transgenes. Neurobiol Dis 7: 87–98, 2000.

Eble AS, Thorpe SR, and Baynes JW. Nonenzymatic glucosylation and glucose-dependent cross-linking of protein. J Biol Chem 258: 9406–9412, 1983.

187 Ebneth A, Godemann R, Stamer K, Illenberger S, Trinczek B, Mandelkow EM, and Mandelkow E. Overexpression of tau protein inhibits kinesin-dependent trafficking of vesicles, mitochondria, and endoplasmic reticulum: implications for Alzheimer’s disease. J Cell Biol 143: 777–794, 1998.

Eidenmuller, J.; Fath, T.; Hellwig, A.; Reed, J.; Sontag, E.; Brandt, R. Structural and functional implications of tau hyperphosphorylation: information from phosphorylation-mimicking mutated tau proteins. Biochemistry 39:13166-13175; 2000.

Elder G. A., Friedrich V. L. Jr., Bosco P., Kang C., Gourov A., Tu P. H., et al. Absence of the mid-sized neurofilament subunit decreases axonal calibers, levels of light neurofilament (NF-L), and neurofilament content. J. Cell Biol. 141: 727-739 1998.

Elder G. A., Friedrich V. L. Jr., Kang C., Bosco P., Gourov A., Tu P. H., et al. Requirement of heavy neurofilament subunit in the development of axons with large calibers. J. Cell Biol. 143: 195-205 1998.

Elder G. A., Friedrich V. L. Jr., Margita A. and Lazzarini R. A. Age-related atrophy of motor axons in mice deficient in the mid-sized neurofilament subunit. J. Cell Biol. 146: 181-192 1999.

Elder G. A., Friedrich V. L. Jr., Pereira D, Tu P. H., Zhang B., Lee V.M., et al. Mice with disrupted midsized and heavy neurofilament genes lack axonal neurofilaments but have unaltered numbers of axonal microtubules. J. Neurosci. Res. 57: 23-32 1999

Elhanany E, Jaffe H, Link WT, Sheeley DM, Gainer H, Pant HC.Identification of endogenously phosphorylated KSP sites in the high-molecular-weight rat neurofilament protein. J Neurochem. 63(6):2324-35. 1994

Engelhart MJ, Geerlings MI, Ruitenberg A, van Swieten JC, Hofman A, Witteman JC, Breteler MM. Dietary intake of antioxidants and risk of Alzheimer disease. JAMA.;287(24):3223-9. 2002

Esterbauer H, Schaur RJ, Zollner H. Chemistry and biochemistry of 4-hydroxynonenal, malonaldehyde and related aldehydes. Free Radic Biol Med. 11(1):81-128. 1991

Evans DA, Funkenstein HH, Albert MS, Scherr PA, Cook NR, Chown MJ, Hebert LE, Hennekens CH, Taylor JO. Prevalence of Alzheimer's disease in a community population of older persons. Higher than previously reported. JAMA. 262(18):2551-6. 1989

188 Eyer J. and Peterson A. Neurofilament-deficient axons and perikaryal aggregates in viable transgenic mice expressing a neurofilament-betagalactosidase fusion protein. Neuron 12: 389-405 1994

Fahn, S. An open trial of high-dosage antioxidants in early Parkinson’s disease. Am J Clin Nutr 53 (1 Suppl): 380S-382S. 1991

Fath T, Eidenmuller J , and Brandt R. Tau-mediated cytotoxicity in a pseudohyperphosphorylation model of Alzheimer’s disease. J Neurosci 22: 9733–9741, 2002.

Fellous A, Francon J, Lennon AM, and Nunez J. Microtubule assembly in vitro. Purification of assembly-promoting factors. Eur J Biochem 78: 167–174, 1977.

Fernyhough P., Gallagher A., Averill S. A., Priestley J. V., Hounsom L., Patel J. et al. Aberrant neurofilament phosphorylation in sensory neurons of rats with diabetic neuropathy. Diabetes 48: 881-889 1999

Ferrante RJ, Browne SE, Shinobu LA et al. Evidence of increased oxidative damage in both sporadic and familial amyotrophic lateral sclerosis. J Neurochem 69 (5): 2064-74 1997

Fields R. The Rapid Determination of Amino Groups with NBS. Methods Enzymology, 25: 464-468, 1972

Figlewicz D. A., Krizus A., Martinoli M. G., Meininger V., Dib M., Rouleau G. A., et al. Variants of the heavy neurofilament subunit are associated with the development of amyotrophic lateral sclerosis. Hum. Mol. Genet. 3: 1757-1761 1994

Fliegner K. H. and Liem R. K. Cellular and molecular biology of neuronal intermediate filaments. Int. Rev. Cytol. 131: 109-167 1991

Fliegner K. H., Kaplan M. P., Wood T. L., Pintar J. E. and Liem R. K. Expression of the gene for the neuronal intermediate filament protein alpha-internexin coincides with the onset of neuronal differentiation in the developing rat nervous system. J. Comp. Neurol. 342: 161-173 1994

Frame S, Cohen P, and Biondi RM. A common phosphate binding site explains the unique substrate specificity of GSK3 and its inactivation by phosphorylation. Mol Cell 7: 1321–1327, 2001.

189 Friede R. and Sarnorajski T. Axon caliber related to neurofilaments and microtubules in sciatic nerve. Anat. Rec. 167: 379-388 1970

Fuchs E. and Weber K. Intermediate filaments: structure, dynamics, function and disease. Annu. Rev. Biochem. 63: 345-382 1994

Fuchs E. and Cleveland D. W. A structural scaffolding of intermediate filaments in health and disease. Science 279: 514-519 1998

Gale CR. Dietary antioxidants and dementia. Int Psychogeriatr 13 (3): 259-62. 2001

Galloway P. G., Mulvihill P. and Perry G. Filaments of Lewy bodies contain insoluble cytoskeletal elements. Am. J. Pathol. 140: 809-822 1992

Gama Sosa MA, Friedrich VL Jr, DeGasperi R, Kelley K, Wen PH, Senturk E, Lazzarini RA, Elder GA. Human midsized neurofilament subunit induces motor neuron disease in transgenic mice. Exp. Neurol. 184: 408-419 2003

Gambetti P et al. In: The Cytoskeleton: A target for Toxic Agents. Plenum Press, New York, pp. 129-142, 1986.

Gamblin, T. C.; King, M. E.; Kuret, J.; Berry, R. W.; Binder, L. I. Oxidative regulation of fatty acid-induced tau polymerization. Biochemistry 39:14203-14210; 2000.

Games D, Adams D, Alessandrini R, Barbour R, Berthelette P, Blackwell C, Carr T, Clemens J, Donaldson T, Gillespie F. Alzheimer-type neuropathology in transgenic mice overexpressing V717F beta-amyloid precursor protein. Nature. 9;373(6514):523-7. 1995

Geddes JF, Hughes AJ, Lees AJ, and Daniel SE. Pathological overlap in cases of parkinsonism associated with neurofibrillary tangles. A study of recent cases of postencephalitic parkinsonism and comparison with progressive supranuclear palsy and Guamanian parkinsonism-dementia complex. Brain 116: 281–302, 1993.

Geisler N. and Weber K. Self-assembly in vitro of the 68,000 molecular weight component of the mammalian neurofilament triplet proteins into intermediate-sized filaments. J. Mol. Biol. 151: 565-571 1981

Georgiou D. M., Zidar J., Korosec M., Middleton L. T., Kyriakides T. and Christodoulou K. A novel NF-L mutation Pro22Ser is associated with CMT2 in a large Slovenian family. Neurogenetics 4: 93-96 2002

190 Ghadge, G. D., Lee, J. P., Bindokas, V. P., Jordan, J., Ma, L., Miller, R. J. & Roos, R. P. Mutant superoxide dismutase-1-linked familial amyotrophic lateral sclerosis: molecular mechanisms of neuronal death and protection. J Neurosci 17 (22): 8756-66. 1997

Ghoshal N, Garcia-Sierra F, Wuu J, Leurgans S, Bennett DA, Berry RW, and Binder LI. Tau conformational changes correspond to impairments of episodic memory in mild cognitive impairment and Alzheimer’s disease. Exp Neurol 177: 475–493, 2002.

Giasson BI, Mushynski WE. Study of proline-directed protein kinases involved in phosphorylation of the heavy neurofilament subunit. J Neurosci. 17(24):9466-72. 1997

Giasson BI, Forman MS, Higuchi M, Golbe LI, Graves CL, Kotzbauer PT, Trojanowski JQ, and Lee VM. Initiation and synergistic fibrillization of tau and alpha-synuclein. Science 300: 636–640, 2003.

Gibb B. J., Robertson J. and Miller C. C. Assembly properties of neurofilament light chain Ser55 mutants in transfected mammalian cells. J. Neurochem. 66: 1306-1311 1996

Gibb B. J., Brion J. P., Brownlees J., Anderton B. H. and Miller C. C. Neuropathological abnormalities in transgenic mice harbouring a phosphorylation mutant neurofilament transgene. J. Neurochem. 70: 492-500 1998

Gill S. R., Wong P. C., Monteiro M. J. and Cleveland D. W. Assembly properties of dominant and recessive mutations in the small mouse neurofilament (NF-L) subunit. J. Cell Biol. 111: 2005-2019 1990

Giunta S and Galeazzi G. Cytochrome c-induces protein-tau aggregates due to the dityrosine crosslinkings (Abstract). Neurobiol Aging 23: S515, 2002.

Goedert M. Filamentous nerve cell inclusions in neurodegenerative diseases: tauopathies and alpha-synucleinopathies. Philos Trans R Soc Lond B Biol Sci 354: 1101–1118, 1999.

Goedert M, Baur CP, Ahringer J, Jakes R, Hasegawa M, Spillantini MG, Smith MJ, and Hill F. PTL-1, a microtubule-associated protein with tau-like repeats from the nematode Caenorhabditis elegans. J Cell Sci 109: 2661–2672, 1996.

Goedert M, Cohen ES, Jakes R, and Cohen P. p42 map kinase phosphorylation sites in microtubule-associated protein tau are dephosphorylated by protein phosphatase 2A1. Implications for Alzheimer’s disease. FEBS Lett 312: 95–99, 1992.

191 Goedert M, Crowther RA, and Spillantini MG. Tau mutations cause frontotemporal dementias. Neuron 21: 955–958, 1998.

Goedert M, Hasegawa M, Jakes R, Lawler S, Cuenda A, and Cohen P. Phosphorylation of microtubule-associated protein tau by stress-activated protein kinases. FEBS Lett 409: 57–62, 1997.

Goedert M, Jakes R, Qi Z, Wang JH, and Cohen P. Protein phosphatase 2A is the major enzyme in brain that dephosphorylates tau protein phosphorylated by proline-directed protein kinases or cyclic AMP-dependent protein kinase. J Neurochem 65: 2804–2807, 1995.

Goedert M, Jakes R, Spillantini MG, Hasegawa M, Smith MJ, and Crowther RA. Assembly of microtubule-associated protein tau into Alzheimer-like filaments induced by sulphated glycosaminoglycans. Nature 383: 550–553, 1996.

Goedert M, Satumtira S, Jakes R, Smith M J, Kamibayashi C,White CL III, and Sontag E. Reduced binding of protein phosphatase 2A to tau protein with frontotemporal dementia and parkinsonism linked to chromosome 17 mutations. J Neurochem 75: 2155–2162, 2000.

Goedert M and Spillantini MG. Tau mutations in frontotemporal dementia FTDP-17 and their relevance for Alzheimer’s disease. Biochim Biophys Acta 1502: 110–121, 2000.

Goedert M, Spillantini MG, Cairns NJ, and Crowther RA. Tau proteins of Alzheimer paired helical filaments: abnormal phosphorylation of all six brain isoforms. Neuron 8: 159–168, 1992.

Goedert M, Spillantini MG, and Crowther RA. Cloning of a big nervous system. Proc Natl Acad Sci USA 89: 1983–1987, 1992.

Goedert M, Spillantini MG, Potier MC, Ulrich J, and Crowther RA. Cloning and sequencing of the cDNA encoding an isoform of microtubule-associated protein tau containing four tandem repeats: differential expression of tau protein mRNAs in human brain. EMBO J 8: 393–399, 1989.

Gomez-Ramos A, Diaz-Nido J, Smith MA, Perry G, and Avila J. Effect of the lipid peroxidation product acrolein on tau phosphorylation in neural cells. J Neurosci Res 71: 863–870, 2003.

Gong CX, Singh TJ, Grundke-Iqbal I, and Iqbal K. Phosphoprotein phosphatase activities

192 in Alzheimer disease brain. J Neurochem 61: 921–927, 1993.

Gong CX, Grundke-Iqbal I, Damuni Z, and Iqbal K. Dephosphorylation of microtubule-associated protein tau by protein phosphatase- 1 and -2C and its implication in Alzheimer disease. FEBS Lett 341: 94–98, 1994.

Gong CX, Singh TJ, Grundke-Iqbal I, and Iqbal K. Alzheimer’s disease abnormally phosphorylated tau is dephosphorylated by protein phosphatase-2B (calcineurin). J Neurochem 62: 803–806, 1994.

Gong CX, Grundke-Iqbal I, and Iqbal K. Dephosphorylation of Alzheimer’s disease abnormally phosphorylated tau by protein phosphatase-2A. Neuroscience 61: 765–772, 1994.

Gong CX, Lidsky T, Wegiel J, Zuck L, Grundke-Iqbal I, and Iqbal K. Phosphorylation of microtubule-associated protein tau is regulated by protein phosphatase 2A in mammalian brain. Implications for neurofibrillary degeneration in Alzheimer’s disease. J Biol Chem 275: 5535–5544, 2000.

Gong C. X., Wang J. Z., Iqbal K. and Grundke-Iqbal I. Inhibition of protein phosphatase 2A induces phosphorylation and accumulation of neurofilaments in metabolically active rat brain slices. Neurosci Lett. 340: 107-110 2003

Gonzalez-Billault C, Engelke M, Jimenez-Mateos EM, Wandosell F, Caceres A, and Avila J. Participation of structural microtubule-associated proteins (MAPs) in the development of neuronal polarity. J Neurosci Res 67: 713–719, 2002.

Good, P. F., Werner, P., Hsu, A., Olanow, C. W., and Perl, D. P. Am. J. Pathol. 149, 21–28 1996

Gotz ME, Kunig G, Reiderer Y. Oxidative stress: Free radical production in neural degeneration. Pharmacol Ther 63 (1): 37-122. 1994

Gordon-Krajcer W, Yang L, and Ksiezak-Reding H. Conformation of paired helical filaments blocks dephosphorylation of epitopes shared with fetal tau except Ser199/202 and Ser202/ Thr205. Brain Res 856: 163–175, 2000.

Gotz J, Probst A, Spillantini MG, Schafer T, Jakes R, Burki K, and Goedert M. Somatodendritic localization and hyperphosphorylation of tau protein in transgenic mice expressing the longest human brain tau isoform. EMBO J 14: 1304–1313, 1995.

193 Gotz J, Tolnay M, Barmettler R, Ferrari A, Burki K, Goedert M, Probst A, and Nitsch RM. Human tau transgenic mice. Towards an animal model for neuro- and glialfibrillary lesion formation. Adv Exp Med Biol 487: 71–83, 2001.

Grant P. and Pant H. C. Neurofilament protein synthesis and phosphorylation. J. Neurocytol. 29: 843-872 2000

Greenberg SG, Davies P, Schein JD, and Binder LI. Hydrofluoric acid-treated tau PHF proteins display the same biochemical properties as normal tau. J Biol Chem 267: 564–569, 1992.

Greenwood JA, Troncoso JC, Costllo AC, Johnson GV. phosphorylation modulate the calsium-mediated proteases and modubulin binding of the 200-KDa and 160 kDa neurofilament proteins. J. Neurochem61,191-199. 1993

Greenwood JA and Johnson GV. Localization and in situ phosphorylation state of nuclear tau. Exp Cell Res 220: 332–337, 1995.

Griffith LM and Pollard TD. The interaction of actin filaments with microtubules and microtubule-associated proteins. J Biol Chem 257: 9143–9151, 1982.

Grover A, Houlden H, Baker M, Adamson J, Lewis J, Prihar G, Pickering-Brown S, Duff K, and Hutton M. 5_ Splice site mutations in tau associated with the inherited dementia FTDP-17 affect a stem-loop structure that regulates alternative splicing of exon 10. J Biol Chem 274: 15134–15143, 1999.

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

Grundke-Iqbal, I.; Iqbal, K. Neuronal cytoskeleton in the biology of Alzheimer disease. Prog. Clin. Biol. Res. 317:745-753; 1989.

Gutteridge JM. Iron promoters of the fenton reaction and lipid peroxidation can be released from hemoglobin by peroxides. FEBS Lett 201: 291-5. 1986

Habeeb A.F.S.A. Determination of Free Amino Groups in Proteins by Trinitrobenzeenesulfonic Acid. Analytical Biochemistry 14:328-336, 1966.

Hall GF, Chu B, Lee G, and Yao J. Human tau filaments induce microtubule and synapse

194 loss in an in vivo model of neurofibrillary degenerative disease. J Cell Sci 113: 1373–1387, 2000.

Haltiwanger R. S., Blomberg M. A. and Hart G. W. Glycosylation of nuclear and cytoplasmic proteins. Purification and characterization of a diphospho-N-acetylglucosamine:polypeptide beta-N-acetylglucosaminyltransferase. J. Biol. Chem. 267: 9005-9013 1992

Halliwell B, Gutteridge JMC. Oxygen toxicity, Oxygen radicals, transition metals and disease. Biochem J 219: 1-14. 1984

Halliwell B, Gutteridge JMC. Oxygen radicals and the nervous system. Trends Neuro Sc 8: 22-6. 1985

Halliwell B, Gutteridge JMC. In free radicals in Biology and Medicine, 2nd Ed, Oxford University Press, Oxford, UK. 1989

Halliwell B, Gutteridge JMC. Role of free radicals and catalytic metal ions in human disease and an overview. Brain Injury 6: 203-12. 1992

Halliwell B. Free radicals, antioxidants and human disease: curiosity, cause or consequences? Lancet 344:721-4. 1994

Halliwell B. Role of free radicals in the neurodegenerative diseases: therapeutic implications for antioxidant treatment. Drugs Aging 18 (9): 685-716. 2001

Haltiwanger R. S., Kelly W. G., Roquemore E. P., Blomberg M. A., Dong L. Y., Kreppel L., et al. Glycosylation of nuclear and cytoplasmic proteins is ubiquitous and dynamic. Biochem. Soc. Trans. 20: 264-269 1992

Hanger DP, Hughes K, Woodgett JR, Brion JP, and Anderton BH. Glycogen synthase kinase-3 induces Alzheimer’s disease-like phosphorylation of tau: generation of paired helical filament epitopes and neuronal localisation of the kinase. Neurosci Lett 147: 58–62, 1992.

Harada A, Oguchi K, Okabe S, Kuno J, Terada S, Ohshima T, Sato-Yoshitake R, Takei Y, Noda T, and Hirokawa N. Altered microtubule organization in small-calibre axons of mice lacking tau protein. Nature 369: 488–491, 1994.

195 Harborne JB. In Flavonoids: Advances in Research Since 1986. Harborne JB, Ed.; Chapman and Hall: London. pp 589-618. 1994

Hardy J and Selkoe JD. The Amyloid Hypothesis of Alzheimer’s Disease: Progress and Problems On the road to Therapeutics. Science 297:353-356, 2002.

Harman, D. “Aging: a theory based on free radical and radiation chemistry” J Gerontal 11: 298-300. 1956

Harman, D. The biologic clock: the mitochondria? J Am Geriatr Soc. 1972 Apr;20(4):145-7 1972

Harman, D. “Free radical theory of aging: the ‘free radical’ diseases” Age 7: 111-31. 1984

Harman, D. “Free radical theory of aging” Mutat Res 275: 257-66. 1992

Hart G. W., Haltiwanger R. S., Holt G. D. and Kelly W. G. Glycosylation in the nucleus and cytoplasm. Annu. Rev. Biochem. 58: 841-874 1989

Hart G. W., Kelly W. G., Blomberg M. A., Roquemore E. P., Dong D. L.-Y., Kreppel L., et al. Colloq. Ges. Biol. Chem. Mosbach 44: 91-103 1993

Hart G. W., Greis K. D., Dong L.-Y. D., Blomberg M. A., Chou T.-Y., Jiang M.-S., et al. Pure Appl. Chem. 67: 1637-1645 1995

Hartmann AM, Rujescu D, Giannakouros T, Nikolakaki E, Goedert M, Mandelkow EM, Gao QS, Andreadis A, and Stamm S. Regulation of alternative splicing of human tau exon 10 by phosphorylation of splicing factors. Mol Cell Neurosci 18: 80–90, 2001.

Hasegawa M, Jakes R, Crowther RA, Lee VMY, Ihara Y, and Goedert M. Characterization of mAb AP422, a novel phosphorylation- dependent monoclonal antibody against tan protein. FEBS Lett 384: 25–30, 1996.

Haynes, R.L., Brune, B. & Townsend, A.J. Apoptosis in RAW 264.7 cells exposed to 4-hydroxy-2-nonenal: dependence on cytochrome c release but not p53 accumulation. Free Radic. Biol. Med. 30, 884–894. 2001

Heins S. and Aebi U. Making heads and tails of intermediate filament assembly, dynamics and networks. Curr. Opin. Cell Biol. 6: 25-33 1994

196 Hellenbrand W, Boeing H, Robra BP et al. Diet and Parkinson’s disease. II: A possible role for the past intake of specific nutrients. Results from a self-administered food-frequency questionnaire in a case-control study. Neurology 47 (3): 644-50. 1996

Hely M. A., Morris J. G., Traficante R., Reid W. G., O'Sullivan D. J. and Williamson P. M. The sydney multicentre study of Parkinson's disease: progression and mortality at 10 years. J. Neurol. Neurosurg. Psychiatry 67: 300-307 1999

Hemnani T, Parihar MS. Reactive oxygen species and oxidative DNA damage. Ind J Physiol Pharmacol 42 (4):44-52. 1998

Herrmann H. and Aebi U. Intermediate filaments and their associates: multi-talented structural elements specifying cytoarchitecture and cytodynamics. Curr. Opin. Cell Biol. 12: 79-90 2000

Herrmann H., Haner M., Brettel M., Ku N. O. and Aebi U. Characterization of distinct early assembly units of different intermediate filament proteins. J. Mol. Biol. 286: 1403-1420 1999

Hernandez F, Borrell J, Guaza C, Avila J, and Lucas JJ. Spatial learning deficit in transgenic mice that conditionally overexpress GSK-3beta in the brain but do not form tau filaments. J Neurochem 83: 1529–1533, 2002.

Hernandez F, Lucas JJ, Cuadros R, and Avila J. GSK-3 dependent phosphoepitopes recognized by PHF-1 and AT-8 antibodies are present in different tau isoforms. Neurobiol Aging. 24(8):1087-94. 2003

Hernandez F, Perez M, Lucas JJ, and Avila J. Are intracellular paired helical filaments (PHFs) from Alzheimer’s disease toxic for a neuron? Brain Aging 1: 12–15, 2001.

Hershko A and Ciechanover A. The ubiquitin system. Annu Rev Biochem 67: 425–479, 1998.

Hetman M, Cavanaugh JE, Kimelman D, and Xia Z. Role of glycogen synthase kinase-3beta in neuronal apoptosis induced by trophic withdrawal. J Neurosci 20: 2567–2574, 2000.

Heutink P. Untangling tau-related dementia. Hum Mol Genet 9: 132. Himmler A. Structure of the bovine tau gene: alternatively spliced transcripts generate a protein family. Mol Cell Biol 9: 1389–1396, 1989.

197 Hill W. D., Arai M., Cohen J. A. and Trojanowski J. Q. Neurofilament mRNA is reduced in Parkinson's disease substantia nigra pars compacta neurons. J. Comp. Neurol. 329: 328-336 1993

Himmler A, Drechsel D, Kirschner MW, and Martin DJ. Tau consists of a set of proteins with repeated C-terminal microtubulebinding domains and variable N-terminal domains. Mol Cell Biol 9: 1381–1388, 1989.

Hirokawa N. The mechanisms of fast and slow transport in neurons: identification and characterization of the new kinesin superfamily motors. Curr. Opin. Neurobiol. 7: 605-614 1997

Hirano A., Donnenfeld H., Sasaki S. and Nakano I. Fine structural observations of neurofilamentous changes in amyotrophic lateral sclerosis. J. Neuropathol. Exp. Neurol. 43: 461-470 1984

Hirokawa N. The mechanisms of fast and slow transport in neurons: identification and characterization of the new kinesin superfamily motors. Curr. Opin. Neurobiol. 7: 605-614 1997

Hisanaga S. and Hirokawa N. Molecular architecture of the neurofilament. II. Reassembly process of neurofilament L protein in vitro. J. Mol. Biol. 211: 871-882 1990

Hisanaga S., Matsuoka Y., Nishizawa K., Saito T., Inagaki M. and Hirokawa N. Phosphorylation of native and reassembled neurofilaments composed of NF-L, NF-M, and NF-H by the catalytic subunit of cAMP-dependent protein kinase. Mol. Biol. Cell 5: 161-172 1994

Hoffman P. N., Griffin J. W. and Price D. L. Control of axonal caliber by neurofilament transport. J. Cell Biol. 99: 705-714 1984

Hoffman P. N. and Lasek R. J. The slow component ofaxonal transport: Identification of major structural polypeptides of the axon and their generality among mammalian neurons. J. Cell Biol. 66: 35 l-366 1975

Hoffmann, R.; Lee, V. M.; Leight, S.; Varga, I.; Otvos, L. Jr. Unique Alzheimer's disease paired helical filament specific epitopes involve double phosphorylation at specific sites. Biochemistry 36:8114-8124; 1997.

198 Holt G. D. and Hart G. W. The subcellular distribution of terminal N-acetylglucosamine moieties. Localization of a novel protein-saccharide linkage, O-linked GlcNAc. J. Biol. Chem. 261: 8049-8057 1986

Hong M, Zhukareva V, Vogelsberg-Ragaglia V, Wszolek Z, Reed L, Miller BI, Geschwind DH, Bird TD, McKeel D, Goate A, Morris JC, Wilhelmsen KC, Schellenberg GD, Trojanowski JQ, and Lee VM. Mutation-specific functional impairments in distinct tau isoforms of hereditary FTDP-17. Science 282: 1914– 1917, 1998.

Hsiao K, Chapman P, Nilsen S, Eckman C, Harigaya Y, Younkin S, Yang F, Cole G. Correlative memory deficits, Abeta elevation, and amyloid plaques in transgenic mice. Science. 4;274(5284):99-102. 1996

Hsieh S. T., Kidd G. J., Crawford T. O., Xu Z., Lin W. M., Trapp B. D., et al. Regional modulation of neurofilament organization by myelination in normal axons. J. Neurosci. 14: 6392-6401 1994

Hutton M, Lendon CL, Rizzu P, Baker M, Froelich S, Houlden H, Pickering-Brown S, Chakraverty S, Isaacs A, Grover A, Hackett J, Adamson J, Lincoln S, Dickson D, Davies P, Petersen RC, Stevens M, de Graaff E, Wauters E, van Baren J, Hillebrand M, Joosse M, Kwon JM, Nowotny P, Heutink P, Che LK, Norton J, Morris JC, Reed LA, Trojanowski J, Basul H, Lannfelt L, Neystar M, Fahn S, Dark F, Tannenberg T, Dodd PR, Hayward N, Kwok JBJ, Schofield PR, Andreadis A, Snowden J, Craufurd D, Neary D, Owen F, Oostra BA, Hardy J, Goate A, van Swieten J, Mann D, Timothy L, and Heutink P. Association of missense and 5_-splice-site mutations in tau with the inherited dementia FTDP-17. Nature 393: 702–705, 1998.

Hwang SC, Jhon DY, Bae YS, Kim JH, and Rhee SG. Activation of phospholipase C-gamma by the concerted action of tau proteins and arachidonic acid. J Biol Chem 271: 18342–18349, 1996.

Ihara Y, Nukina N, Miura R, and Ogawara M. Phosphorylated tau protein is integrated into paired helical filaments in Alzheimer’s disease. J Biochem 99: 1807–1810, 1986.

Ikegami S, Harada A, and Hirokawa N. Muscle weakness, hyperactivity, and impairment in fear conditioning in tau-deficient mice. Neurosci Lett 279: 129–132, 2000.

Imahori K and Uchida T. Physiology and pathology of tau protein kinases in relation to Alzheimer’s disease. J Biochem 121: 179–188, 1997.

Ingram EM and Spillantini MG. Tau gene mutations: dissecting the pathogenesis of

199 FTDP-17. Trends Mol Med 8: 555–562, 2002.

Iqbal K, Zaidi T, Bancher C, and Grundke-Iqbal I. Alzheimer paired helical filaments. Restoration of the biological activity by dephosphorylation. FEBS Lett 349: 104–108, 1994.

Iqal, K.; Grundke-Iqbal, I. Elevated levels of tau and ubiquitin in brain and cerebrospinal fluid in Alzheimer's disease. Int. Psychogeriatr. 9(Suppl 1):289-296; 1997.

Iqbal, K.; Alonso, A. C.; Gong, C. X.; Khatoon, S.; Pei, J. J.; Wang, J. Z.; Grundke-Iqbal, I. Mechanisms of neurofibrillary degeneration and the formation of neurofibrillary tangles. J. Neural Transm. Suppl. 53:169-180; 1998.

Ishihara T, Hong M, Zhang B, Nakagawa Y, Lee MK, Trojanowski JQ, and Lee VM. Age-dependent emergence and progression of a tauopathy in transgenic mice overexpressing the shortest human tau isoform. Neuron 24: 751–762, 1999.

Ishihara T, Zhang B, Higuchi M, Yoshiyama Y, Trojanowski JQ, and Lee VM. Age-dependent induction of congophilic neuro- fibrillary tau inclusions in tau transgenic mice. Am J Pathol 158: 555–562, 2001.

Jackson GR, Wiedau-Pazos M, Sang TK, Wagle N, Brown CA, Massachi S, and Geschwind DH. Human wild-type tau interacts with wingless pathway components and produces neurofibrillary pathology in Drosophila. Neuron 34: 509–519, 2002.

Jacob RA. The integrated antioxidant system. Nutr Res 15 (5): 755-66. 1995

Jacques Alary, Franc_oise Gu_eraud, Jean-Pierre Cravedi Fate of 4-hydroxynonenal in vivo: disposition and metabolic pathways Molecular Aspects of Medicine 24 177–187 2003

Jacomy H., Zhu Q., Couillard-Despres S., Beaulieu J. M. and Julien J. P. Disruption of type IV intermediate filament network in mice lacking the neurofilament medium and heavy subunits. J. Neurochem. 73: 972-984 1999

Jaffe H., Veeranna, Shetty K.T. and Pant H. C. Characterization of the phosphorylation sites of human high molecular weight neurofilament protein by electrospray ionization tandem mass spectrometry and database searching. Biochemistry 37: 3931-3940 1998

200 Jama JW, Launer LJ, Witteman JC et al. Dietary antioxidants and cognitive function in a population-based sample of older persons: the Rotterdam study. Am J Epidemiol 144 (3): 275-80. 1996

Jicha, G. A.; Bowser, R.; Kazam, I. G.; Davies, P. Alz-50 and MC-1, a new monoclonal antibody raised to paired helical filaments, recognize conformational epitopes on recombinant tau. J. Neurosci. Res. 48:128-132; 1997.

Jicha, G. A.; Lane, E.; Vincent, I.; Otvos, L. Jr.; Hoffmann, R.; Davies, P. A conformation- and phosphorylation-dependent antibody recognizing the paired helical filaments of Alzheimer's disease. J. Neurochem. 69:2087-2095; 1997.

Jicha GA, Rockwood JM, Berenfeld B, Hutton M, and Davies P. Altered conformation of recombinant frontotemporal dementia- 17 mutant tau proteins. Neurosci Lett 260: 153–156, 1999.

Joachim, C. L.; Morris, J. H.; Selkoe, D. J.; Kosik, K. S. Tau epitopes are incorporated into a range of lesions in Alzheimer's disease. J. Neuropathol. Exp. Neurol. 46:611-622; 1987.

Johnson GV, Jope RS, and Binder LI. Proteolysis of tau by calpain. Biochem Biophys Res Commun 163: 1505–1511, 1989.

Johnson JV, Greenwood JA, Costello AC, Troncoso . The regulatory role of Calmodulin in the proteolysis of individual neurofilament proteins by calpain. Neurochem Res. 16,869-873. 1991

Jordanova A., De Jonghe P., Boerkoel C. F., Takashima H., De Vriendt E., Ceuterick C., et al. Mutations in the neurofilament light chain gene (NEFL) cause early onset severe Charcot-Marie-Tooth disease. Brain 126: 590-597 2003

Julien JP, Mushynski WE. Neurofilaments in health and disease. Prog Nucleic Acid Res Mol Biol, 61: 1-23, 1998

Julien J. P., Tretjakoff I., Beaudet L. and Peterson A. Expression and assembly of a human neurofilament protein in transgenic mice provide a novel neuronal marking system. Genes Dev. 1: 1085-1095 1987

Julien J. P., Cote F. and Collard J. F. Mice overexpressing the human neurofilament heavy gene as a model of ALS. Neurobiol. Aging 16: 487-492

201 Julien J. P. (1999) Neurofilament functions in health and disease. Curr. Opin. Neurobiol. 9: 554-560 1995

Jung C., Yabe J. T., Lee S. and Shea T. B. Hypophosphorylated neurofilament subunits undergo axonal transport more rapidly than more extensively phosphorylated subunits in situ. Cell Motil. Cytoskeleton 47: 120-129 2000

Jung C., Yabe J. T. and Shea T. B. C-terminal phosphorylation of the high molecular weight neurofilament subunit correlates with decreased neurofilament axonal transport velocity. Brain Res. 856: 12-19 2000

Kampers, T.; Friedhoff, P.; Biernat, J.; Mandelkow, E. M.; Mandelkow, E. RNA stimulates aggregation of microtubule-associated protein tau into Alzheimer-like paired helical filaments. FEBS Lett. 399:344-349; 1996.

Kampfl A., Posmantur R.M., Zhao X., Schmutzhard E., Clifton G.L. and Hayes R.L. Mechanisms of calpain proteolysis following traumatic brain injury: implications for pathology and therapy: a review and update. J. Neurotrauma 14: 121-134 1997

Kaplan M. P., Chin S. S., Fliegner K. H. and Liem R. K. Alpha-internexin, a novel neuronal intermediate filament protein, precedes the low molecular weight neurofilament protein (NF-L) in the developing rat brain. J. Neurosci. 10: 2735-2748 1990

Kearse K. P. and Hart G. W. Lymphocyte activation induces rapid changes in nuclear and cytoplasmic glycoproteins. Proc. Natl. Acad. Sci. USA 88: 1701-1705 1991

Khachaturian ZS. Diagnosis of Alzheimer’s disease. Arch Neurol 42:1097-1105, 1985.

Khatoon S, Grundke-Iqbal I, and Iqbal K. Brain levels of microtubule- associated protein tau are elevated in Alzheimer’s disease: a radioimmuno-slot-blot assay for nanograms of the protein. J Neurochem 59: 750–753, 1992.

Khatoon S, Grundke-Iqbal I, and Iqbal K. Levels of normal and abnormally phosphorylated tau in different cellular and regional compartments of Alzheimer disease and control brains. FEBS Lett 351: 80–84, 1994.

Kidd M. Paired helical filaments in electron microscopy of Alzheimer’s disease. Nature 197: 192–193, 1963.

Kleinsek, D. “Preventing macromolecular cross-linking and aging” Vit Res News 16(7):

202 1, 4-7.169. 2002

Knops J, Kosik KS, Lee G, Pardee JD, Cohen-Gould L, and McConlogue L. Overexpression of tau in a nonneuronal cell induces long cellular processes. J Cell Biol 114: 725–733, 1991.

Knowles R, LeClerc N, and Kosik KS. Organization of actin and microtubules during process formation in tau-expressing Sf9 cells. Cell Motil Cytoskeleton 28: 256–264, 1994.

Komiya Y., Cooper N. A. and Kidman A. D. The long-term effects of a single injection of beta,beta'-iminodipropionitrile on slow axonal transport in the rat. J. Biochem. (Tokyo) 100: 1241-1246 1986

Kong J., Tung V. W., Aghajanian J. and Xu Z. Antagonistic roles of neurofilament subunits NF-H and NF-M against NF-L in shaping dendritic arborization in spinal motor neurons. J. Cell Biol. 140: 1167-1176 1998

Kong J, Xu Z. Overexpression of neurofilament subunit NF-L and NF-H extend survival of a mouse model for amyotrophic lateral sclerosis. Neuroscience Letter 281(1): 72-74, 2000

Kopke E, Tung YC, Shaikh S, Alonso AC, Iqbal K, and Grundke-Iqbal I. Microtubule-associated protein tau. Abnormal phosphorylation of a non-paired helical filament pool in Alzheimer disease. J Biol Chem 268: 24374–24384, 1993.

Kosik KS. Tau: structure and function. In: Brain Microtubule Associated, edited by Avila J, Brandt R, and Kosik KS. Chur, Switzerland: Harwood Academic, p. 43–52. 1997,

Kosik, K. S.; Crandall, J. E.; Mufson, E. J.; Neve, R. L. Tau in situ hybridization in normal and Alzheimer brain: localization in the somatodendritic compartment. Ann. Neurol. 26:352-361; 1989.

Kosik KS, Orecchio LD, Bakalis S, and Neve RL. Developmentally regulated expression of specific tau sequences. Neuron 2: 1389–1397, 1989.

Kriz J., Zhu Q., Julien J.P. and Padjen A. L. Electrophysiological properties of axons in mice lacking neurofilament subunit genes: disparity between conduction velocity and axon diameter in absence of NF-H. Brain Res. 885: 32-44 2000

203 Ksiezak-Reding, H.; Liu, W. K.; Yen, S. H. Phosphate analysis and dephosphorylation of modified tau associated with paired helical filaments. Brain Res. 597:209-219; 1992.

Lang, E.; Szendrei, G. I.; Lee, V. M.; Otvos, L. Jr. Spectroscopic evidence that monoclonal antibodies recognize the dominant conformation of medium-sized synthetic peptides. J. Immunol. Methods 170:103-115; 1994.

Landry F, Lombardo RC, Smith WJ A Method for Application of Samples to Matrix-Assisted Laser Desorption Ionization Time-of-Flight Targets That Enhances Peptide Detection Analyt Bioch 297:1-8,2000

Lariviere RC, Julien JP. Functions of intermediate filaments in neuronal development and disease. J Neurobiol. 58(1):131-48. 2004

Lasek R. J. and Black M. M. How do axons stop growing? Some clues from the metabolism of the proteins in the slow component of axonal transport. In: Mechanisms, Regulation and Special Functions of Protein Synthesis in Brain, Roberts S., Lajtha A. and Gispen W. H. (eds.), pp. 16 l-l 69, Elsevier/North-Holland, New York 1977

Lasek R. J. and Hoffman P. N. The neuronal cytoskeleton, axonal transport and axonal growth. Cold Spring Harbor Conf. Cell Prolif. 3: 1021-1049 1976

Lavedan C., Buchholtz S., Nussbaum R. L., Albin R. L. and Polymeropoulos M. H. A mutation in the human neurofilament M gene in Parkinson’s disease that suggests a role for the cytoskeleton in neuronal degeneration. Neurosci. Lett. 322: 57-61 2002

Ledesma MD, Bonay P, and Avila J. Tau protein from Alzheimer’s disease patients is glycated at its tubulin-binding domain. J Neurochem 65: 1658–1664, 1995.

Ledesma MD, Bonay P, Colac¸o C, and Avila J. Analysis of icrotubule-associated protein tau glycation in paired helical filaments. J Biol Chem 269: 21614–21619, 1994.

Ledesma, M. D.; Perez, M.; Colaco, C.; Avila, J. Tau glycation is involved in aggregation of the protein but not in the formation of filaments. Cell. Mol. Biol. 44:1111-1116; 1998.

Lee G, Cowan N, and Kirschner M. The primary structure and heterogeneity of tau protein from mouse brain. Science 239: 285– 288, 1988.

Lee G, Newman ST, Gard DL, Band H, and Panchamoorthy G. Tau interacts with src-family non-receptor tyrosine kinases. J Cell Sci 111: 3167–3177, 1998.

204 Lee G and Rook SL. Expression of tau protein in non-neuronal cells: microtubule binding and stabilization. J Cell Sci 102: 227–237, 1992.

Lee JY, Je JH, Kim DH, Chung SW, Zou Y, Kim ND, Ae Yoo M, Suck Baik H, Yu BP, Chung HY.Induction of endothelial apoptosis by 4-hydroxyhexenal. Eur J Biochem. 271(7):1339-47. 2004

Lee M. K., Xu Z., Wong P. C. and Cleveland D. W. Neurofilaments are obligate heteropolymers in vivo. J. Cell Biol. 122: 1337-1350 1993

Lee M. K., Marszalek J. R. and Cleveland D. W. A mutant neurofilament subunit causes massive, selective motor neuron death: implications for the pathogenesis of human motor neuron disease. Neuron 13: 975-988 1994

Lee MK, Cleveland DW. Lee, M., Hyun, D. H., Halliwell, B., and Jenner, P. J. Neurochem. 78, 209-220 2001

Lee SC, Kuan CY, Wen ZD, and Yang SD. The naturally occurring PKC inhibitor sphingosine and tumor promoter phorbol ester potentially induce tyrosine phosphorylation/activation of oncogenic proline-directed protein kinase F-A/GSK-3 alpha in a common signalling pathway. J Protein Chem 17: 15–27, 1998.

Lees J. F., Shneidman P. S., Skuntz S. F., Carden M. J. and Lazzarini R. A. The structure and organization of the human heavy neurofilament subunit (NF-H) and the gene encoding it. EMBO J. 7: 1947-1955 1988

Lee V. M., Otvos L. Jr., Schmidt M. L. and Trojanowski J. Q. Alzheimer disease tangles share immunological similarities with multiphosphorylation repeats in the two large neurofilament proteins. Proc. Natl. Acad. Sci. USA 85: 7384-7388 1988

Lee, V. M.; Balin, B. J.; Otvos, L. Jr.; Trojanowski, J. Q. A68: a major subunit of paired helical filaments and derivatized forms of normal Tau. Science 251:675-678; 1991.

Lee VM, Goedert M, and Trojanowski JQ. Neurodegenerative tauopathies. Annu Rev Neurosci 24: 1121–1159, 2001.

Lefebvre T, Ferreira S, Dupont-Wallois L, Bussiere T, Dupire MJ, Delacourte A, Michalski JC, and Caillet-Boudin ML. Evidence of a balance between phosphorylation and O-GlcNAc glycosylation of Tau proteins: a role in nuclear localization. Biochim Biophys Acta 1619: 167–176, 2003.

205 Leibovitz, B. & Siegel, B. “Aspects of free radical reactions in biological systems: aging” J Gerontal 35: 45-56. 1980

Leigh P. N., Dodson A., Swash M., Brion J. P. and Anderton B. H. Cytoskeletal abnormalities in motor neuron disease. An immunocytochemical study. Brain 112: 521-535 1989

Leroy E., Anastasopoulos D., Konitsiotis S., Lavedan C. and Polymeropoulos M. H. Deletions in the Parkin gene and genetic heterogeneity in a Greek family with early onset Parkinson's disease. Hum. Genet. 103: 424-427 1998

Leski M. L., Bao F., Wu L., Qian H., Sun D. and Liu D. Protein and DNA oxidation in spinal injury: neurofilaments--an oxidation target. Free Radic. Biol. Med. 30: 613-624 2001

Levavasseur F., Zhu Q. and Julien J. P. No requirement of α-internexin for nervous system development and for radial growth of axons. Mol. Brain Res. 69: 104-112 1999

Levine RL, Williams JA, Stadtman ER, Shacter E. Carbonyl assays for determination of oxidatively modified proteins. Methods Enzymol. 233:346-57. 1994

Levine SA, Kidd PM, San Leandro CA. Antioxidant adaptation: Its role in Free Radical Pathology. Biocurrents 171-218. 1985

Lew J., Huang Q. Q., Qi Z., Winkfein R. J., Aebersold R., Hunt T., et al. A brain-specific activator of cyclin-dependent kinase 5. Nature 371: 423-426 1994

Lewis S. E. and Nixon R. A. Multiple phosphorylated variants of high molecular weight subunit of neurofilaments of retinal cell neurons: characterization and evidence for their differential association with stationary and moving neurofilaments. J. Cell Biol. 107: 2689-2701 1988

Lewis J, McGowan E, Rockwood J, Melrose H, Nacharaju P, Van Slegtenhorst M, Gwinn-Hardy K, Paul Murphy M, Baker M, Yu X, Duff K, Hardy J, Corral A, Lin WL, Yen SH, Dickson DW, Davies P, and Hutton M. Neurofibrillary tangles, amyotrophy and progressive motor disturbance in mice expressing mutant (P301L) tau protein. Nature Genet 25: 402–405, 2000.

Liem R. K. and Hutchison S. B. Purification of individual components of the neurofilament triplet: filament assembly from the 70 000-dalton subunit. Biochemistry 21: 3221-3226 1982

206 Lim F, Hernandez F, Lucas JJ, Gomez-Ramos P, Moran MA, and Avila J. FTDP-17 mutations in tau transgenic mice provoke lysosomal abnormalities and Tau filaments in forebrain. Mol Cell Neurosci 18: 702–714, 2001.

Lim K. L., Dawson V. L. and Dawson T. M. The genetics of Parkinson's disease. Curr. Neurol. Neurosci. Rep. 2: 439-446 2002

Lindsay J, Laurin D, Verreault R, Hebert R, Helliwell B, Hill GB, McDowell I. Risk factors for Alzheimer's disease: a prospective analysis from the Canadian Study of Health and Aging. Am J Epidemiol. 156(5):445-53. 2002

Lindwall, G.; Cole, R. D. Phosphorylation affects the ability of tau protein to promote microtubule assembly. J. Biol. Chem. 259:5301-5305; 1984.

Lippa CF, Zhukareva V, Kawarai T, Uryu K, Shafiq M, Nee LE, Grafman J, Liang Y, St George-Hyslop PH, Trojanowski JQ, and Lee VM. Frontotemporal dementia with novel tau pathology and a Glu342Val tau mutation. Ann Neurol 48: 850–858, 2000.

Liu F, Iqbal K, Grundke-Iqbal I, and Gong CX. Involvement of aberrant glycosylation in phosphorylation of tau by cdk5 and GSK- 3beta. FEBS Lett 530: 209–214, 2002a.

Liu F, Zaidi T, Iqbal K, Grundke-Iqbal I, and Gong CX. Aberrant glycosylation modulates phosphorylation of tau by protein kinase A and dephosphorylation of tau by protein phosphatase 2A and 5. Neuroscience 115: 829–837, 2002b.

Liu F, Zaidi T, Iqbal K, Grundke-Iqbal I, Merkle RK, and Gong CX. Role of glycosylation in hyperphosphorylation of tau in Alzheimer’s disease. FEBS Lett 512: 101–106, 2002c.

Liu Q., Raina A. K., Smith M. A., Sayre L. M. and Perry G. Hydroxynonenal, toxic carbonyls, and Alzheimer disease. Mol. Aspects Med. 24: 305-313 2003

Lovestone S and Reynolds CH. The phosphorylation of tau: a critical stage in neurodevelopment and neurodegenerative processes. Neuroscience 78: 309–324, 1997.

Lovestone, S.; McLoughlin, D. M. Protein aggregates and dementia: is there a common toxicity? J. Neurol. Neurosurg. Psychiatry 72:152-161; 2002.

Lu PJ, Wulf G, Zhou XZ, Davies P, and Lu K P. The prolylisomerase Pin1 restores the function of Alzheimer-associated phosphorylated tau protein. Nature 399: 784–788,

207 1999.

Lucas JJ, Hernandez F, Gomez-Ramos P, Moran MA, Hen R, and Avila J. Decreased nuclear beta-catenin, tau hyperphosphorylation and neurodegeneration in GSK-3beta conditional transgenic mice. EMBO J 20: 27–39, 2001.

Lucking C. B., Durr A., Bonifati V., Vaughan J., De Michele G., Gasser T., et al. Association between early-onset Parkinson's disease and mutations in the parkin gene. French Parkinson's Disease Genetics Study Group. N. Engl. J. Med. 342: 1560-1567 2000

Maccioni R. B., Otth C., Concha I. I. and Munoz J. P. The protein kinase Cdk5. Structural aspects, roles in neurogenesis and involvement in Alzheimer's pathology. Eur. J. Biochem. 268: 1518-1527 2001

Maccioni RB, Vera JC, Dominguez J, and Avila J. A discrete repeated sequence defines a tubulin binding domain on microtubule- associated protein tau. Arch Biochem Biophys 275: 568–579, 1989.

Ma D. M., Descarries L., Micheva K. D., Lepage Y., Julien J. P. and Doucet G. Severe neuronal losses with age in the parietal cortex and ventrobasal thalamus of mice transgenic for the human NF-L neurofilament protein. J. Comp. Neurol. 406: 433-448 1999

Maillard LC. Action des acides aminé´s sur les sucres; formationdes melanoidines par voie mé´thodique. CR Hebd Sé´ances Acad Sci 154: 66–68, 1912.

Mailliot C, Trojanowski JQ, and Lee VM. Impaired tau protein function following nitration-induced oxidative stress in vitro and in vivo (Abstract). Neurobiol Aging 23: S415, 2002.

Mandelkow E. M., Drewes G., Biernat J., Gustke N., Van Lint J., Vandenheede J. R., et al. Glycogen synthase kinase-3 and the Alzheimer-like state of microtubule-associated protein tau. FEBS Lett. 314: 315-321 1992

Manetto V., Sternberger N. H., Perry G., Sternberger L. A. and Gambetti P. Phosphorylation of neurofilaments is altered in amyotrophic lateral sclerosis. J. Neuropathol. Exp. Neurol. 47: 642-653 1988

Mark, R.J., Lovell, M.A., Markesbery, W.R., Uchida, K. & Mattson, M.P. A role for 4-hydroxynonenal, an aldehydic product of lipid peroxidation, in disruption of ion

208 homeostasis and neuronal death induced by amyloid beta-peptide. J. Neurochem. 68, 255–264. 1997

Markesbery W. R. Oxidative stress hypothesis in Alzheimer's disease. Free Radic. Biol. Med. 23: 134-147 1997

Marszalek J. R., Williamson T. L., Lee M. K., Xu Z., Hoffman P. N., Becher M. W., et al. Neurofilament subunit NF-H modulates axonal diameter by selectively slowing neurofilament transport. J. Cell Biol. 135: 711-724 1996

Mattson, M. P. Degenerative and protective signaling mechanisms in the neurofibrillary pathology of AD. Neurobiol. Aging 16:447-457; 1995.

Mattson MP. Neuronal death and GSK-3beta: a tau fetish? Trends Neurosci 24: 255–256, 2001.

Matus A. Microtubule-associated proteins: their potential role in determining neuronal morphology. Annu Rev Neurosci 11: 29–44, 1988.

Mayer A, Siegel NR, Schwartz AL, and Ciechanover A. Degradation of proteins with acetylated amino termini by the ubiquitin system. Science 244: 1480–1483, 1989.

Mayeux R., Honig L. S., Tang M. X., Manly J., Stern Y., Schupf N., et al. Plasma A[beta]40 and A[beta]42 and Alzheimer's disease: relation to age, mortality, and risk. Neurology 61: 1185-1190 2003

McGeer P. L. and McGeer E. G. Inflammation, autotoxicity and Alzheimer disease. Neurobiol. Aging 22: 799-809 2001

McLachlan D. R., Lukiw W. J., Wong L., Bergeron C. and Bech-Hansen N. T. Selective messenger RNA reduction in Alzheimer's disease. Brain Res. 427: 255-261 1988

McLean W. G. The role of axonal cytoskeleton in diabetic neuropathy. Neurochem. Res. 22: 951-956 1997

Meister A. Minireview: Glutathione- ascorbic acid antioxidant system in animals. J Biol Chem 269: 9397- 9400. 1994

209 Meier J., Couillard-Despres S., Jacomy H., Gravel C. and Julien J. P. Extra neurofilament NF-L subunits rescue motor neuron disease caused by overexpression of the human NF-H gene in mice. J. Neuropathol. Exp. Neurol. 58: 1099-1110 1999

Menzies F. M., Grierson A. J., Cookson M. R., Heath P. R., Tomkins J., Figlewicz D. A., et al. Selective loss of neurofilament expression in Cu/Zn superoxide dismutase (SOD1) linked amyotrophic lateral sclerosis. J. Neurochem. 82: 1118-1128 2002

Merchant C, Tang MX, Albert S, Manly J, Stern Y, Mayeux R. The influence of smoking on the risk of Alzheimer's disease. Neurology. 52(7):1408-12. 1999

Mersiyanova I. V., Perepelov A. V., Polyakov A. V., Sitnikov V. F., Dadali E. L., Oparin R. B., et al. A new variant of Charcot-Marie- Tooth disease type 2 is probably the result of a mutation in the neurofilament-light gene. Am. J. Human Genet. 67: 37-46 2000

Miller ML and Johnson GVW. Transglutaminase cross-linking of the tau protein. J Neurochem 65: 1760–1770, 1995.

Minoura K, Tomoo K, Ishida T, Hasegawa H, Sasaki M, and Taniguchi T. Amphipathic helical behavior of the third repeat fragment in the tau microtubule-binding domain, studied by (1)H NMR spectroscopy. Biochem Biophys Res Commun 294: 210–214, 2002.

Mitchison T and Kirschner M. Cytoskeletal dynamics and nerve growth. Neuron 1: 761–772, 1988.

Mitchell JJ, Paiva M, Heaton MB. Vitamin E and beta-carotene protect against ethanol combined with ischaemia in an embryonic rat hippocampal culture model of fetal alcohol syndrome. Neurosci Lett 263 (2-3): 189-92. 1999

Montejo de Garcini E and Avila J. In vitro conditions for the self-polymerization of the microtubule-associated protein, tau factor. J Biochem 102: 1415–1421, 1987.

Montejo de Garcini E, Carrascosa JL, Correas I, Nieto A, and Avila J. Tau factor polymers are similar to paired helical filaments of Alzheimer’s disease. FEBS Lett 236: 150–154, 1988.

Montejo de Garcini E, Dı´ez JC, and Avila J. Quantitation and characterization of tau factor in porcine tissues. Biochim Biophys Acta 881: 456–461, 1986.

Montejo de Garcini E, Serrano L, and Avila J. Self assembly of microtubule associated

210 protein tau into filaments resembling those found in Alzheimer disease. Biochem Biophys Res Commun 141: 790–796, 1986.

Montine, T. J., Amarnath, V., Martin, M. E., Strittmatter, W. J., and Graham, D. G. Am. J. Pathol. 148, 89–93 1996

Montine, K. S.; Kim, P. J.; Olson, S. J.; Markesbery, W. R.; Montine, T. J. 4-hydroxy-2-nonenal pyrrole adducts in human neurodegenerative disease. J. Neuropathol. Exp. Neurol. 56:866-871; 1997.

Montine, T. J.; Neely, M. D.; Quinn, J. F.; Beal, M. F.; Markesbery, W. R.; Roberts, L. J.; Morrow, J. D. Lipid peroxidation in aging brain and Alzheimer’s disease. Free Radic. Biol. Med. 33:620-626; 2002.

Mori, H.; Kondo, J.; Ihara, Y. Ubiquitin is a component of paired helical filaments in Alzheimer’s disease. Science 235:1641-1644; 1987.

Mori H., Oda M. and Mizuno Y. Cortical ballooned neurons in progressive supranuclear palsy. Neurosci. Lett. 209: 109-112 1996

Morishima KM, Hasegawa M, Takio K, Suzuki M, Titani K, and Ihara Y. Ubiquitin is conjugated with amino-terminally processed tau in paired helical filaments. Neuron 10: 1151–1160, 1993.

Morris J. R. and Lasek R. J. Stable polymers of the axonal cytoskeleton: the axoplasmic ghost. J. Cell Biol. 92: 192-198 1982

Morris J. R. and Lasek R. J. Monomer-polymer equilibria in the axon: direct measurement of tubulin and actin as polymer and monomer in axoplasm. J. Cell Biol. 98: 2064-2076 1984

Morsch R, Simon W, and Coleman PD. Neurons may live for decades with neurofibrillary tangles. J Neuropathol Exp Neurol 58: 188–197, 1999.

Munoz D. G., Greene C., Perl D. P. and Selkoe D. J. Accumulation of phosphorylated neurofilaments in anterior horn motoneurons of amyotrophic lateral sclerosis patients. J. Neuropathol. Exp. Neurol. 47: 9-18 1988

211 Munoz DG, Mann D, Lang AE, Bergeron C, Bigio EH, Litvan I, Bhatia KP, Dickson D, Wood NW, and Hutton M. Corticobasal degeneration and progressive supranuclear palsy share a common tau haplotype. Neurology 56: 1702– 1706, 2001.

Murayama S., Bouldin T. W. and Suzuki K. Immunocytochemical and ultrastructural studies of upper motor neurons in amyotrophic lateral sclerosis. Acta Neuropathol. (Berl) 83: 518-524 1992

Murphey RK and Godenschwege TA. New roles for ubiquitin in the assembly and function of neuronal circuits. Neuron 36: 5–8, 2002.

Murrell JR, Spillantini MG, Zolo P, Guazzelli M, Smith MJ, Hasegawa M, Redi F, Crowther RA, Pietrini P, Ghetti B, and Goedert M. Tau gene mutation G389R causes a tauopathy with abundant pick body-like inclusions and axonal deposits. J Neuropathol Exp Neurol 58: 1207–1226, 1999.

Myers M. W., Lazzarini R. A., Lee V. M., Schlaepfer W. W. and Nelson D. L. The human mid-size neurofilament subunit: a repeated protein sequence and the relationship of its gene to the intermediate filament gene family. EMBO J. 6: 1617-1626 1987

Nakamura Y., Hashimoto R., Kashiwagi Y., Aimoto S., Fukusho E., Matsumoto N., et al. Major phosphorylation site (Ser55) of neurofilament L by cyclic AMP-dependent protein kinase in rat primary neuronal culture. J. Neurochem. 74: 949-959 2000

Nakagawa T., Chen J., Zhang Z., Kanai Y. and Hirokawa N. Two distinct functions of the carboxyl-terminal tail domain of NF-M upon neurofilament assembly: cross-bridge formation and longitudinal elongation of filaments. J. Cell Biol. 129: 411-429 1995

National Library of Medicine's Unified Medical Language System (UMLS), http://ghr.nlm.nih.gov/ghr/glossary/neurofilament, accessed 21 June 2004

Nadkarni DV, Sayre LM. Structural definition of early lysine and histidine adduction chemistry of 4-hydroxynonenal. Chem Res Toxicol. 8(2):284-91. 1995

Nakagawa T., Chen J., Zhang Z., Kanai Y. and Hirokawa N. Two distinct functions of the carboxyl-terminal tail domain of NF-M upon neurofilament assembly: cross-bridge formation and longitudinal elongation of filaments. J. Cell Biol. 129: 411-429 1995

Nixon R. A. and Sihag R. K. Neurofilament phosphorylation: a new look at regulation and function. Trends Neurosci. 14(11):501-6. Review. 1991

212 Neely M. D., Sidell K. R., Graham D. G., and Montine T. J. The Lipid Peroxidation Product 4-Hydroxynonenal Inhibits Neurite Outgrowth, Disrupts Neuronal Microtubules, and Modifies Cellular Tubulin. J Neurochem 72: 2323-2333, 1999.

Nelson PT, Stefansson K, Gulcher J, and Saper CB. Molecular evolution of tau protein: implications for Alzheimer’s disease. J Neurochem 67: 1622–1632, 1996.

Neve RL, Harris P, Kosik KS, Kurnit DM, and Donlon TA. Identification of cDNA clones for the human microtubule-associated protein tau and chromosomal localization of the genes for tau and microtubule-associated protein 2. Brain Res 387: 271–280, 1986.

New Alzheimer’s Treatments That May Ease the Mind. Science: 297: 1260-1262, 2002

Nguyen M. D., Lariviere R. C. and Julien J. P. Reduction of axonal caliber does not alleviate motor neuron disease caused by mutant superoxide dismutase 1. Proc. Natl. Acad. Sci. USA 97: 12306-12311 2000

Nguyen M. D., Lariviere R. C. and Julien J. P. Deregulation of Cdk5 in a mouse model of ALS: toxicity alleviated by perikaryal neurofilament inclusions. Neuron 30: 135-147 2001

Nguyen M. D., Boudreau M., Kriz J., Couillard-Despres S., Kaplan D. R. and Julien J. P. Cell cycle regulators in the neuronal death pathway of amyotrophic lateral sclerosis caused by mutant superoxide dismutase 1. J. Neurosci. 23: 2131-2140 2003

Nixon R. A. and Sihag R. K. Neurofilament phosphorylation: a new look at regulation and function. Trends Neurosci. 14: 501-506 1991

Nixon R. A. and Shea T. B. Dynamics of neuronal intermediate filaments: a developmental perspective. Cell Motil. Cytoskeleton 22: 81-91 1992

Nixon R. A. The regulation of neurofilament protein dynamics by phosphorylation: clues to neurofibrillary pathobiology. Brain Pathol. 3: 29-38 1993

Nixon R. A., Paskevich P. A., Sihag R. K. and Thayer C. Y. Phosphorylation on carboxyl terminus domains of neurofilament proteins in retinal ganglion cell neurons in vivo: influences on regional neurofilament accumulation, interneurofilament spacing, and axon caliber. J. Cell Biol. 126: 1031-1046 1994

213 Nixon R. A. Dynamic behavior and organization of cytoskeletal proteins in neurons: reconciling old and new findings. Bioessays 20: 798-807 1998

Nixon R. A. The slow axonal transport debate. Trends Cell Biol. 8: 100 1998

Nixon R. A. The slow axonal transport of cytoskeletal proteins. Curr. Opin. Cell Biol. 10: 87-92 1998

Nunomura A, Perry G, Pappolla MA, Wade R, Hirai K, Chiba S, Smith MA. RNA oxidation is a prominent feature of vulnerable neurons in Alzheimer's disease. J Neurosci. 19(6):1959-64. 1999

Nunomura A, Perry G, Hirai K, Aliev G, Takeda A, Chiba S, Smith MA. Neuronal RNA oxidation in Alzheimer's disease and Down's syndrome. Ann N Y Acad Sci. 893:362-4. 1999

Nunomura A, Perry G, Aliev G, Hirai K, Takeda A, Balraj EK, Jones PK, Ghanbari H, Wataya T, Shimohama S, Chiba S, Atwood CS, Petersen RB, Smith MA. Oxidative damage is the earliest event in Alzheimer disease. J Neuropathol Exp Neurol. 60(8):759-67. 2001

Oblinger M. M., Brady S. T., McQuarrie I. G. and Lasek R. J. Cytotypic differences in the protein composition of the axonally transported cytoskeleton in mammalian neurons. J. Neurosci. 7: 453-462 1987

Ohara O., Gahara Y., Miyake T., Teraoka H. and Kitamura T. Neurofilament deficiency in quail caused by nonsense mutation in neurofilament-L gene. J. Cell Biol. 121: 387-395 1993

Okabe S. and Hirokawa N. Turnover of fluorescently labelled tubulin and actin in the axon. Nature 343: 479-482 1990

Orgogozo JM, Dartigues JF, Lafont S, Letenneur L, Commenges D, Salamon R, Renaud S, Breteler MB. Wine consumption and dementia in the elderly: a prospective community study in the N(epsilon)-(carboxymethyl) lysine and hexitol-lysine. Free Radic Biol Med. 31(2):175-80. 2001

Ott A, Slooter AJ, Hofman A, van Harskamp F, Witteman JC, Van Broeckhoven C, van Duijn CM, Breteler MM. Smoking and risk of dementia and Alzheimer's disease in a population-based cohort study: the Rotterdam Study. Lancet. 351(9119):1840-3. 1998

214 Otvos, L. Jr.; Feiner, L.; Lang, E.; Szendrei, G. I.; Goedert, M.; Lee, V. M. Monoclonal antibody PHF-1 recognizes tau protein phosphorylated at serine residues 396 and 404. J. Neurosci. Res. 39:669-673; 1994.

Paggi P. and Lasek R. J. Axonal transport of cytoskeletal proteins in oculomotor axons and their residence times in the axon terminals. J. Neurosci. 7: 2397-2411 1987

Panda D, Goode BL, Feinstein SC, and Wilson L. Kinetic stabilization of microtubule dynamics at steady state by tau and microtubule-binding domains of tau. Biochemistry 34: 11117–11127, 1995.

Pant H. C. and Gainer H. Properties of a calcium-activated protease in squid axoplasm which selectively degrades neurofilament proteins. J. Neurobiol. 11: 1-12 1980

Pant H. C., Veeranna Neurofilament phosphorylation. Biochem. Cell Biol. 73: 575-592 1995

Papasozomenos, S. C.; Binder, L. I. Phosphorylation determines two distinct species of Tau in the central nervous system. Cell Motil. Cytoskeleton 8:210-226; 1987.

Pappert EJ, Tangney CC, Goetz CG et al. Alpha tocopherol in the ventricular cerebrospinal fluid of Parkinson’s disease patients: dose-response study and correlations with plasma levels. Neurology 47 (4): 1037-42. 1996

Perez M, Valpuesta JM, Medina M, Montejo de Garcini E, and Avila J. Polymerization of tau into filaments in the presence of heparin: the minimal sequence required for tau-tau interaction. J Neurochem 67: 1183–1190, 1996.

Perez, M.; Valpuesta, J. M.; de Garcini, E. M.; Quintana, C.; Arrasate, M.; Lopez Carrascosa, J. L.; Rabano, A.; Garcia de Yebenes, J.; Avila, J. Ferritin is associated with the aberrant tau filaments present in progressive supranuclear palsy. Am. J. Pathol. 152:1531-1539; 1998.

Perez M, Lim F, Arrasate M, and Avila J. The FTDP-17-linked mutation R406W abolishes the interaction of phosphorylated tau with microtubules. J Neurochem 74: 2583–2589, 2000.

Perez M, Cuadros R, Smith MA, Perry G, and Avila J. Phosphorylated, but not native, tau protein assembles following reaction with the lipid peroxidation product, 4-hydroxy-2-nonenal. FEBS Lett 486: 270–274, 2000.

215 Perez M, Arrasate M, Montejo De Garcini E, Munoz V, and Avila J. In vitro assembly of tau protein: mapping the regions involved in filament formation. Biochemistry 40: 5983–5991, 2001.

Perez M, Hernandez F, Gomez-Ramos A, Smith M, Perry G, and Avila J. Formation of aberrant phosphotau fibrillar polymers in neural cultured cells. Eur J Biochem 269: 1484–1489, 2002.

Perrone Capano C., Pernas-Alonso R. and di Porzio U. Neurofilament homeostasis and motoneurone degeneration. Bioessays 23: 24-33 2001

Perry, G.; Friedman, R.; Shaw, G.; Chau, V. Ubiquitin is detected in neurofibrillary tangles and senile plaque neurites of Alzheimer disease brains. Proc. Natl. Acad. Sci. USA 84:3033-3036; 1987.

Perry G, Lipphardt S, Mulvihill P, Kancherla M, Mijares M, Gambetti P, Sharma S, Maggiora L, Cornette J, Lobl T. Amyloid precursor protein in senile plaques of Alzheimer disease. Lancet. 2(8613):746. 1988

Perry, G.; Mulvihill, P.; Fried, V. A.; Smith, H. T.; Grundke-Iqbal, I.; Iqbal, K. Immunochemical properties of ubiquitin conjugates in the paired helical filaments of Alzheimer disease. J. Neurochem. 52:1523-1528; 1989.

Perry G., Castellani R. J., Hirai K. and Smith M. A. Reactive oxygen species mediate cellular damage in Alzheimer disease. J. Alzheimers Dis. 1: 45-55 1998

Perry G, Nunomura A, Hirai K, Zhu X, Perez M, Avila J, Castellani RJ, Atwood CS, Aliev G, Sayre LM, Takeda A, Smith MA. Is oxidative damage the fundamental pathogenic mechanism of Alzheimer's and other neurodegenerative diseases? Free Radix Biol Med. 33(11):1475-9. Review. 2002

Perry G, Sayre LM, Atwood CS et al. The role of iron and copper in the aetiology of neurodegenerative disorders: therapeutic implications. CNS Drugs 16 (5): 339-52. 2002

Perry G, Cash AD, Smith MA. Alzheimer Disease and Oxidative Stress. J Biomed Biotechnol. 2(3):120-123. 2002

Peter M. Eckl Genotoxicity of HNE Molecular Aspects of Medicine 24 161–165 2003 HNE––signaling pathways leading to its elimination Henry Jay Forman , Dale A. Dickinson, Karen E. Iles Molecular Aspects of Medicine 24 189–194 2003

216 Phillips M, Sabas M, Greenberg J. Increased pentane and carbon disulfide in the breath of patients with schizophrenia. J Clin Pathol 46: 861-4. 1993

Pietta PG. Flavonoids as antioxidants. J Nat Prod 63: 1035-42. 2000

Picklo M. J., Montine T. J., Amarnath V. and Neely M. D. Carbonyl toxicology and Alzheimer's disease. Toxicol. Appl. Pharmacol. 184: 187-197 2002

Picklo MJ, Montine TJ, Amarnath V, Neely MD. Carbonyl toxicology and Alzheimer's disease. Toxicol Appl Pharmacol. 184(3):187-97. Review. 2002

Plassman BL, Havlik RJ, Steffens DC, Helms MJ, Newman TN, Drosdick D, Phillips C, Gau BA, Welsh-Bohmer KA, Burke JR, Guralnik JM, Breitner JC. Documented head injury in early adulthood and risk of Alzheimer's disease and other dementias. Neurology. 55(8):1158-66. 2000

Prahlad V., Helfand B. T., Langford G. M., Vale R. D. and Goldman R. D. (2000) Fast transport of neurofilament protein along microtubules in squid axoplasm. J. Cell Sci. 113: 3939-3946

Prasad KN, Cole WC, Kumar B. Multiple antioxidants in the prevention and treatment of Parkinson’s disease. J Am Coll Nutr 18 (5): 413-23. 1999

Praprotnik D, Smith MA, Richey PL, Vinters HV, Perry G. Plasma membrane fragility in dystrophic neurites in senile plaques of Alzheimer’s disease: an index of oxidative stress. Acta Neuropathol 91:1-5, 1996.

Pratico D, Uryu K, Leight S, Trojanoswki JQ, Lee VM. Increased lipid peroxidation precedes amyloid plaque formation in an animal model of Alzheimer amyloidosis. J Neurosci. 21(12):4183-7 2001

Probst A, Gotz J, Wiederhold KH, Tolnay M, Mistl C, Jaton AL, Hong M, Ishihara T, Lee VM, Trojanowski JQ, Jakes R, Crowther RA, Spillantini MG, Burki K, and Goedert M. Axonopathy and amyotrophy in mice transgenic for human fourrepeat tau protein. Acta Neuropathol 99: 469–481, 2000.

Purkinje J. E. Untersuchungen aus der nerven und himanatomie. Ber. Versamml. deut. Naturforsch. u. Arzte (Prague), p.177 1838

217 Rao M. V., Houseweart M. K., Williamson T. L., Crawford T. O., Folmer J. and Cleveland D. W. Neurofilament-dependent radial growth of motor axons and axonal organization of neurofilaments does not require the neurofilament heavy subunit (NF-H) or its phosphorylation. J. Cell Biol. 143: 171-181 1998

Rao MV, Garcia ML, Miyazaki Y, Gotow T, Yuan A, Mattina S, Ward CM, Calcutt NA, Uchiyama Y, Nixon RA and Cleveland DW Gene replacement in mice reveals that the heavily phosphorylated tail of neurofilament heavy subunit does not affect axonal caliber or the transit of cargoes in slow axonal transport. J. Cell Biol. 158: 681-693 2002

Rao M. V., Campbell J., Yuan A., Kumar A., Gotow T., Uchiyama Y. et al. The neurofilament middle molecular mass subunit carboxyl-terminal tail domains is essential for the radial growth and cytoskeletal architecture of axons but not for regulating neurofilament transport rate. J. Cell Biol. 163: 1021-1031(2003

Rapoport M, Dawson HN, Binder LI, Vitek MP, and Ferreira A. Tau is essential to beta-amyloid-induced neurotoxicity. Proc Natl Acad Sci USA 99: 6364–6369, 2002.

Rebeck GW, Reiter JS, Strickland DK, Hyman BTApolipoprotein E in sporadic Alzheimer's disease: allelic variation and receptor interactions. Neuron. 11(4):575-80. 1993

Reynolds, C. H.; Betts, J. C.; Blackstock, W. P.; Nebreda, A. R.; Anderton, B. H. Phosphorylation sites on tau identified by nanoelectrospray mass spectrometry: differences in vitro between the mitogen-activated protein kinases ERK2, c-Jun N-terminal kinase and P38, and glycogen synthase kinase-3β. J. Neurochem. 74:1587-1595, 2000.

Richard C, Mahboub S, Delacoute A, Hemon B and Han K. Comparative Studies of the Neurofilament Triplet Protein Peptide mapping by Chemical Cleavage Comp. Biochem. Physiol., 80B:707-712,1985.

Rizzu P, Van Swieten JC, Joosse M, Hasegawa M, Stevens M, Tibben A, Niermeijer MF, Hillebrand M, Ravid R, Oostra BA, Goedert M, van Duijn CM, and Heutink P. High prevalence of mutations in the microtubule-associated protein tau in a population study of frontotemporal dementia in the Netherlands. Am J Hum Genet 64: 414–421, 1999.

Roder HM, Ingram VM. Two novel kinases phosphorylate tau and the KSP site of heavy neurofilament subunits in high stoichiometric ratios. J Neurosci 11:3325-3343, 1991.

218 Rogaev EI, Sherrington R, Rogaeva EA, Levesque G, Ikeda M, Liang Y, Chi H, Lin C, Holman K, Tsuda T. Familial Alzheimer's disease in kindreds with missense mutations in a gene on chromosome 1 related to the Alzheimer's disease type 3 gene. Nature. 376(6543):775-8. 1995

Rohn T and Davis M. Caspase cleavage of tau in the Alzheimer’s disease brain (Abstract). Neurobiol Aging 23: S519, 2002.

Roots B. I. Neurofilament accumulation induced in synapses by leupeptin. Science 221: 971-972(1983

Roy S., Coffee P., Smith G., Liem R. K., Brady S. T. and Black M. M. Neurofilaments are transported rapidly but intermittently in axons: implications for slow axonal transport. J. Neurosci. 20: 6849-6861 (2000

Ruben GC, Iqbal K, Grundke II, and Johnson JJ. The organization of the microtubule associated protein tau in Alzheimer paired helical filaments. Brain Res 602: 1–13, 1993.

Ruben GC, Iqbal K, Grundke-Iqbal I, Wisniewski HM, Ciardelli TL, and Johnson JE Jr. The microtubule-associated protein tau forms a triple-stranded left-hand helical polymer. J Biol Chem 266: 22019–22027, 1991.

Sadqi M, Hernandez F, Pan U, Perez M, Schaeberle MD, Avila J, and Munoz V. Alpha-helix structure in Alzheimer’s disease aggregates of tau-protein. Biochemistry 41: 7150–7155, 2002.

Sahara N, Lewis J, DeTure M, McGowan E, Dickson DW, Hutton M, and Yen SH. Assembly of tau in transgenic animals expressing P301L tau: alteration of phosphorylation and solubility. J Neurochem 83: 1498–1508, 2002.

Saggu H, Cooksey J, Dexter D et al. A selective increase in particulate superoxide dismutase activity in parkinsonian substantial mgra. J Neurochem 53: 692-7. 1989

Sakaguchi T., Okada M., Kitamura T. and Kawasaki K. Reduced diameter and conduction velocity of myelinated fibers in the sciatic nerve of a neurofilament-deficient mutant quail. Neurosci. Lett. 153: 65-68(1993

Sanchez I., Hassinger L., Paskevich P. A., Shine H.D. and Nixon R. A. Oligodendroglia regulate the regional expansion of axon caliber and local accumulation of neurofilaments

219 during development independently of myelin formation. J. Neurosci. 16: 5095-5105 (1996

Sanchez I., Hassinger L., Sihag R. K., Cleveland D. W., Mohan P. and Nixon R. A. Local control of neurofilament accumulation during radial growth of myelinating axons in vivo. Selective role of site-specific phosphorylation. J. Cell Biol. 151: 1013-1024(2000

Sano M, Ernesto C, Thomas RG, Klauber MR, Schafer K, Grundman M, Woodbury P, Growdon J, Cotman CW, Pfeiffer E, Schneider LS, Thal LJ. A controlled trial of selegiline, alpha-tocopherol, or both as treatment for Alzheimer's disease. The Alzheimer's Disease Cooperative Study. N Engl J Med. 24;336(17):1216-22. 1997

Saunders AM, Strittmatter WJ, Schmechel D, George-Hyslop PH, Pericak-Vance MA, Joo SH, Rosi BL, Gusella JF, Crapper-MacLachlan DR, Alberts MJ. Association of apolipoprotein E allele epsilon 4 with late-onset familial and sporadic Alzheimer's disease. Neurology. 43(8):1467-72. 1993

Sayre LM, Zelasko DA, Harris PL, Perry G, Salomon RG, Smith MA. 4-Hydroxynonenal-derived advanced lipid peroxidation end products are increased in Alzheimer's disease. J Neurochem. 68(5):2092-7. 1997

Sayre LM, Arora PK, Iyer RS, Salomon RG.Pyrrole formation from 4-hydroxynonenal and primary amines.Chem Res Toxicol. 6(1):19-22. 1993

Sayre L. M., Sha W., Xu G., Kaur K., Nadkarni D., Subbanagounder G., et al. Immunochemical evidence supporting 2-pentylpyrrole formation on proteins exposed to 4-hydroxy-2-nonenal. Chem. Res. Toxicol. 9: 1194-12011996.

Scheuner D, Eckman C, Jensen M, Song X, Citron M, Suzuki N, Bird TD, Hardy J, Hutton M, Kukull W, Larson E, Levy-Lahad E, Viitanen M, Peskind E, Poorkaj P, Schellenberg G, Tanzi R, Wasco W, Lannfelt L, Selkoe D, Younkin S. Secreted amyloid beta-protein similar to that in the senile plaques of Alzheimer's disease is increased in vivo by the presenilin 1 and 2 and APP mutations linked to familial Alzheimer's disease. Nat Med. 2(8):864-70. 1996

Schmechel DE, Saunders AM, Strittmatter WJ, Crain BJ, Hulette CM, Joo SH, Pericak-Vance MA, Goldgaber D, Roses AD. Increased amyloid beta-peptide deposition in cerebral cortex as a consequence of apolipoprotein E genotype in late-onset Alzheimer disease. Proc Natl Acad Sci U S A. 90(20):9649-53. 1993

220 Schmidt M. L., Martin J. A., Lee V. M. Y. and Trojanowski J. Q. (1996) Convergence of Lewy bodies and nenrofibrillary tangles in amygdala neurons of Alzheimer’s disease and Lewy body disorders. Acta Neuropathol. (Berl) 91: 475-481

Schmidt R. E., Beaudet L. N., Plurad S. B. and Dorsey D. A. (1997) Axonal cytoskeletal pathology in aged and diabetic human sympathetic autonomic ganglia. Brain Res. 769: 375-383

Schmidt R, Hayn M, Roob B et al. Plasma antioxidants and cognitive performance in middle-aged and older adults: Results of the Austrian Stroke Prevention Study. J Am Geriat Soc 1998; 46: 1407-10.

Schneider, A.; Biernat, J.; von Bergen, M.; Mandelkow, E.; Mandelkow, E. M. Phosphorylation that detaches tau protein from microtubules (Ser262, Ser214) also protects it against aggregation into Alzheimer paired helical filaments. Biochemistry 38:3549-3558; 1999.

Schweers O, Mandelkow EM, Biernat J, and Mandelkow E. Oxidation of cysteine-322 in the repeat domain of microtubuleassociated protein tau controls the in vitro assembly of paired helical filaments. Proc Natl Acad Sci USA 92: 8463–8467, 1995.

Selkoe DJ. Alzheimer's disease: genotypes, phenotypes, and treatments. Science. 275(5300):630-1. 1997

Selkoe D. J. Toward a comprehensive theory for Alzheimer's disease. Hypothesis: Alzheimer's disease is caused by the cerebral accumulation and cytotoxicity of amyloid beta-protein. Ann. NY Acad. Sci. 924: 17-25(2000

Selkoe DJ (2001) Presenilin, Notch, and the genesis and treatment of Alzheimer's disease. Proc Natl Acad Sci ;98(20):11039-41. Review.

Seshadri S, Wolf PA, Beiser A, Au R, McNulty K, White R, D'Agostino RB. (Lifetime risk of dementia and Alzheimer's disease. The impact of mortality on risk estimates in the Framingham Study. Neurology 49: 1498-1504 1997

Shaw G. Neurofilament proteins. In: The Neuronal Cytoskeleton, pp. 5-74, Burgoyne R. D. (ed.), Wiley, New York 1991

Shaw G. and Weber K. Differential expression of neurofilament triplet proteins in brain development. Nature 298: 277-279 1982

221 Shackelford, R.E., Kaufmann, W.K. & Paules, R.S. Oxidative stress and cell cycle checkpoint function. Free Radic. Biol. Med. 28, 1387–1404. 2000

Shea T. B. (Microtubule motors, phosphorylation and axonal transport of neurofilaments. J. Neurocytol. 29: 873-887 2000

Shecket, G. and Lasek, R.J. (1980). Preparation of neurofilament protein from guinea pig peripheral nerve and spinal cord. J. Neurochem. 35, 1335-1344.

Shepherd C. E., McCann H., Thiel E. and Halliday G. M. Neurofilament-immunoreactive neurons in Alzheimer’s disease and dementia with Lewy bodies. Neurobiol. Dis. 9: 249-257 9.2002

Shelanski ML, Gaskin F, and Cantor CR. Microtubule assembly in the absence of added . Proc Natl Acad Sci USA 70: 765-768, 1973

Sherrington R, Rogaev EI, Liang Y, Rogaeva EA, Levesque G, Ikeda M, Chi H, Lin C, Li G, Holman K. Cloning of a gene bearing missense mutations in early-onset familial Alzheimer's disease. Nature. 375(6534):754-60. 1995

Shu YZ. Recent natural products based drug development: A pharmaceutical industry perspective. J Nat Prod 61:1053-71. 1998

Siedlak, S. L.; Cras, P.; Kawai, M.; Richey, P.; Perry, G. Basic fibroblast growth factor binding is a marker for extracellular neurofibrillary tangles in Alzheimer disease. J. Histochem. Cytochem. 39:899-904; 1991.

Sies H, Stahl W. Vitamin E and C, beta carotene and other carotenoids as antioxidants. Am J Clin Nutr 62 (suppl): 13115S-21S. 1995

Sihag R. K. and Nixon R. A. In vivo phosphorylation of distinct domains of the 70-kilodalton neurofilament subunit involves different protein kinases. J. Biol. Chem. 264: 457-464 1989

Sihag R. K. and Nixon R. A. Identification of Ser-55 as a major protein kinase A phosphorylation site on the 70-kDa subunit of neurofilaments. Early turnover during axonal transport. J. Biol. Chem. 266: 18861-18867 1991

Singh TJ, Zaidi T, Grundkeiqbal I, and Iqbal K. Modulation of GSK-3-catalyzed phosphorylation of microtubule-associated protein tau by non-proline-dependent protein

222 kinases. FEBS Lett 358: 4–8, 1995.

Singh RP, Khanna R, Kaw JL et al. Comparative e ffect of Benzanthrone and 3-Bromobenzanthrone on hepatic xenobiotic metabolism and antioxidative defense system in guinea pigs. Arch Toxicol 77: 94-9. 2003

Smith MA, Taneda S, Richey PL, Miyata S, Yan SD, Stern D, Sayre LM, Monnier VM, Perry G. Advanced Maillard reaction end products are associated with Alzheimer disease pathology. Proc Natl Acad Sci U S A. 91(12):5710-4. 1994a

Smith MA, Rudnicka-Nawrot M, Richey PL, Praprotnik D, Mulvihill P, Miller CA, Sayre LM, Perry G. Carbonyl-related posttranslational modification of neurofilament protein in the neurofibrillary pathology of Alzheimer’s disease. J Neurochem 64:2660-2666, 1995.

Smith MA, Sayre LM, Monnier VM, and Perry G. Radical AGEing in Alzheimer’s disease. Trends Neurosci 18: 172–176, 1995.

Smith MA, Siedlak SL, Richey PL, Nagaraj RH, Elhammer A, Perry G. Quantitative solubilization and analysis of insoluble paired helical filaments from Alzheimer disease. Brain Res 717:99-108, 1996.

Smith, M. A., Perry, G., Richey, P. L., Sayre, L. M., Anderson, V. E., Beal, M. F., and Kowall, N. Nature 382, 120–121 1996

Smith, M. A., Harris, P. L. R., Sayre, L. M., Beckman, J. S., and Perry, G. J. Neurosci. 17, 2653–2657 1997

Smith et al, Journal of Histochemistry and Cytochemistry, 46:731-735, 1998

Smith MA, Perry G. What are the facts and artifacts of the pathogenesis and etiology of Alzheimer disease? J Chem Neuroanat. 16(1):35-41. Review. 1998

Smith M.A. Alzheimer disease. Int Rev Neurobiol. 42:1-54. 1998

Smith MA, Monnier VM, Sayre LM, Perry G. Amyloidosis, advanced glycation end products and Alzheimer disease.Neuroreport. 6(12):1595-6. 1995

Smith, M. A.; Casadesus, G.; Joseph, J. A.; Perry, G. Amyloid-beta and tau serve antioxidant functions in the aging and Alzheimer brain. Free Radic. Biol. Med. 33:1194-1199; 2002.

223

Smith, M. A.; Drew, K. L.; Nunomura, A.; Takeda, A.; Hirai, K.; Zhu, X.; Atwood, C. S.; Raina, A. K.; Rottkamp, C. A.; Sayre, L. M.; Friedland, R. P.; Perry, G. Amyloid-beta, tau alterations and mitochondrial dysfunction in Alzheimer disease: the chickens or the eggs? Neurochem. Int. 40:527-531, 2002.

Sobue K, Agarwal-Mawal A, Li W, Sun W, Miura Y, and Paudel HK. Interaction of neuronal Cdc2-like protein kinase with microtubule- associated protein tau. J Biol Chem 275: 16673–16680, 2000.

Sotelo-Silveira J. R., Calliari A., Kun A., Benech J. C., Sanguinetti C., Chalar C. and Sotelo J. R. Neurofilament mRNAs are present and translated in the normal and severed sciatic nerve. J. Neurosci. Res. 62: 65-74 2000

Spina MB, Cohen G. Dopamine turnover and glutathione oxidation implications for Parkinson’s disease. Proc Natl Acad Sci USA 86: 1398-1400. 1989

Spillantini MG, Murrell JR, Goedert M, Farlow MR, Klug A, and Ghetti B. Mutation in the tau gene in familial multiple system tauopathy with presenile dementia. Proc Natl Acad Sci USA 95: 7737–7741, 1998.

Sternberger L. A. and Sternberger N. H. Monoclonal antibodies distinguish phosphorylated and nonphosphorylated forms of neurofilaments in situ. Proc. Natl. Acad. Sci. USA 80: 6126-6130 1983

Sternberger N. H., Sternberger L. A. and Ulrich J. Aberrant neurofilament phosphorylation in Alzheimer disease. Proc. Natl. Acad. Sci. USA 82: 4274-4276 1985

Sternberger, L.A., ed. Immunocytochemistry. New York: Wiley; 1986.

Strausberg R. L., Feingold E. A., Grouse L. H., Derge J. G., Klausner R. D., Yan SD, Chen X, Schmidt AM, Brett J, Godman G, Zou YS, Scott CW, Caputo C, Frappier T, Smith MA. Glycated tau protein in Alzheimer disease: a mechanism for induction of oxidant stress. Proc Natl Acad Sci U S A. 91(16):7787-911994

Steinert_PM. Intermediate filaments in health and disease. Expe and Mole Med, 28:55- 63,1996.

224 Sternberger L. A. and Sternberger N. H. Monoclonal antibodies distinguish phosphorylated and nonphosphorylated forms of neurofilaments in situ. Proc. Natl. Acad. Sci. USA 80: 6126-6130 1983

Sternberger N. H., Sternberger L. A. and Ulrich J. Aberrant neurofilament phosphorylation in Alzheimer disease. Proc. Natl. Acad. Sci. USA 82: 4274-4276 1985

Sternberger LA. Immunocytochemistry, Wiley, New York. 1986

Stewart WF, Kawas C, Corrada M, Metter EJ. Risk of Alzheimer's disease and duration of NSAID use. Neurology. 48(3):626-32. 1997

St George-Hyslop PH, Tanzi RE, Polinsky RJ, Haines JL, Nee L, Watkins PC, Myers RH, Feldman RG, Pollen D, Drachman D. The genetic defect causing familial Alzheimer's disease maps on chromosome 21. Science. 235(4791):885-90. 1987

Strittmatter WJ, Weisgraber KH, Huang DY, Dong LM, Salvesen GS, Pericak-Vance M, Schmechel D, Saunders AM, Goldgaber D, Roses AD. Binding of human apolipoprotein E to synthetic amyloid beta peptide: isoform-specific effects and implications for late-onset Alzheimer disease. Proc Natl Acad Sci U S A. 90(17):8098-102. 1993

Stohs SJ. The role of free radicals in toxicity and disease. J Basic Clin Physiol Pharmacol 6 (3-4): 205-28. 1995

Sun AY, Simonyi A, Sun GY. The “French Paradox” and beyond: neuroprotective effects of polyphenols. Free Radic Biol Med 32 (4): 314-8. 2002

Szucs K, Ledesma MD, Dombradi V, Gergely P, Avila J, and Friedrich P. Dephosphorylation of tau protein from Alzheimer’s disease patients. Neurosci Lett 165: 175–178, 1994.

Takashima A, Murayama M, Murayama O, Kohno T, Honda T, Yasutake K, Nihonmatsu N, Mercken M, Yamaguchi H, Sugihara S, and Wolozin B. Presenilin 1 associates with glycogen synthase kinase-3 beta and its substrate tau. Proc Natl Acad Sci USA 95: 9637–9641, 1998.

Takeda, A.; Perry, G.; Abraham, N. G.; Dwyer, B. E.; Kutty, R. K.; Laitinen, J. T.; Petersen, R. B.; Smith, M. A. Overexpression of heme oxygenase in neuronal cells, the possible interaction with Tau. J. Biol. Chem. 275:5395-5399; 2000.

225

Takeda, A.; Smith, M. A.; Avila, J.; Nunomura, A.; Siedlak, S. L.; Zhu, X.; Perry, G.; Sayre, L. M. In Alzheimer’s disease, heme oxygenase is coincident with Alz50, an epitope of τ induced by 4-hydroxy-2-nonenal modification. J. Neurochem. 75:1234-1241; 2000.

Takei Y, Teng J, Harada A, and Hirokawa N. Defects in axonal elongation and neuronal migration in mice with disrupted tau and map1b genes. J Cell Biol 150: 989–1000, 2000.

Tanaka Y, Kanai Y, Okada Y, Nonaka S, Takeda S, Harada A, and Hirokawa N. Targeted disruption of mouse conventional kinesin heavy chain, kif5B, results in abnormal perinuclear clustering of mitochondria. Cell 93: 1147–1158, 1998.

Tanemura K, Murayama M, Akagi T, Hashikawa T, Tominaga T, Ichikawa M, Yamaguchi H, and Takashima A. Neurodegeneration with tau accumulation in a transgenic mouse expressing V337M human tau. J Neurosci 22: 133–141, 2002.

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. 235(4791):880-4. 1987

Tatebayashi Y, Miyasaka T, Chui DH, Akagi T, Mishima K, Iwasaki K, Fujiwara M, Tanemura K, Murayama M, Ishiguro K, Planel E, Sato S, Hashikawa T, and Takashima A. Tau filament formation and associative memory deficit in aged mice expressing mutant (R406W) human tau. Proc Natl Acad Sci USA 99: 13896–13901, 2002.

Terwel D, Dewachter L, and Van Leuven F. Axonal transport, tau protein, and neurodegeneration in Alzheimer’s disease. Neuromol Med 2: 151–165, 2002.

Teter B ApoE-dependent plasticity in Alzheimer's disease. J Mol Neurosci. 23(3):167-79. Review. 2004

Tomkins J., Usher P., Slade J. Y., Ince P. G., Curtis A., Bushby K., et al. Novel insertion in the KSP region of the neurofilament heavy gene in amyotrophic lateral sclerosis (ALS). NeuroReport 9: 3967-3970 1998

Torres C. R. and Hart G. W. Topography and polypeptide distribution of terminal N-acetylglucosamine residues on the surfaces of intact lymphocytes. Evidence for O-linked GlcNAc. J. Biol. Chem. 259: 3308-3317 1984

226 Trimmer P. A., Borland M. K., Keeney P. M., Bennett J. P. Jr. and Parker W. D. Jr. Parkinson's disease transgenic mitochondrial cybrids generate Lewy inclusion bodies. J. Neurochem. 88: 800-812 2004

Trinczek B, Biernat J, Baumann K, Mandelkow EM, and Mandelkow E. Domains of tau protein, differential phosphorylation, and dynamic instability of microtubules. Mol Biol Cell 6: 1887–1902, 1995.

Trojanowski J. Q., Mawal-Dewan M., Schmidt M. L., Martin J. and Lee V. M. Localization of the mitogen activated protein kinase ERK2 in Alzheimer's disease neurofibrillary tangles and senile plaque neurites. Brain Res. 618: 333-337 1993

Trojanowski, J. Q.; Lee, V. M. Phosphorylation of paired helical filament tau in Alzheimer's disease neurofibrillary lesions: focusing on phosphatases. FASEB J. 9:1570-1576; 1995.

Troncoso JC, Costello A, Watson AL Jr, and Johnson GV. In vitro polymerization of oxidized tau into filaments. Brain Res 613: 313–316, 1993.

Truong AB, Masters SC, Yang H, and Fu H. Role of the 14–3-3 C-terminal loop in ligand interaction. Proteins 49: 321–325, 2002.

Tseng HC, Lu Q, Henderson E, and Graves DJ. Phosphorylated tau can promote tubulin assembly. Proc Natl Acad Sci USA 96: 9503–9508, 1999.

Tu, P.-H., Elder G., Lazzarini R. A., Nelson D., Trojanowski J. Q. and Lee V. M.-Y. (1995) Overexpression of the human NFM subunit in transgenic mice modifies the level of endogenous NFL and the phosphorylation state of NFH subunits. J. Cell Biol. 129: 1629-1640

Uchida K, Stadtman ER. Covalent attachment of 4-hydroxynonenal to glyceraldehyde-3-phosphate dehydrogenase. A possible involvement of intra- and intermolecular cross-linking reaction. J Biol Chem 268:6388-6393, 1993.

Uchida k. (2003) 4-Hydroxy-2-nonenal: a product and mediator of oxidative stress Progress in Lipid Research 42 (2003) 318–343

Vafai SB and Stock JB. Protein phosphatase 2A methylation: a link between elevated plasma homocysteine and Alzheimer’s Disease. FEBS Lett 518: 1–4, 2002.

227 Valentin G. Uber den verlauf und die letzten enden der nerven. Nova Acta physico-medica Acdemiae Caesareae Leopoldino-Carolinae 18: 51-71 1836

Vanier MT, Duthel S, Rodriguez-Lafrasse C, Pentchev P, and Carstea ED. Genetic heterogeneity in Niemann-Pick C disease: a study using somatic cell hybridization and linkage analysis. Am J Hum Genet 58: 118–125, 1996.

Vanier MT, Rodriguez-Lafrasse C, Rousson R, Gazzah N, Juge MC, Pentchev PG, Revol A, and Louisot P. Type C Niemann- Pick disease: spectrum of phenotypic variation in disruption of intracellular LDL-derived cholesterol processing. Biochim Biophys Acta 1096: 328–337, 1991.

Varani L, Hasegawa M, Spillantini MG, Smith MJ, Murrell JR, Ghetti B, Klug A, Goedert M, and Varani G. Structure of tau exon 10 splicing regulatory element RNA and destabilization by mutations of frontotemporal dementia and arkinsonism linked to chromosome 17. Proc Natl Acad Sci USA 96: 8229–8234, 1999.

Veeranna, Amin N. D., Ahn N. G., Jaffe H., Winters C. A., Grant P., et al. Mitogen-activated protein kinases (Erk1,2) phosphorylate Lys-Ser-Pro (KSP) repeats in neurofilament proteins NF-H and NF-M. J. Neurosci. 18: 4008-4021 1998

Vendemiale G, Grattagliano I, Altomare E. An update on the role of free radicals and antioxidant defense in human disease. In J Clin Lab Res 29: 49-55. 1999

Verpillat P, Ricard S, Hannequin D, Dubois B, Bou J, Camuzat A, Pradier L, Frebourg T, Brice A, Clerget-Darpoux F, Deleuze JF, and Campion D. Is the Saitohin gene involved in neurodegenerative diseases? Ann Neurol 52: 829–832, 2002.

Vincent, I.; Rosado, M.; Davies, P. Mitotic mechanisms in Alzheimer's disease? J. Cell Biol. 132:413-425; 1996.

Vitek MP, Bhattacharya K, Glendening JM, Stopa E, Vlassara H, Bucala R, Manogue K, Cerami A. Advanced glycation end products contribute to amyloidosis in Alzheimer disease. Proc Natl Acad Sci U S A. 91(11):4766-701994

Von Bergen M, Friedhoff P, Biernat J, Heberle J, Mandelkow EM, and Mandelkow E. Assembly of tau protein into Alzheimer paired helical filaments depends on a local sequence motif [(306)VQIVYK(311)] forming beta structure. Proc Natl Acad Sci USA 97: 5129–5134, 2000.

Wandosell F and Avila J. Microtubule-associated proteins present in different

228 developmental stages of Drosophila melanogaster. J Cell Biochem 35: 83–92, 1987.

Wang JZ, Gong CX, Zaidi T, Grundke-Iqbal I, and Iqbal K. Dephosphorylation of Alzheimer paired helical filaments by protein phosphatase-2A and -2B. J Biol Chem 270: 4854–4860, 1995.

Wang JZ, Grundkeiqbal I, and Iqbal K. Glycosylation of microtubule- associated protein tau: an abnormal posttranslational modification in Alzheimer’s disease. Nature Med 2: 871–875, 1996.

Wang L., Ho C. L., Sun D., Liem R. K. and Brown A. (2000) Rapid movement of axonal neurofilaments interrupted by prolonged pauses. Nat. Cell Biol. 2: 137-141

Waring SC, Rocca WA, Petersen RC, O'Brien PC, Tangalos EG, Kokmen E. Postmenopausal estrogen replacement therapy and risk of AD: a population-based study. Neurology. 52(5):965-70. 1999

Watanabe A, Takio K, and Ihara Y. Deamidation and isoaspartate formation in smeared tan in paired helical filaments: unusual properties of the microtubule-binding domain of tau. J Biol Chem 274: 7368–7378, 1999.

Watanabe T, Ihara N, Itoh T, Fujita T, and Sugimoto Y. Deletion mutation in Drosophila ma-l homologous, putative molybdopterin cofactor sulfurase gene is associated with bovine xanthinuria type II. J Biol Chem 275: 21789–21792, 2000.

Wataya, T., Nunomura, A., Smith, M.A., Siedlak, S.L., Harris, P.L.R., Shimohama, S., Szweda, L.I., Kaminski, M.A., Price, D.L., Cleveland, D.W., Sayre, L.M., Perry, G. High molecular weight neurofilament proteins are physiological substrates of adduction by the lipid peroxidation product hydroxynonenal. J. Biol. Chem. 277, 4644-4648. 2002

Watson D. F., Nachtman F. N., Kuncl R. W. and Griffin J. W. Altered neurofilament phosphorylation and beta tubulin isotypes in Charcot-Marie-Tooth disease type 1. Neurology 44: 2383-2387 1994

Weaver, C. L.; Espinoza, M.; Kress, Y.; Davies, P. Conformational change as one of the earliest alterations of tau in Alzheimer's disease. Neurobiol. Aging 21:719-727; 2000.

Wedemeyer WJ, Welker E, and Scheraga HA. Proline cis-trans isomerization and protein folding. Biochemistry 41: 14637–14644, 2002.

229 Weingarten MD, Lockwood AH, Hwo SY, and Kirschner MW. A protein factor essential for microtubule assembly. Proc Natl Acad Sci USA 72: 1858–1862, 1975.

Weiner H. L. and Selkoe D. J. Inflammation and therapeutic vaccination in CNS diseases. Nature 420: 879-884(2002

Weisburger JH. Dietary fat and risk of chronic disease: Mechanistic insights from experimental studies. J Am Dietetic Assoc 97: S16-S23. 1997

Weggen S, Eriksen JL, Das P, Sagi SA, Wang R, Pietrzik CU, Findlay KA, Smith TE, Murphy MP, Bulter T, Kang DE, Marquez-Sterling N, Golde TE, Koo EH. A subset of NSAIDs lower amyloidogenic Abeta42 independently of cyclooxygenase activity. Nature. 414(6860):212-6. 2001

Wilhelmsen KC. The tangled biology of tau. Proc Natl Acad Sci USA 96: 7120–7121, 1999.

Willard M and Simon, C. Modulations of neurofilament axonal transport during the development of rabbit retinal ganglion cells. Cell 35,551-559. 1983

Wille H, Drewes G, Biernat J, Mandelkow EM, and Mandelkow E. Alzheimer-like paired helical filaments and antiparallel dimers formed from microtubule-associated protein tau in vitro. J Cell Biol 118: 573–584, 1992.

Williamson TL, Bruijn LI, Zhu Q, Anderson KL, Anderson SD, Julien JP, Cleveland DW. Absence of neurofilaments reduces the selective vulnerability of motor neurons and slows disease caused by a familial amyotrophic lateral sclerosis-linked superoxide dismutase 1 mutant. Proc Natl Acad Sci U S A. 95(16):9631-6. 1998

Williamson R, Scales T, Clark BR, Gibb G, Reynolds CH, Kellie S, Bird IN, Varndell IM, Sheppard PW, Everall I, and Anderton BH. Rapid tyrosine phosphorylation of neuronal proteins including tau and focal adhesion kinase in response to amyloid- beta peptide exposure: involvement of Src family protein kinases. J Neurosci 22: 10–20, 2002.

Wilson DM and Binder LI. Free fatty acids stimulate the polymerization of tau and amyloid beta peptides. In vitro evidence for a common effector of pathogenesis in Alzheimer’s disease. Am J Pathol 150: 2181–2195, 1997.

Wischik CM, Crowther RA, Stewart M, and Roth M. Subunit structure of paired helical filaments in Alzheimer’s disease. J Cell Biol 100: 1905–1912, 1985.

230 Wischik, C. M.; Novak, M.; Edwards, P. C.; Klug, A.; Tichelaar, W.; Crowther, R. A. Structural characterization of the core of the paired helical filament of Alzheimer disease. Proc. Natl. Acad. Sci. USA 85:4884-4888; 1988.

Wischik CM, Lai RYK, and Harrington CR. Modelling prion-like processing of tau protein in Alzheimer’s disease for pharmaceutical development. In: Brain Microtubule Associated, edited by Avila J, Brandt R, and Kosik KS. Chur, Switzerland: Harwood Academic, p. 185–241. 1997

Wisniewski HM, Narang HK, and Terry RD. Neurofibrillary tangles of paired helical filaments. J Neurol Sci 27: 173–181, 1976.

Wittmann CW, Wszolek MF, Shulman JM, Salvaterra PM, Lewis J, Hutton M, and Feany MB. Tauopathy in Drosophila: neurodegeneration without neurofibrillary tangles. Science 293: 711–714, 2001.

Wong N. K., He B. P. and Strong M. J. Characterization of neuronal intermediate filament protein expression in cervical spinal motor neurons in sporadic amyotrophic lateral sclerosis (ALS). J. Neuropathol. Exp. Neurol. 59: 972-982 2000

Wong P. C. and Cleveland D. W. Characterization of dominant and recessive assembly-defective mutations in mouse neurofilament NF-M. J. Cell Biol. 111: 1987-2003 1990

Wong PC, Marszalek J, Crawford TO, Xu Z, Hsieh ST, Griffin JW, Cleveland DW. Increasing neurofilament subunit NF-M expression reduces axonal NF-H, inhibits radial growth, and results in neurofilamentous accumulation in motor neurons. J Cell Biol. 130(6):1413-22. 1995

Xie L, Helmerhorst E, Taddei K, Plewright B, Van Bronswijk W, and Martins R. Alzheimer’s beta-amyloid peptides compete for insulin binding to the insulin receptor. J Neurosci 22: RC221, 2002.

Xu, G.; Liu, Y.; Sayre, L. M. Independent synthesis, solution behavior, and studies on the mechanism of formation of a primary-amine-derived fluorophore representing crosslinking of proteins by (E)-4-hydroxy-2-nonenal (HNE). J. Org. Chem. 64:5732-5745; 1999.

Xu G, Liu Y, Kansal MM, Sayre LM. Rapid cross-linking of proteins by 4-ketoaldehydes and 4-hydroxy-2-alkenals does not arise from the lysine-derived monoalkylpyrroles. Chem Res Toxicol. 12(9):855-61.300. 1999

231 Xu Z., Cork L. C., Griffin J. W. and Cleveland D. W. Increased expression of neurofilament subunit NF-L produces morphological alterations that resemble the pathology of human motor neuron disease. Cell 73: 23-33 1993

Xu Z., Dong D. L. and Cleveland D.W. Neuronal intermediate filaments: new progress on an old subject. Curr. Opin. Neurobiol. 4: 655-6611994

Yamamoto A, Lucas JJ, and Hen R. Reversal of neuropathology and motor dysfunction in a conditional model of Huntington’s disease. Cell 101: 57–66, 2000.

Yamamoto H, Hasegawa M, Ono T, Tashima K, Ihara Y, and Miyamoto E. Dephosphorylation of fetal-tau and paired helical filaments-tau by protein phosphatases 1 and 2A and calcineurin. J Biochem 118: 1224–1231, 1995.

Yan SD, Chen X, Schmidt AM, Brett J, Godman G, Zou YS, Scott CW, Caputo C, Frappier T, Smith MA, Perry G, Yen SH, and Stem D. Glycated tau protein in Alzheimer disease: a mechanism for induction of oxidant stress. Proc Natl Acad Sci USA 91: 7787–7791, 1994.

Yankner BA. New clues to Alzheimer’s disease: unraveling the roles of amyloid and tau. Nature Med 2: 850–852, 1996.

Yin HS, Chou HC, and Chiu MM. Changes in the microtubule proteins in the developing and transected spinal cords of the bullfrog tadpole: induction of microtubule-associated protein 2c and enhanced levels of Tau and tubulin in regenerating central axons. Neuroscience 67: 763–775, 1995.

Yoo BC and Lubec G. p25 protein in neurodegeneration. Nature 411: 763–765, 2001.

Yoshida H and Goedert M. Molecular cloning and functional characterization of chicken brain tau: isoforms with up to five tandem repeats. Biochemistry 41: 15203–15211, 2002.

Yabe J. T., Pimenta A. and Shea T. B. Kinesin-mediated transport of neurofilament protein oligomers in growing axons. J. Cell Sci. 112: 3799-3814 1999

Yabe J. T., Chylinski T., Wang F. S., Pimenta A., Kattar S. D., Linsley M. D., et al. Neurofilaments consist of distinct populations that can be distinguished by C-terminal phosphorylation, bundling, and axonal transport rate in growing axonal neurites. J. Neurosci. 21: 2195-2205 2001

232 Yamasaki H., Itakura C., and Mizutani M. Hereditary hypotrophic axonopathy with neurofilament deficiency in a mutant strain of the Japanese quail. Acta Neuropathol. 82: 427-434 1991

Yin X., Crawford T. O., Griffin J. W., Tu P., Lee V. M., Li C., et al. Myelin-associated glycoprotein is a myelin signal that modulates the caliber of myelinated axons. J. Neurosci. 18: 1953-1962 1998

Yoshihara T., Yamamoto M., Hattori N., Misu K., Mori K., Koike H., et al. Identification of novel sequence variants in the neurofilament-light gene in a Japanese population: analysis of Charcot-Marie-Tooth disease patients and normal individuals. J. Peripher. Nerv. Syst. 7: 221-224 2002

Yu ZF, Bruce-Keller AJ, Goodman Y, Mattson, MP. Uric acid protects neurons against excitotoxic and metabolic insults in cell culture, and against focal ischaemic brain injury in vivo. J Neurosci Res 53 (5): 613-25. 1998

Zhao J. X., Ohnishi A., Itakura C., Mizutani M., Yamamoto T., Hayashi H., et al. Greater number of microtubules per axon of unmyelinated fibers of mutant quails deficient in neurofilaments: possible compensation for the absence of neurofilaments. Acta Neuropathol. (Berl) 87: 332-336 1994

Zheng Y. L., Li B. S., Veeranna and Pant H. C. Phosphorylation of the head domain of neurofilament protein (NF-M): a factor regulating topographic phosphorylation of NF-M tail domain KSP sites in neurons. J. Biol. Chem. 278: 24026-24032 2003

Zhou ZX, Kops O, Werner A, Lu JP, Shen M, Stoller G, Kullertz G, Stark M, Fischer G, and Lu PK. Pin1-dependent prolyl isomerization regulates dephosphorylation of Cdc25C and tau proteins. Mol Cell 6: 873–883, 2000.

Zs-Nagy, I. “A membrane hypothesis of aging” J Theor Biol 75: 189-95. 1978

Zhu Q., Lindenbaum M., Levavasseur F., Jacomy H. and Julien J. P. Disruption of the NF-H gene increases axonal microtubule content and velocity of neurofilament transport: relief of axonopathy resulting from the toxin beta, beta’- iminodipropionitrile. J. Cell Biol. 143: 183-193 1998

Zhu Q., Couillard-Despres S. and Julien J.P. Delayed maturation of regenerating myelinated axons in mice lacking neurofilaments. Exp. Neurol. 148: 299-316 1997

233 Zhu, X.; Rottkamp, C. A.; Boux, H.; Takeda, A.; Perry, G.; Smith, M. A. Activation of p38 kinase links tau phosphorylation, oxidative stress, and cell cycle-related events in Alzheimer disease. J. Neuropathol. Exp. Neurol. 59:880-888; 2000.

Zhu, X.; Castellani, R. J.; Takeda, A.; Nunomura, A.; Atwood, C. S.; Perry, G.; Smith, M. A. Differential activation of neuronal ERK, JNK/SAPK and p38 in Alzheimer disease: the “two hit” hypothesis. Mech. Ageing Dev. 123:39-46; 2001.

Zhu, X.; Lee, H. G.; Raina, A. K.; Perry, G.; Smith, M. A. The role of mitogen-activated protein kinase pathways in Alzheimer's disease. Neurosignals 11:270-281; 2002

234

APPENDIX

Publications, Abstracts, and Conferences

235 Publications:

Published:

1. Bishop GM, Robinson SR, Liu Q, Perry G, Atwood CS, Smith MA. Iron: a

pathological mediator of Alzheimer disease? Dev Neurosci. 2002;24(2-3):184-7.

2. Liu Q, Raina AK, Smith MA, Sayre LM, Perry G. Hydroxynonenal, toxic carbonyls,

and Alzheimer disease. Mol Aspects Med. 2003, 24(4-5):305-13.

3. Perry G, Taddeo MA, Petersen RB, Castellani RJ, Harris PL, Siedlak SL, Cash AD,

Liu Q, Nunomura A, Atwood CS, Smith MA. Adventiously-bound redox active iron

and copper are at the center of oxidative damage in Alzheimer disease. Biometals.

2003;16(1):77-81.

4. Taddeo MA, Smith MA, Liu Q, Atwood CS, Sayre LM, Perry G. Metal-catalyzed

redox activity in neurodegenerative disease. In: Metal Ions and neurodegenerative

Disorders, Zatta P, Ed, World Scientific Publishing Co., Singapore, 2003,pp1-14.

5. Liu Q, Xie F, Siedlak LS, Smith MA, Perry G. High Molecular Weight

Neurofilament Proteins in neurodegenerative diseases. Cell Mol Life Scie, in press

2004.

6. Liu Q, Siedlak SL, Harris PLR, Lee HG, Zhu X, Avila J, Takeda A, Smith MA,

Perry G. Tau modifiers as therapeutic targets for Alzheimer disease. Biochem

Biophys Acta, in press. 2004.

7. Moreira PI, Nunomura A, Honda K, Liu Q, Aliev G, Oliveira CR, Santos MS, Zhu

X, Smith MA, Perry G. Stress and homeostatic regulation of oxidative damage in

neurodegenerative disease. Proceedings of the XII Biennial Meeting of the Society

for Free Radical Research, in press. 2004

236 8. Liu Q, Smith MA, Avilá J, Bernardis JD, Kansal M, Takeda A, Zhu X, Nunomura

A, Honda K, Moreira1 PI, Oliveira CR, Santos MS, Shimohama S, Aliev G, de la

Torre J, Ghanbari HA, Siedlak SL, Harris PLR, Sayre LM, Perry G.

Alzheimer-specific epitopes of tau represent lipid peroxidation induced

conformations. Free Radical biology and medicine In press 2004

Comments

9. Liu Q, Honda K, Perry G, Smith M. on NEWS: Radical Development: Reactive

Oxygen Species Critical in Development, Is Redox Regulation of Development in

Caenorhabditis elegans a Critical... 2 Mar 2004. Alzheimer Research Forum, 2004,

10. Liu Q, Honda K, Moreira P, Perry G, Smith M, Zhu X. on PAPER: Palacino JJ. et

al., 2004, Parkin and Mitochondria: Are They Allies in the War Against

Parkinson’s... 30 Apr 2004. Alzheimer Research Forum, 2004.

11. Liu Q, Honda K, Moreira P, Perry G, Smith M. on PAPER: Zhang Q. et al., 2004,

Embalming Amyloid-β: The Role for Aldehyde Stress in Alzheimer's Disease...

2 May 2004. Alzheimer Research Forum, 2004,

12. Liu Q, Honda K, Moreira P, Perry G, Smith M, Zhu X. on PAPER: Coskun PE. et

al., 2004, Mitochondria and Alzheimer’s Disease: A Complex Interrelationship

Recently, Coskun... 27 Aug 2004. Alzheimer Research Forum, 2004,

Abstracts:

13. Perry G, Liu Q, Wataya T, Shimohama, S, Nunomura A, Siedlak SL, Sayre LM,

Smith MA. Neurofilament proteins are major targets of oxidative damage in the

nervous system. J Neuropathol Exp Neurol 61:456, 2002.

14. Liu Q, Smith MA, Siedlak SL, Harris PLR, Sayre LM, Perry G.

237 Phosphorylation-dependent regulation of neurofilament protein oxidation in

Alzheimer disease. 25th Graduate Student Symposium, p. 25, 2002.

15. Liu Q, Smith MA, Siedlak SL, Harris PLR, Sayre LM, Perry G.

Phosphorylation-dependent regulation of neurofilament protein oxidation. Neurobiol

Aging 23 (Suppl 1):S511-S512, 2002.

16. Liu Q, Smith MA, Sayre LM, Perry G. Neurofilament protein oxidation in

Alzheimer’s disease. 2003 Biomedical Graduate Student Symposium, P.46, 2003.

17. Liu Q, Smith MA, Siedlak SL, Harris PLR, Maccioni RB, Sayre LM, Perry G.

Phosphorylation and sequence dependency of neurofilament protein oxidation in

Alzheimer’s disease. Soc Neurosci Abstr, Program No.628.14, 2003.

18. Liu Q, Sayre LM, Siedlak SL, Harris PLR, Maccioni RB, Smith MA, Perry G.

Phosphorylation and sequence dependency of neurofilament protein oxidation.

Research ShowCASE 2004.

19. Liu Q, Smith MA, Siedlak SL, Harris PLR, Maccioni RB, Sayre LM, Perry G.

Phosphorylation modulated oxidative modification of neurofilament protein. J

Neuropathol Exp Neurol 63:523, 2004

20. Liu Q, Smith MA, Siedlak SL, Harris PLR, Maccioni RB, Sayre LM, Perry G.

Phosphorylation dependent control of oxidative modification in Alzheimer’s disease.

Soc Neurosci Abstr, 2004

Submitted:

21. Moreira PI, Honda K, Liu Q, Aliev G, Oliveira CR, Santos MS, Zhu X, Smith MA,

Perry G. Alzheimer’s disease and oxidative stress: the old problem remains unsolved.

Curr Med Chem-Central Nervous System Agents (Submitted)

238 In preparation:

22. Quan Liu, Mark A. Smith, Sandra L. Siedlak, Ricardo B. Maccioni†, Don W.

Cleveland, Lawrence M. Sayre, George Perry. Overexpression of neurofilament

heavy subunit protects neurons from oxidative stress. (Prepared to be submitted to

JBC).

23. Quan Liu, Mark A. Smith, Dandan Wang, De Lin, Sandra L. Siedlak, Peggy L.R.

Harris, Ricardo B. Maccioni†, Robert G. Salomon, Shu G. Chen, Lawrence M. Sayre,

George Perry. Phosphorylation and sequence dependency of neurofilament protein

oxidation in Alzheimer disease. (Prepared to be submitted to JBC or higher journals).

24. Quan Liu, Peggy L.R. Harris, Mark A. Smith, George Perry. High molecular weight

neurofilament subunits are the major pathophysiological targets of HNE modification

in brain. (Prepared to be submitted to journal of neurochemistry).

Conferences / Meetings Attended and Presented:

The 25th Graduate Student Symposium in CWRU, 2002

The 33th annual meeting of Society for Neuroscience in New Oreland 2003

The 26th Graduate Student Symposium in CWRU, 2003

The 2nd Research Showcase in CWRU 2004

The 80th American Association of Neuropathologists annual meeting in

Cleveland, Ohio 2004

The Society for Neuroscience's 34th Annual Meeting in San Diego, California

2004

239