CELLULAR AND MOLECULAR STUDIES OF

ALSIN/ALS2

Elizabeth Lilian Tudor BSc Hons

Thesis submitted in fulfilment of the degree of Doctor of Philosophy,

University of London (Kings College, Institute of Psychiatry)

September 2005

ABSTRACT

Mutations in alsin/ALS2 have been identified that cause several forms of motor neuron disease including a rare recessive form of juvenile ALS (ALS2), juvenile primary lateral sclerosis (jPLS) and infantile onset ascending hereditary spastic paraplegia (IAHSP).

Although the function of the ALS2 is unknown, ALS2 contains three functional domains that show sequence to guanine nucleotide exchange factors (GEFs) for the of . The first domain shows most similarity to a

GEF (RCC1-like domain), the second to Rho family GEFs (DH/PH domain) and the third to a GEF (VPS9 domain).

In these studies, a polyclonal antibody to ALS2 was generated and used to investigate

ALS2 localisation in the mammalian central nervous system. ALS2 was present in several neuronal populations in the brain and in motor neurons of the spinal cord. Since the DH/PH domain of ALS2 shows close homology to Rho family GEFs, in vivo assays were performed to investigate whether ALS2 acts as a GEF for the Rho family members

RhoA, Rac1 and Cdc42. ALS2 activated Rac1 but not RhoA or Cdc42. A major downstream effector of Rac1 is p21 activated kinase 1 (PAK1) and in vitro kinase assays revealed that ALS2 also activates PAK1. Thus, ALS2 stimulates Rac1-PAK1 signalling.

Rac1 is known to regulate neurite outgrowth during development. Immunofluorescence studies revealed that ALS2 is present in growth cones of rat embryonic neurons where it co-localised with Rac1. Furthermore, overexpression of ALS2 stimulated neurite outgrowth.

2 Finally, the phosphorylation state of ALS2 was investigated. ALS2 was found to be a phosphoprotein in vivo and five serine/threonine phosphorylation sites were identified.

However, mutation of the identified phosphorylation sites did not alter the ability of

ALS2 to activate Rac1 or PAK1, or to stimulate neurite outgrowth. The findings reported in this thesis provide an insight into ALS2 function, which may contribute to our understanding of the molecular mechanisms involved in motor neuron disease.

3 ACKNOWLEDGEMENTS

I would like to thank my supervisors Chris Miller and Chris Shaw for their constant advice and support, everyone in the Miller Group for help in the lab (especially Steve

Ackerley and Kwok-Fai Lau), Steve Banner for help with the immunohistochemical studies presented in Chapter 3, Anja Schmidt for constructs and advice towards the work presented in Chapter 4, Mike Perkinton for his encouragement, advice and help with the phosphorylation studies presented in Chapter 6 and Helen Byers for the mass spectrometry studies in Chapter 6. This research was funded by a Jim Tew Memorial

Prize Studentship from the Motor Neurone Disease Association UK.

4 TABLE OF CONTENTS

ABSTRACT...... 2

ACKNOWLEDGEMENTS ...... 4

TABLE OF CONTENTS ...... 5

LIST OF FIGURES AND TABLES...... 13

PUBLICATIONS ARISING FROM THIS WORK ...... 15

ABBREVIATIONS...... 16

CHAPTER 1: INTRODUCTION...... 21

1.1 Motor Neuron Disease ...... 22

1.1.1 Clinical features and pathology of MND...... 23

1.1.2 Genetics of MND...... 25

1.1.3 Animal models of MND ...... 29

1.1.3.1 Spontaneous mutants...... 29

1.1.3.2 Targeted mutants ...... 31

1.1.4 Mechanisms of neurodegeneration in MND...... 34

1.1.4.1 Toxicity of intracellular aggregates...... 35

1.1.4.2 Oxidative stress ...... 39

1.1.4.3 Defects in axonal transport...... 42

5 1.1.4.4 Glutamatergic excitotoxicity ...... 46

1.1.4.5 Neuroinflammation and autoimmunity ...... 51

1.1.4.6 ...... 52

1.1.4.7 Involvement of cell signalling pathways...... 54

1.2 Alsin/ALS2...... 56

1.2.1 Expression of ALS2...... 58

1.2.2 ALS2 exhibits GEF activity...... 60

1.2.3 ALS2 regulates endosomal morphology ...... 62

1.2.4 ALS2 binds to mutant SOD1 and displays neuroprotective activity...... 63

1.3 The Ras Superfamily of GTPases...... 64

1.3.1 Rho family GTPases ...... 67

1.3.2 Rab family GTPases ...... 72

1.4 Guanine nucleotide exchange factors (GEFs) ...... 74

1.4.1 DH/PH (Dbl-family) GEFs...... 74

1.4.2 Rab GEFs...... 77

CHAPTER 2: MATERIALS AND METHODS ...... 78

2.1 Materials ...... 79

2.1.1 Stock solutions...... 79

2.1.2 General molecular biology reagents ...... 81

2.1.2.1 Plasmids...... 81

2.1.2.2 Primers...... 81

2.1.2.3 Growth of E. coli for DNA purification: media ...... 82

6 2.1.2.4 Plasmid DNA preparation from E. coli ...... 82

2.1.2.5 Agarose gel electrophoresis of nucleic acids...... 83

2.1.2.6 Polymerase chain reaction (PCR) ...... 83

2.1.3 Site-directed mutagenesis ...... 84

2.1.4 Purification of GST fusion ...... 85

2.1.5 Protein analysis...... 85

2.1.5.1 Protein sample preparation...... 85

2.1.5.2 Sodium dodecyl sulphate-polyacrylamide gel electrophoresis (SDS-PAGE)

...... 87

2.1.5.3 Immunoblotting (Western blotting) solutions ...... 89

2.1.5.4 Protein staining...... 89

2.1.5.5 Treatment of protein samples with λ protein phosphatase...... 90

2.1.6 Antibodies...... 90

2.1.6.1 Primary antibodies...... 90

2.1.6.2 Secondary antibodies...... 90

2.1.7 Mammalian cell culture ...... 92

2.1.7.1 Chinese Hamster Ovary (CHO) cell culture...... 92

2.1.7.2 Primary neuronal culture ...... 92

2.1.8 In vitro protein kinase assays...... 93

2.1.9 Immunohistochemistry ...... 94

2.2 Methods...... 95

2.2.1 General molecular biology methods...... 95

2.2.1.1 Quantitation of nucleic acids...... 95

2.2.1.2 Restriction digestion of DNA ...... 95

2.2.1.3 Alkaline phosphatase treatment...... 96

7 2.2.1.4 Agarose gel electrophoresis of DNA...... 96

2.2.1.5 Recovery of DNA from agarose gels ...... 96

2.2.1.6 Ethanol precipitation of double-stranded DNA...... 97

2.2.1.7 Purification of nucleic acids ...... 97

2.2.1.8 Ligation of prepared vectors and DNA fragments ...... 97

2.2.1.9 Preparation of electrocompetent bacteria ...... 98

2.2.1.10 Electroporation of DH5α cells...... 99

2.2.1.11 Screening recombinant clones...... 99

2.2.1.12 Large scale preparation of plasmid DNA...... 100

2.2.1.13 DNA sequencing ...... 101

2.2.1.14 Polymerase chain reaction (PCR)...... 101

2.2.2 PCR-based site-directed mutagenesis...... 102

2.2.2.1 Digesting and polishing the PCR product ...... 104

2.2.2.2 Ligating the PCR product...... 104

2.2.2.3 Transformation into XL1-Blue supercompetent cells ...... 105

2.2.2.4 Screening for mutated plasmids ...... 105

2.2.3 Purification of GST fusion proteins...... 106

2.2.4 Protein analysis...... 108

2.2.4.1 Protein concentration determination...... 108

2.2.4.2 SDS-PAGE of protein samples ...... 109

2.2.4.3 Immunoblotting (Western blotting)...... 109

2.2.4.4 Treatment of samples with λ protein phosphatase ...... 110

2.2.5 Preparation of rabbit polyclonal antibodies...... 111

2.2.6 Mammalian cell culture and transfection...... 111

2.2.6.1 CHO cell culture...... 111

2.2.6.2 Primary embryonic rat cortical and hippocampal neuron culture ...... 112

8 2.2.6.3 Transient transfection ...... 113

2.2.7 GTPase activation assays...... 114

2.2.8 Immunoprecipitation and in vitro protein kinase assays ...... 115

2.2.8.1 Immunoprecipitation from cell lysates...... 115

2.2.8.2 Immunoprecipitation from rat brain homogenate...... 116

2.2.8.3 In vitro kinase assays...... 117

2.2.9 CHO cell fractionation...... 118

2.2.10 Immunohistochemistry ...... 118

2.2.11 Immunofluorescence...... 120

CHAPTER 3: PREPARATION OF AN ALS2 ANTIBODY

AND LOCALISATION OF ALS2 IN NEURONAL TISSUES

...... 121

3.1 Introduction...... 122

3.2 Methods...... 123

3.2.1 PCR and cloning into pGEX expression vector...... 123

3.2.2 Preparation of GST fusion protein...... 123

3.2.3 Production of a polyclonal antibody to ALS2 ...... 124

3.2.4 Affinity purification of ALS2 polyclonal antibody ...... 124

3.2.5 Cell culture and transfection...... 125

3.2.6 CHO cell fractionation...... 125

3.2.7 Antibodies...... 126

3.3 Results ...... 126

9 3.3.1 GST fusion protein preparation ...... 126

3.3.2 GST fusion protein expression and purification...... 127

3.3.3 Characterisation of ALS2 antibody ...... 128

3.3.4 Localisation of ALS2 in brain and spinal cord...... 129

3.3.5 Overexpressed ALS2 is localised in cytoplasmic and membranous fractions of

mammalian cells ...... 130

3.4 Discussion...... 130

CHAPTER 4: ALS2 ACTS AS GEF FOR RAC1 AND

ACTIVATES PAK1...... 140

4.1 Introduction...... 141

4.2 Methods...... 142

4.2.1 Plasmids...... 142

4.2.2 Cell culture and transfection...... 143

4.2.3 GTPase activation assays...... 143

4.2.4 In vitro kinase assays ...... 144

4.3 Results ...... 144

4.3.1 ALS2 acts as a GEF for Rac1 but not RhoA or Cdc42 ...... 144

4.3.2 ALS2 activates PAK1 and this is dependent on a functional DH domain ...... 145

4.4 Discussion...... 146

10 CHAPTER 5: ALS2 IS PRESENT IN NEURONAL

GROWTH CONES AND PROMOTES NEURITE

OUTGROWTH...... 153

5.1 Introduction...... 154

5.2 Materials and Methods...... 155

5.2.1 Antibodies and immunofluorescence microscopy...... 155

5.2.2 Preparation of mouse brain homogenates...... 155

5.2.3 Neurite outgrowth measurements ...... 156

5.2.4 Statistics...... 157

5.3 Results ...... 157

5.3.1 Developmental expression of ALS2...... 157

5.3.2 ALS2 is present in growth cones of hippocampal neurons, where it is co-

localised with Rac1, F-actin and α-...... 157

5.3.3 Overexpression of ALS2 (but not ALS2ΔDH) promotes neurite outgrowth in rat

embryonic cortical neurons via a Rac-dependent mechanism...... 158

5.3.4 Overexpression of ALS2 does not affect the number of neurites per cell, or the

extent of neurite branching, in cultured cortical neurons ...... 161

5.4 Discussion...... 161

CHAPTER 6: ALS2 IS A PHOSPHOPROTEIN...... 173

6.1 Introduction...... 174

11 6.2 Materials and Methods...... 175

6.2.1 Cell culture and transfection...... 175

6.2.2 Mass spectrometric sequencing of ALS2 ...... 175

6.2.3 Rac activation assay...... 176

6.2.4 PAK1 kinase assay ...... 177

6.2.5 Neurite outgrowth measurements ...... 177

6.3 Results ...... 177

6.3.1 ALS2 is a phosphoprotein ...... 177

6.3.2 Mutation of serine/threonine-proline sites to alanine does not affect the activity

of ALS2 ...... 178

6.4 Discussion...... 178

CHAPTER 7: DISCUSSION & FUTURE DIRECTIONS . 183

7.1 Summary of findings...... 184

7.2 The role of Alsin/ALS2 in motor neuron disease...... 185

7.3 Future directions ...... 187

REFERENCES...... 190

12 LIST OF FIGURES AND TABLES

Figures

Figure 1.1 Schematic of ALS2 (Long-form and Short-form) and predicted disease mutants...... 59

Figure 1.2 Schematic diagram of the small GTPase activation cycle...... 66

Figure 3.1 Schematic of ALS2 Long-form, ALS2 Short-form and GST-ALS2452-668..132

Figure 3.2 PCR amplification of sequences encoding GST-ALS2452-668...... 133

Figure 3.3 GST-ALS2452-668 and GST protein expression ...... 134

Figure 3.4 Purification of GST-ALS2452-668...... 135

Figure 3.5 Characterisation of ALS2 antibody ...... 136

Figure 3.6 Localisation of ALS2 in adult rat brain and spinal cord sections...... 137

Figure 3.7 Affinity-purified ALS2 antibody immunoprecipitates ALS2 from rat brain

...... 138

Figure 3.8 ALS2 is present in the cytosolic and membrane fractions of transfected CHO cells ...... 139

Figure 4.1 ALS2 stimulates Rac activity ...... 150

Figure 4.2 ALS2ΔDH does not activate Rac1 ...... 151

Figure 4.3 ALS2 activates PAK1...... 152

Figure 5.1 Developmental expression of ALS2 in mouse brain and cultured rat embryonic cortical neurons...... 167

Figure 5.2 Subcellular localisation of ALS2 in cultured embryonic neurons...... 168

Figure 5.3 Subcellular localisation of ALS2 in neuronal growth cones ...... 169

Figure 5.4 Transfected ALS2 and ALS2ΔDH display identical subcellular localisation to endogenous ALS2...... 170

Figure 5.5 ALS2 promotes neurite outgrowth in cultured rat cortical neurons ...... 171

13 Figure 5.6 Overexpression of ALS2 does not affect the number of neurites or the extent of neurite branching in cortical neurons...... 172

Figure 6.1 ALS2 is phosphorylated on serine/threonine-proline residues in embryonic rat cortical neurons...... 180

Figure 6.2 Phosphorylation of ALS2 ...... 181

Figure 6.3 Mutation of ALS2 phosphorylation sites does not affect ALS2 Rac GEF activity, PAK1 activation or neurite outgrowth ...... 182

Tables

Table 1.1 Classification of Motor Neuron Diseases ...... 22

Table 1.2 Loci identified in MND...... 26

Table 1.3 Loci identified in HSP...... 28

Table 1.4 Mammalian Ras GTPase Superfamily...... 65

Table 1.5 Selected effectors of the Rho GTPase family ...... 69

Table 2.1 Vectors ...... 81

Table 2.2 Mammalian expression plasmids...... 81

Table 2.3 Primary antibodies ...... 91

Table 2.4 PCR Cycling Parameters...... 102

Table 2.5 ExSite™ Mutagenesis Cycling Parameters ...... 104

Table 2.6 Transient transfection of CHO cells with Lipofectamine™ Reagent...... 113

14 PUBLICATIONS ARISING FROM THIS WORK

Peer Reviewed Research Paper

Tudor, E. L., Perkinton, M. S., Schmidt, A., Ackerley, S., Brownlees, J., Jacobsen, N. J.,

Byers, H. L., Ward, M., Hall, A., Leigh, P. N., Shaw, C. E., McLoughlin, D. M., and

C.C. Miller. 2005. ALS2/Alsin Regulates Rac-PAK Signalling and Neurite Outgrowth.

J. Biol. Chem. 280:34735-34740.

Published Abstracts

Elizabeth Tudor, Steven Ackerley, Nicholas O Jacobsen, Janet Brownlees, Steven

Banner, Anja Schmidt, Chris Shaw and Chris Miller (2003) “Alsin/ALS2 Is A Guanine

Nucleotide Exchange Factor That Is Present In Motor Neurons And Regulates Rho/Rac

Signalling” 33rd Annual Meeting of the Society for Neuroscience, New Orleans, USA.

Program No. 528.2.

Elizabeth Tudor, Steven Ackerley, Nicholas O Jacobsen, Janet Brownlees, Steven

Banner, Anja Schmidt, Chris Shaw and Chris Miller (2004) “Alsin/ALS2 Is A Guanine

Nucleotide Exchange Factor That Is Present In Motor Neurons And Regulates Rho/Rac

Signalling” 2nd European ALS Research Workshop (and European ALS Consortium

Young Investigators Meeting), Nice, France. Program No. P6.

15 ABBREVIATIONS

Amino acids: single letter code and abbreviations

Single Abbreviation Amino Acid letter code A Ala Alanine C Cys Cysteine D Asp Aspartic acid E Glu Glutamic acid F Phe Phenylalanine G Gly Glycine H His Histidine I Ile Isoleucine K Lys Lysine L Leu Leucine M Met Methionine N Asn Asparagine P Pro Proline Q Gln Glutamine R Arg Arginine S Ser Serine T Thr Threonine V Val Valine W Trp Tryptophan X Xxx Any amino acid Y Tyr Tyrosine

Other abbreviations

ABC avidin-biotinylated enzyme complex

ALS amyotrophic lateral sclerosis

16 Amp ampicillin

AMPA alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid

APS ammonium persulphate

ATP adenosine 5′-triphosphate

β-gal β-galactosidase bp nucleotide

BSA bovine serum albumin

Ca2+ calcium ion

CAT chloramphenicol acetyl

CCD charge coupled device

C. elegans

CHO Chinese hamster ovary cell line cDNA complementary DNA

CIAP calf intestinal alkaline phosphatase

CRIB Cdc42/Rac interactive binding

C-terminus carboxyl-terminus

Cu copper

DAB 3’,3’-diaminobenzidine tetrahydrochloride

Dbl diffuse B-cell lymphoma-associated protein ddH2O double distilled water

DH dbl homology

DH5α Douglas Hanahan bacterial strain 5α

DIV day(s) in vitro

DMSO dimethyl sulfoxide

DNA deoxyribonucleic acid dNTPs deoxyribonucleotide triphosphates

17 DTT dithiothreitol

E embryonic day

ECL enhanced chemiluminescence

E. coli Escherichia coli

EDTA ethylenediaminetetraacetic acid

EGTA ethyleneglycol-bis (β-aminoethylether) N,N,N′,N′-tetraacetic acid

ENU N-Ethyl-N-nitrosurea

FALS familial amyotrophic lateral sclerosis

FBS fetal bovine serum

Fgd1 faciogenital dysplasia 1-associated protein xg centrifugal force

GAP GTPase activating protein

GDI guanine nucleotide dissociation inhibitor

GDP guanosine 5’-diphosphate

GEF guanine nucleotide exchange factor

GFP green fluorescent protein

GST glutathione S-transferase

GTP guanosine 5’-triphosphate

GTPase guanosine 5’-

HEPES 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid

HRP horse radish peroxidase

HSP hereditary spastic paraplegia

IAHSP infantile-onset ascending hereditary spastic paraplegia

IF intermediate filament

IgG immunoglobulin gamma

IP immunoprecipitation

18 IPTG isothiopropyl-1-thio-β-D-galactopyranoside kDa kiloDalton

LB Luria Bertani

LMN lower motor neuron

MBP myelin basic protein

Mg2+ magnesium ion

MND motor neuron disease

MLB magnesium lysis buffer mRNA messenger RNA

Net1 neuroepithelial cell transforming 1

NF neurofilament

NF-H neurofilament heavy chain

NF-L neurofilament light chain

NF-M neurofilament medium chain

NGS normal goat serum

NMDA N-methyl-D-aspartate

NP-40 Nonidet P40

N-terminus amino-terminus

OD optical density

PAK p21 associated kinase

PBD p21 binding domain

PBS phosphate buffered saline

PCR polymerase chain reaction pH inverse log of hydrogen ion concentration

PH pleckstrin homology

PI(3)K phosphatidylinositol 3-kinase

19 PLS primary lateral sclerosis

PMSF phenylmethyl sulphonyl fluoride

RBD Rho binding domain

RCC1 regulator of condensation 1 protein

RNA ribonucleic acid

ROS reactive oxygen species

SALS sporadic amyotrophic lateral sclerosis

SDS-PAGE sodium dodecyl sulphate-polyacrylamide gel electrophoresis

SMA spinal muscular atrophy

SMN survival motor neuron

SOD1 Cu/Zn superoxide dismutase 1

TBS tris-buffered saline

Tris tris(hydroxymethyl)aminomethane

Tween-20 polyoxyethylene-sorbitan monolaurate

U units

UMN upper motor neuron

UV ultraviolet

VPS9 vacuolar protein sorting 9

X-gal 5-Bromo-4-chloro-3-indoyl-β-D-galactopyranoside

Zn zinc

20

CHAPTER 1: INTRODUCTION

21 1.1 Motor Neuron Disease

The term ‘Motor Neuron Disease’ (MND) describes a group of progressive neurodegenerative disorders involving selective loss of upper motor neurons (UMN) in the cortex and/or lower motor neurons (LMN) in the brainstem and spinal cord. Clinical features include paralysis and spasticity of the limbs and bulbar muscles (UMN involvement) and muscle weakness, wasting and fasciculation (LMN involvement)

(Devon et al., 2003; Talbot, 2002). Amyotrophic Lateral Sclerosis (ALS) is the most common MND, with a prevalence of 2-3 per 100,000 people (Cleveland and Rothstein,

2001). Following progressive degeneration of both UMNs and LMNs, death usually occurs within 1-5 years. By contrast, other forms of MND involve only the upper or lower motor neurons, resulting in disorders with varying severity (Table 1.1).

Table 1.1 Classification of Motor Neuron Diseases

Combined UMN and LMN involvement • Amyotrophic lateral sclerosis (ALS)

Proximal hereditary motor neuropathy Pure LMN involvement • • Hereditary bulbar palsy • Hexosaminidase deficiency • Multifocal motor neuropathies • Post polio syndrome • Post irradiation syndrome • Monomelic, focal or segmental spinal muscular atrophy (SMA)

Pure UMN involvement • Primary lateral sclerosis (PLS) • Hereditary spastic paraplegia (HSP) • Neurolathyrism • Konzo

Mixed Motor and Sensory involvement • ALS with fronto-temporal dementia (ALS- FTD) • Charcot Marie Tooth Disease (CMT) • Distal hereditary motor neuropathy

(Donaghy, 1999; Talbot, 2002)

22 1.1.1 Clinical features and pathology of MND

Amyotrophic Lateral Sclerosis (ALS)

The ‘Classical’ form of ALS is a sporadic disease of mid- to late- life, with a mean age at onset of 58 years (Ringel et al., 1993), although rare juvenile forms have also been described with age at onset varying from 3-23 years, and with a slower progression of about 10-15 years (Ben Hamida et al., 1990). The major pathological features of classical ALS are degeneration of the corticospinal tracts and extensive loss of anterior horn cells (Ghatak et al., 1986; Hughes, 1982; Leigh and Garofolo, 1995), degeneration and loss of Betz cells and other pyramidal cells in the primary motor cortex (Hammer et al., 1979; Maekawa et al., 2004; Udaka et al., 1986), and reactive gliosis in the motor cortex and spinal cord (Ekblom et al., 1994; Kawamata et al., 1992; Murayama et al.,

1991; Schiffer et al., 1996).

An established pathological hallmark of ALS is the presence of various inclusion bodies in degenerating neurons and surrounding reactive astrocytes (Barbeito et al., 2004).

Ubiquitinated inclusions are the most common and specific type of inclusion in ALS and are found in LMNs of the spinal cord and brainstem (Matsumoto et al., 1993) and in corticospinal UMNs (Sasaki and Maruyama, 1994). They are classified as ‘Lewy body- like inclusions’ (LBIs) and ‘Skein-like inclusions’ (SLIs) (He and Hays, 2004;

Kawashima et al., 2000). The exact composition of such inclusions is not known, although proteins identified so far include (in varying amounts) (Leigh et al.,

1991; Murayama et al., 1989), Cu/Zn superoxide dismutase 1 (SOD1) (Shibata et al.,

1996a; Shibata et al., 1994), peripherin (He and Hays, 2004) and Dorfin (a RING-finger type E3 ubiquitin ) (Niwa et al., 2002). Accumulations of intermediate filament proteins (mainly hyperphosphorylated neurofilament subunits and peripherin) are found in hyaline conglomerate inclusions (HCIs) and axonal ‘spheroids’ in spinal cord motor

23 neurons (Corbo and Hays, 1992; Munoz et al., 1988; Sobue et al., 1990) and pyramidal cells of the motor cortex (Troost et al., 1992) in ALS post-mortem tissue. Additionally,

Bunina bodies (BBs), which are Cystatin C-containing inclusions, are found in the cell bodies of motor neurons in ALS (Okamoto et al., 1993; Sasaki and Maruyama, 1994), although these are now thought to be less specific for ALS than the ubiquitinated and neurofilamentous inclusions, as they are similar to structures found in neurons of aged rats and humans (Kusaka, 1999). Other neuropathological features seen in ALS include fragmentation of the (Fujita et al., 2000; Fujita et al., 2002; Gonatas et al., 1998), mitochondrial vacuolisation (Okamoto et al., 1990) and ultrastructural abnormalities of synaptic terminals (Sasaki and Iwata, 1996).

Primary lateral sclerosis (PLS)

PLS is characterised by predominantly UMN degeneration resulting in spinobulbar spasticity. LMN and extra-motor system involvement has been reported in some cases, although if present this is usually mild or only seen at a late stage in the disease (Le

Forestier et al., 2001). The majority of PLS has an average age at onset of 50 years

(Pringle et al., 1992), and rarer juvenile forms have also been described (jPLS) (Gascon et al., 1995; Grunnet et al., 1989; Yang et al., 2001). In contrast to ALS, PLS has a very slow progression of about 15 years (Pringle et al., 1992) and cortical atrophy is more extensive (Kuipers-Upmeijer et al., 2001).

Hereditary spastic paraplegia (HSP)

Hereditary spastic paraplegia or hereditary spastic paraparesis (HSP) is a group of inherited neurodegenerative disorders of lower limb spastic paralysis, caused either by failure of development or progressive degeneration of the UMNs of the corticospinal tract. HSP is described as either ‘pure’ or ‘complex’, with autosomal dominant,

24 autosomal recessive and X-linked recessive patterns of inheritance (see section 1.1.2).

Complex HSP is associated with other clinical features such as distal amyotrophy, mental retardation, pigmentary retinopathy, sensory neuropathy or ataxia (Donaghy,

1999). Degenerating neurons of the corticospinal tract exhibit a dying back axonopathy in which synaptic terminal and axonal degeneration precedes cell body degeneration, and this typically occurs in cells with the longest axons first (Donaghy, 1999).

1.1.2 Genetics of MND

Familial ALS (FALS) subgroups

Whereas approximately 90% of ALS cases are sporadic (SALS) with no known genetic linkage, 10% are inherited and known as familial ALS (FALS) (Strong et al., 1991).

SALS and FALS share indistinguishable clinical and pathological features (Mulder et al., 1986) which suggests the possibility of common mechanisms underlying motor neuron degeneration in both sporadic and familial ALS. FALS inheritance is mainly autosomal dominant, but autosomal recessive and X-linked modes of inheritance have also been reported (Table 1.2). SALS and the majority of FALS is adult-onset, although three loci have been described that lead to juvenile-onset disease (ALS2, ALS4 and

ALS5). Approximately 20% of FALS cases are caused by point mutations in the Cu/Zn

SOD1 gene on chromosome 21 (Deng et al., 1993; Rosen et al., 1993). Over 100 SOD1 mutations have been described so far (for an updated list see www.alsod.org) resulting in highly variable phenotypes, even within families carrying the same mutation. Since the discovery of SOD1 mutations most studies on the pathogenesis of ALS have focused on mutant SOD1 mediated motor neuron death, utilising transgenic in vivo and in vitro models.

25 Table 1.2 Loci identified in MND

DISEASE ONSET INHERITANCE LOCUS GENE REFERENCE

Combined UMN and LMN involvement ALS1 Adult AD/AR 21q22.1 SOD1 (Rosen et al., 1993) ALS2/ (Hadano et al., 2001; Yang ALS2 Juvenile AR 2q33.2 Alsin et al., 2001) ALS3 Adult AD 18q21 - (Hand et al., 2002) (Blair et al., 2000; Chen et ALS4 Juvenile AD 9q34 SETX al., 2004) ALS5 Juvenile AR 15q15-22 - (Hentati et al., 1998) (Ruddy et al., 2003; Sapp ALS6 Adult AD 16q12 - et al., 2003; Abalkhail et al., 2003) ALS7 Adult AD 20ptel - (Sapp et al., 2003) ALS8 Adult AD 20q13.33 VAPB (Nishimura et al., 2004) ALS X Adult X-linked Xp11-q12 - (Siddique et al., 1998) Pure LMN involvement (Puls et al., 2003; Puls et SMA Adult AD 2p13 DCTN1 al., 2005) (Lefebvre et al., 1995; SMA Juvenile AR 5q13 SMN1 Rodrigues et al., 1995) SBMA/ Androgen (La Spada et al., 1991; Kennedy’s Adult X-linked Xq11-q12 Biancalana et al., 1992) Disease Pure UMN involvement jPLS Juvenile AR 2q33.2 ALS2 (Yang et al., 2001) HSP See Table 1.3 Mixed sensory and motor involvement ALS- FTD/ (Siddique and Hentati, Adult AD 17q21 MAPT Parkinsonism 1995) ALS- FTD/ Adult AD 17q - (Wilhelmsen et al., 2004) Parkinsonism ALS- FTD Adult AD 9q21-22 - (Hosler et al., 2000)

Abbreviations: ALS= amyotrophic lateral sclerosis; AD= autosomal dominant; AR= autosomal recessive; SOD1= Cu/Zn superoxide dismutase 1; SETX= Senataxin; VAPB= vesicle associated membrane protein (VAMP)/ synaptobrevin-associated membrane protein B; SMA= spinal muscular atrophy; SBMA= spinobulbar muscular atrophy; jPLS= juvenile primary lateral sclerosis; HSP= hereditary spastic paraplegia; FTD= frontotemporal dementia; MAPT= microtubule associated protein tau; DCTN1= dynactin 1; SMN1= survival motor neuron 1

26 Genetic risk factors for SALS

Several genetic risk factors, or susceptibility , have been implicated in sporadic

ALS. These include neurofilament heavy chain (NF-H) (Al-Chalabi et al., 1999;

Figlewicz et al., 1994; Skvortsova et al., 2004; Tomkins et al., 1998), vascular endothelial growth factor (VEGF) (Greenway et al., 2004), survival motor neuron 2

(SMN2) (Veldink et al., 2001), apolipoprotein E epsilon 4 (ApoEε4) (Drory et al., 2001;

Moulard et al., 1996), the gene encoding the p150 subunit of dynactin (DCTN1) (Munch et al., 2004), the mitochondrial gene cytochrome c oxidase (Comi et al., 1998), the gene encoding DNA repair enzyme abasic endonuclease (APEX) (Hayward et al., 1999) and aberrant mRNA processing of excitatory amino acid transporter 2 (EAAT2) (Aoki et al.,

1998; Lin et al., 1998). The absence of mutations in FALS suggests that the mutations found in SALS patients are susceptibility factors rather then a direct cause of the disease

(Aoki et al., 1998; Vechio et al., 1996).

Hereditary spastic paraplegia (HSP)

21 HSP genes have been mapped so far, 9 of which have been identified as follows: cell adhesion molecule L1 (L1-CAM), the gene encoding proteolipid protein (PLP1), atlastin, spastin, the neuronal specific protein gene KIF5A, the mitochondrial chaperone heat shock protein gene HSP60 (mitochondrial chaperonin), paraplegin, non imprinted in Prader-Willi/Angelman syndrome 1 (NIPA1), spartin, Berardinelli-Seip congenital lipodystrophy (BSCL2), maspardin and the ALS2/alsin gene ALS2 (Table

1.3).

27 Table 1.3 Loci identified in HSP

GENE LOCUS DISEASE INHERITANCE REFERENCE

(De Jonghe et al., 1996; SPG4/ Spastin 2p22 Pure HSP AD Hazan et al., 1999)

SPG13/ Hsp60 2q24-34 Pure HSP AD (Hansen et al., 2002)

Pure or (Reid et al., 1999a; Reid SPG10/ KIF5A 12q13 Complicated AD et al., 2002) HSP (Dalpozzo et al., 2003; SPG3A/ Atlastin 14q11- 21 Pure HSP AD Zhao et al., 2001b)

SPG6/ NIPA1 15q11.1 Pure HSP AD (Rainier et al., 2003)

Pure or (Casari et al., 1998; SPG7/ Paraplegin 16q24.3 Complicated AR Wilkinson et al., 2004) HSP Troyer SPG20/ Spartin 13q12.3 AR (Patel et al., 2002) syndrome Complicated SPG1/ L1CAM Xq28 X-linkedR (Jouet et al., 1994) HSP Complicated SPG2/ PLP Xq21 X-linkedR (Bonneau et al., 1993) HSP (Eymard-Pierre et al., ALS2/alsin 2q33.2 IAHSP AR 2002; Gros-Louis et al., 2003b) (Patel et al., 2001; Silver SPG17/ BSCL2 11q12- 14 AD Windpassinger et al., Syndrome 2004) (Reid et al., 1999b; SPG8 8q23- 24 Pure HSP AD Rocco et al., 2000) SPG21/ Mast 15q22.31 AR (Simpson et al., 2003) Maspardin Syndrome 10q23.3- Complicated SPG9 AD (Lo Nigro et al., 2000) 24.2 HSP (Ashley-Koch et al., SPG12 19q13 Pure HSP AD 2001; Reid et al., 2000) SPG19 9q33- 34 Pure HSP AD (Valente et al., 2002) Complicated SPG14 3q27- 28 AR (Vazza et al., 2000) HSP SPG5 8q12-13 Pure HSP AR (Muglia et al., 2004) (Martinez Murillo et al., SPG11 15q13-15 ARHSP AR 1999; Shibasaki et al., 2000) Complicated SPG15 14q22-24 AR (Hughes et al., 2001) HSP Pure or SPG16 Xq11.2 Complicated X- linkedR (Steinmuller et al., 1997) HSP 12p11.1– Complicated SPG26 AR (Wilkinson et al., 2005) 12q14 HSP

Abbreviations: SPG= Spastic Gait gene locus; AD= Autosomal dominant; AR= Autosomal recessive;

28 HSP60= heat shock protein 60 (mitochondrial Chaperonin); KIF5A= kinesin heavy chain; NIPA1= non imprinted in Prader-Willi/Angelman syndrome 1; L1CAM= L1 cell adhesion molecule; PLP= proteolipid protein; IAHSP= infantile-onset ascending HSP; BSCL2= Berardinelli-Seip congenital lipodystrophy; ARHSP= autosomal recessive HSP.

1.1.3 Animal models of MND

Animal models of neurodegeneration are an invaluable tool for studying the pathogenic mechanisms involved in vivo at all stages of the disease, even before symptom onset.

Murine models of MND in particular are used to study the cellular and molecular pathways involved, as such pathways are usually highly conserved between mouse and human (Hafezparast et al., 2003a). A selection of the most widely used mouse models are described below. Other animal models of MND include mutant SOD1 rats

(Howland et al., 2002) and survival motor neuron (SMN) knockdown

(McWhorter et al., 2003).

Transgenic mice have been created to investigate the function of proteins already known or suspected to be involved in MND, such as SOD1 and neurofilaments (NFs). This is known as a ‘genotype driven’ approach. In contrast, mice have been found

(‘natural/spontaneous mutants’) or created (by the use of irradiation or chemical mutagens) that exhibit a MND pathology, in a ‘phenotype driven’ approach. The causal gene has been identified in many of these mice and this has brought useful information regarding the proteins and pathways involved in motor neuron degeneration.

1.1.3.1 Spontaneous mutants

Wobbler mouse

The Wobbler mouse has an unsteady gait with progressive weakness, typically dying by

3 months of age (Duchen and Strich, 1968), and is often used as a model of MND.

29 However, recent studies have suggested that the disease in Wobbler mice may be a more generalised neurodegeneration than MND, as degeneration is seen in the thalamus, cerebellum and brainstem, and this precedes degeneration of the motor neurons and onset of gliosis (Rathke-Hartlieb et al., 1999). The gene mutated in this autosomal recessive disease model is not yet known but has been mapped to chromosome 11

(Kaupmann et al., 1992).

Progressive motor neuronopathy (pmn) mouse

The progressive motor neuronopathy (pmn) mutant mouse is a widely used model of

MND (most closely resembling SMA) that develops hind-limb paralysis and displays progressive degeneration of motor neurons until death occurs in the early postnatal period (6-8 weeks) (Schmalbruch et al., 1991). The autosomal recessive mutation causing this pathology was originally mapped to chromosome 13 (Brunialti et al., 1995) and has been subsequently identified as a point mutation in the tubulin-specific chaperone E (Tbce) gene (Bommel et al., 2002; Martin et al., 2002). Tbce encodes the protein E (CofE) which is a tubulin-specific chaperone and is essential in the correct assembly of microtubules (Bommel et al., 2002; Martin et al., 2002). This discovery is extremely interesting as defects of microtubule function and impairment of axonal transport have been extensively implicated in pathogenic mechanisms of MND, as discussed in section 1.1.4.3.

ENU-induced mutants

N-Ethyl-N-nitrosurea (ENU) is a chemical mutagen that produces point mutations in the genome, which enables identification of the causal gene by methods such as ‘positional candidate cloning’, and is currently being used to generate new mouse models of MND

(Brown and Balling, 2001; Chen et al., 2000; Hafezparast et al., 2003a). So far, this

30 approach has successfully created the mouse Legs at odd angles (Loa) which develops a late onset progressive motor neuron disease with neuropathological features similar to those seen in other MND mouse models, including mitochondrial swelling, Golgi fragmentation and cytoplasmic inclusions (Hafezparast et al., 2003a; Rogers et al.,

2001). The causal mutation in the Loa mouse is inherited in an autosomal dominant fashion, and has been found to occur in the cytoplasmic heavy chain gene

Dnchc1, which suggests involvement of the dynein-dynactin complex in MND.

Furthermore, a different mutation in dynein heavy chain has been identified in a second

ENU mutagenesis-generated mouse, Cramping1 (Cra1). The role of the dynein-dynactin complex in the pathogenesis of MND is discussed in section 1.1.4.3.

1.1.3.2 Targeted mutants

SOD1 transgenic mice

SOD1 is a ubiquitous enzyme and is highly expressed in motor neurons (Pardo et al.,

1995). The primary function of SOD1 within the cell is to catalyse the conversion

(dismutation) of superoxide radicals (by-products of normal cellular metabolism) to hydrogen peroxide, which is then eliminated by other free radical-scavenging enzymes

(glutathione peroxidase and catalase). It also has other activities in the cell including peroxidase activity (resulting in generation of hydroxyl radicals from hydrogen peroxide or superoxide, or production of nitronium species from peroxynitrite) and protection of the enzyme calcineurin from inactivation (Hodgson and Fridovich, 1975; Wang et al.,

1996). Transgenic mice carrying the human SOD1 mutants found in FALS (the most widely-used mutants are G37R, G85R, G93A, G93R) on a wild-type mouse SOD1 background develop progressive muscle weakness and atrophy and have a pathology that highly resembles the human disease, including loss of motor neurons and interneurons, reactive astrocytosis, and inclusion bodies immunoreactive for ubiquitin,

31 NFs and SOD1 (Bruijn et al., 1997b; Cha et al., 1998; Dal Canto and Gurney, 1995;

Gonatas et al., 1998; Gurney et al., 1994; Kong and Xu, 1998; Mourelatos et al., 1996;

Tu et al., 1996; Wong and Borchelt, 1995; Zhang et al., 1997). Mice with high copy numbers (high expresser mice) show early onset of disease, whereas mice with low copy numbers (low expresser mice) are affected by a late-onset disease (Dal Canto and

Gurney, 1997), and the type of mutation seems to affect the severity/rapidity of progression (Gurney, 1997; Shibata, 2001).

SOD1 mutant mouse models have therefore been widely used to study the molecular and cellular processes occurring both before disease onset and during disease progression. It is believed that such research may lead to identification of therapeutic targets and/or elucidation of the neurodegenerative mechanisms involved in all forms of

MND. The use of mutant SOD1 mice has enabled the identification of pre-symptomatic neuropathology such as fragmentation of the Golgi apparatus and mitochondrial vacuolation (Dal Canto and Gurney, 1995; Mourelatos et al., 1996; Wong and Borchelt,

1995), and biological processes disrupted before and during disease progression such as axonal transport (Borchelt et al., 1998; Warita et al., 1998; Williamson and Cleveland,

1999; Zhang et al., 1997).

The means by which SOD1 exerts its toxicity is unknown, although it is thought to occur by a ‘gain of toxic function’ rather than loss of the dismutase activity of the protein. This is supported by several findings from the study of mutant SOD1 mice, including: SOD1 wild-type knockout mice do not develop a disease pathology (although they display subtle motor defects and their motor neurons show an increased sensitivity to axonal injury) (Flood et al., 1999; Reaume et al., 1996; Shefner et al., 1999); the toxicity of mutant SOD1 is not accelerated or reduced by loss of wild-type SOD1 in

32 transgenic mutant SOD1 mice (Bruijn et al., 1998), and is either unaffected (Bruijn et al., 1998) or enhanced (Jaarsma et al., 2000) by increasing wild-type SOD1 activity; some mutants still cause disease despite retaining their dismutase activity (Borchelt et al., 1994; Bowling et al., 1993).

Neurofilament and peripherin mouse models

Neurofilaments are type IV intermediate filament (IF) proteins that are a major component of the neuronal cytoskeleton, and are composed of 3 subunits that range in size according to the size of their tail domain; neurofilament light chain (NF-L; 68 kDa), neurofilament medium chain (NF-M; 95 kDa), neurofilament heavy chain (NF-H;

110 kDa). They are responsible for the maintenance of neuronal calibre and are particularly abundant in large myelinated neurons, such as those that are preferentially affected in MND, and are found in filamentous inclusions in spinal cord motor neurons of MND post-mortem tissue. Peripherin, a type III neuronal IF protein, is also found in the majority of motor neuron neurofilament inclusions in MND, although its expression is usually (in non-disease cases) most abundant in autonomic nerves and peripheral sensory neurons, with low levels in spinal motor neurons (Escurat et al., 1990; Parysek and Goldman, 1988; Troy et al., 1990a; Troy et al., 1990b). The discovery of intermediate filament inclusions in MND has led to the proposal that they are involved in the pathogenesis of the disease, and so transgenic mouse models have been created to investigate this possibility. Transgenic mice overexpressing human NF-H (Cote et al.,

1993), mouse NF-L (Xu et al., 1993) or mouse peripherin (Beaulieu et al., 1999) all develop an MND-like pathology with muscle atrophy, reduced axonal calibre and IF inclusions. This has led to extensive research into the role of IF proteins in the disease mechanisms that cause MND, as discussed in sections 1.1.4.1 and 1.1.4.3.

33 Mutant VEGF mice

Mice carrying a targeted deletion of the hypoxia response element (HRE) within the vascular endothelial growth factor (VEGF) promoter have a phenotype similar to that of human ALS (Oosthuyse et al., 2001). This has led to the proposal that this gene may be involved in ALS, although linkage studies have as yet failed to support this hypothesis

(Gros-Louis et al., 2003a). The mechanism by which the mouse phenotype occurs is unknown, although it has been proposed that motor neuron death could occur due to reduced perfusion under low oxygen conditions (with motor neurons being particularly vulnerable due to their large size and high energy/oxygen requirements), and/or VEGF could have neuroprotective effects on motor neurons under normal circumstances

(Oosthuyse et al., 2001). Indeed, treatment with VEGF has been found to protect motor neurons against cell death following spinal cord ischemia in mice (Lambrechts et al.,

2003) and in a mutant SOD1 rat model (Storkebaum et al., 2005).

1.1.4 Mechanisms of neurodegeneration in MND

The pathogenic mechanisms underlying the neuronal degeneration and death in MND are as yet unknown. There are currently several hypotheses concerning the pathogenic processes involved in MND, none of which are mutually exclusive. The main hypotheses include protein misfolding and aggregation, oxidative stress, disruption of axonal transport, and glutamatergic excitotoxicity; there is also evidence that inflammation, autoimmunity and apoptotic cell death pathways are involved, all of which will be discussed.

Cell types involved in MND

Activated/reactive astrocytes and microglia are often found in ALS post-mortem tissue

(Ekblom et al., 1994; Murayama et al., 1991; Kawamata et al., 1992; Schiffer et al.,

34 1996) and in mutant SOD1 transgenic mice (Bruijn et al., 1997b; Cha et al., 1998; Tu et al., 1996), and although the clinical and molecular pathology of MND indicates an obvious involvement of motor neurons, it is increasingly being recognised that glial cells also play an important role in the pathogenesis of the disease. Important insights into the interdependence of different cell types in MND pathogenesis have come from several studies in mutant SOD1 transgenic mice. The discovery that cell-specific expression of mutant SOD1 in either neurons or astrocytes did not cause disease in mice

(Gong et al., 2000; Lino et al., 2002; Pramatarova et al., 2001) indicated that both neuronal and non-neuronal cells could play a role in disease pathogenesis. A chimeric

SOD1 mouse model was subsequently created, composed of a mixture of normal cells and mutant SOD1-expressing cells in order to investigate the relationship between different cell types (Clement et al., 2003). The presence of wild-type cells delayed disease onset and extended the lifespan of the chimeric mice compared with those overexpressing mutant SOD1 ubiquitously, and more specifically, degeneration and death of mutant SOD1-expressing motor neurons was reduced when the neurons were surrounded with a sufficient number of normal non-neuronal cells, whereas normal motor neurons surrounded by mutant SOD1-expressing non-neuronal cells showed signs of degeneration such as ubiquitinated inclusions. These results indicate that the expression of mutant SOD1 in both neuronal and non-neuronal cell populations, and the interactions between them, is of fundamental importance in MND.

1.1.4.1 Toxicity of intracellular aggregates

Mutant SOD1 aggregates in FALS

The mechanism(s) by which mutant SOD1 causes FALS are not known, although a popular hypothesis is that the mutant protein forms abnormal toxic aggregates within the cell, leading to neuronal degeneration and death. It has been suggested that the

35 toxicity of these aggregates may be effected by disruption of axonal transport (Borchelt et al., 1998; Williamson and Cleveland, 1999), sequestration of heat shock proteins and other chaperones (Bruening et al., 1999; Okado-Matsumoto and Fridovich, 2002;

Shinder et al., 2001), dysfunction of the (Hoffman et al., 1996; Urushitani et al., 2002) and/or damage to mitochondria (Jaarsma et al., 2001; Kong and Xu, 1998;

Takeuchi et al., 2002a).

SOD1 aggregates are found in motor neurons and surrounding astrocytes in FALS post- mortem tissue (Kato et al., 1997; Shibata et al., 1996b); they are also an early indicator of disease, occurring before the onset of symptoms, in mutant SOD1 transgenic mice

(Bruijn et al., 1997b; Johnston et al., 2000; Stieber et al., 2000). They are not, however, a characteristic feature of sporadic ALS (Shibata et al., 1996a). Further support for a role of abnormal SOD1 aggregation in FALS pathogenesis comes from the finding that delaying the formation of such abnormal aggregates, which are high molecular weight

‘insoluble protein complexes’ (IPCs), delays the disease onset in mutant SOD1 transgenic mice. This delay in aggregate formation was achieved by the creation of double transgenic mice carrying both human SOD1G93A and chat-GluR2 (in which the

GluR2 subunit of the alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid

(AMPA) receptor is overexpressed in a cholinergic neuron-specific manner, resulting in a large reduction in calcium-permeability of AMPA receptors, the significance of which will be discussed in section 1.1.4.4) (Tateno et al., 2004).

The way in which mutant SOD1 causes abnormal protein aggregation is unknown, although structural evidence suggests that it may be by means of destabilisation of normal dimers (Hough et al., 2004), and this has subsequently led to the proposal of dimer stabilisation as a potential therapeutic intervention (Ray and Lansbury, 2004).

36 Abnormal mutant SOD1 aggregates are not only present in the cytosol but have also been found in spinal cord mitochondria in both human cases and mouse models of mutant SOD1-mediated FALS (Liu et al., 2004; Pasinelli et al., 2004; Vijayvergiya et al., 2005). Together with the existence of extensive mitochondrial degeneration and vacuolation in mutant SOD1 mice at pre-clinical stages of the disease (Higgins et al.,

2002; Jaarsma et al., 2000; Wong et al., 1995), the localisation of such aggregates suggests an involvement of mitochondrial dysfunction in the pathogenesis of MND.

Intermediate filament aggregates in MND

Neurofilament subunits (NF-L, NF-M, NF-H) and peripherin are major components of the neuronal cytoskeleton and are frequently found in MND inclusion bodies (within neuronal cell bodies and axons) (Carpenter, 1968; Corbo and Hays, 1992; Hirano et al.,

1984; Migheli et al., 1993; Rouleau et al., 1996). Transgenic mice overexpressing IF proteins such as NF-L (Xu et al., 1993), human NF-H (hNF-H) (Cote et al., 1993), peripherin (Beaulieu et al., 1999) or alpha-internexin (Ching et al., 1999) all develop a motor neuron pathology with IF inclusions. In humans, mutations in NF-H have been reported in SALS patients (Al-Chalabi et al., 1999; Figlewicz et al., 1994; Tomkins et al., 1998), mutations in the NF-L gene have been found to cause the motor/sensory neuropathy Charcot Marie Tooth disease type 2 (CMT2) (De Jonghe et al., 2001;

Fabrizi et al., 2004; Georgiou et al., 2002; Jordanova et al., 2003; Mersiyanova et al.,

2000; Yoshihara et al., 2002; Zuchner et al., 2004) and two different mutations in the peripherin gene have recently been reported to cause MND (Gros-Louis et al., 2004;

Leung et al., 2004).

It has therefore been proposed that the abnormal IF accumulations found in ALS may contribute to the pathogenesis of the disease. It is unknown, however, how these

37 accumulations form and whether they play a detrimental or protective role in ALS. For example, crossing SOD1G37R mice with mice overexpressing hNF-H resulted in an extended lifespan compared with mice overexpressing SOD1G37R alone, and this correlated with the degree of perikaryal NF accumulation (Couillard-Després et al.,

1998). Similar results were found when the phenotype of peripherin overexpressing mice was rescued by overexpressing hNF-H (Beaulieu and Julien, 2003). The overexpression of hNF-H shifted the intracellular localisation of NF-H from the axonal to the perikaryal compartment, which suggests that the protective effect may be due to a reduction of axonal NF accumulation or an increase in perikaryal accumulation.

There is evidence that an increase in perikaryal accumulation is more important in protection from degeneration than a decrease in axonal accumulation, as decreasing the axonal NF content and axonal calibre in mutant SOD1 mice (by disruption in one allele of the NF-L gene), did not affect the disease severity or lifespan of the mice (Nguyen et al., 2000). It has been suggested that perikaryal NF inclusions may act as a

‘phosphorylation sink’ in which NFs are preferentially phosphorylated, thereby protecting other proteins that could be detrimental to the cell when hyperphosphorylated by kinases such as Cdk5 (Nguyen et al., 2001). In support of this hypothesis, it has been found that the perikaryal NF accumulations in the SOD1G37R/hNF-H mice are hyperphosphorylated and co-localise with Cdk5; and hyperphosphorylation of tau at

Cdk5 sites is reduced in these mice, suggesting that the NFs are preferentially phosphorylated by Cdk5 (Nguyen et al., 2001). In contrast, axonal accumulations of IF proteins are thought to have a detrimental effect, which may be caused by sequestration of essential proteins and organelles such as mitochondria, and/or by disruption of axonal transport, as discussed in Section 1.1.4.3.

38 Alternatively, sequestration of neuronal nitric oxide synthase (nNOS) in NF aggregates has been proposed to occur in neurons of mice overexpressing NF-L, leading to enhanced N-methyl-D-aspartate (NMDA)-mediated calcium influx that may cause neuronal cell death (Sanelli et al., 2004).

It is believed that NF subunit stoichiometry is a major factor in the formation of NF aggregates. This is supported by the finding that the motor neuron pathology found in hNF-H overexpressing mice was rescued by co-expressing human NF-L (hNF-L) at levels that restored the correct NF-L:NF-H stoichiometry. Additionally, reduced NF-L mRNA levels and selective alterations of NF expression have been observed in inclusion-bearing spinal cord motor neurons of ALS patients (Bergeron et al., 1994;

Menzies et al., 2002; Wong et al., 2000). Interestingly, in peripherin overexpressing mice, the disease was dramatically accelerated by a deficiency in NF-L (a phenomenon seen in ALS and in normal ageing) (Bergeron et al., 1994; Krekoski et al., 1996).

1.1.4.2 Oxidative stress

In ALS patients, biochemical changes indicative of oxidative damage, such as lipid peroxidation, free carbonyls, protein nitration and protein glycosylation, in motor neurons and spinal cord astrocytes suggests the involvement of oxidative stress in the disease (Abe et al., 1995; Beal et al., 1997; Bowling et al., 1993; Niebroj-Dobosz et al.,

2004; Shaw et al., 1995b; Shibata et al., 2001). For reviews see (Cookson and Shaw,

1999; Robberecht, 2000). Furthermore, the elevation of hydroxyl radicals in the spinal cord of pre-symptomatic SOD1 mutant mice has been reported in several studies

(Andrus et al., 1998; Bogdanov et al., 2000; Ferrante et al., 1997; Hall et al., 1998; Liu et al., 1998a), which suggests vulnerability of a particular subset of motor neurons at pre-clinical stages of the disease. The role of oxidative stress mechanisms leading to cell

39 death in MND has thus been investigated, particularly in relation to mutant SOD1- mediated FALS.

As mentioned earlier (Section 1.1.3), the toxicity of mutant SOD1 is thought to occur by means of a gain of (unknown) function rather than loss of its superoxide dismutase activity. This is suggested by the dominant inheritance pattern of FALS SOD1 mutations, the absence of any mutations causing severe truncations/loss of the protein and the finding that many of the mutants, for example D90A, retain their dismutase activity despite causing ALS. Although protein aggregation is a widely-favoured hypothesis for toxicity caused by mutant SOD1 in FALS, oxidative damage to the cell by mutant SOD1 has also been suggested as a pathogenic mechanism in FALS. It has also been postulated that the presence of such protein aggregates could be a consequence of reactive oxygen species (ROS) generation and oxidative modifications of the proteins (Gélinas et al., 2000; Rakhit et al., 2002; Valentine and Hart, 2003).

Studies using the spin trapping molecule 5’,5’-dimethylpyrroline-N-oxide (DMPO)

(which reduces oxidative stress propagation in NSC-34 motor neuron-like cells overexpressing SOD1G93A) (Liu et al., 2002) revealed that mutant SOD1 can become a source of free radicals by means of enhanced peroxidase activity, and can induce oxidative damage leading to lipid peroxidation, mitochondrial dysfunction and cell death (Liu et al., 2002; Wiedau-Pazos et al., 1996; Yim et al., 1996). This peroxidase activity has not, however, been observed in all studies (Singh et al., 1998). Alternatively it has been suggested that mutant SOD1 causes oxidative stress by production of nitronium ions, which may then nitrate tyrosine residues (Beckman et al., 1994), and this may be due to the mutants’ decreased affinity for zinc (Crow et al., 1997). This is supported by the presence of free 3-nitotyrosine in both human FALS cases and in

40 mutant SOD1 mice (Beal et al., 1997; Ferrante et al., 1997). These nitrotyrosines appear not to be protein bound, however, which calls into question the significance of this finding (Bruijn et al., 1997a), although it has recently been found that free nitrotyrosine can induce apoptosis (Peluffo et al., 2004). However, both of the above oxidative damage hypotheses (enhanced peroxidase and nitration) require the presence of an active-site copper bound to SOD1, but MND still occurred in a transgenic mouse expressing a mutant in which all four copper-coordinating histidines were mutated

(Wang et al., 2003). Furthermore, the disease phenotype seen in several mutant SOD1 mouse models was unaffected when the copper chaperone for SOD1 (CCS; a protein that is essential for copper loading of SOD1) was eliminated (Subramaniam et al.,

2002).

Although the origins of oxidative stress in MND are not clear, the localisation of mutant

SOD1 in the intermembrane space (IMS) (Liu et al., 2004) and matrix (Vijayvergiya et al., 2005) of mitochondria could cause mitochondrial dysfunction (membrane depolarisation, decreased activity of respiratory complexes and Cytochrome c release), and this may contribute to oxidative stress pathways leading to neuronal death in FALS

(Beretta et al., 2003; Higgins et al., 2002; Jaarsma et al., 2001; Sturtz et al., 2001).

Mitochondrial dysfunction is also implicated in Hereditary spastic paraplegia, as two of the known mutations that cause the disease (SPG7/paraplegin and SPG13/HSP60) are found in genes encoding mitochondrial chaperone proteins, which are involved in protein folding in the IMS. Indeed, several patients with the SPG7/paraplegin mutation show signs of impaired mitochondrial oxidative phosphorylation, which is shown by the presence of abnormal mitochondria (ragged red fibers) and Cytochrome c oxidase deficient-fibers in muscle biopsies (Casari et al., 1998). Furthermore, mitochondrial abnormalities such as hypertrophy, concentric cristae, herniations and giant

41 mitochondria, found in the synaptic terminals of spinal cord motor neurons, is the earliest pathological feature seen in Paraplegin knockout mice (Gelbard, 2004).

Together these studies suggest an involvement of mitochondria in the pathogenesis of the MND.

1.1.4.3 Defects in axonal transport

The presence of IF accumulations in ALS has led to the proposal that NF transport is somehow perturbed, and subsequent studies in mouse models of MND have indicated that defects in slow axonal transport is one of the earliest pathological events in the disease (Williamson and Cleveland, 1999). In particular, axonal transport is abnormally reduced in mice with the SOD1 G93A, G37R and G85R mutations, and this precedes the onset of neuropathology (Borchelt et al., 1998; Warita et al., 1999; Williamson and

Cleveland, 1999; Zhang et al., 1997). Additionally, axonal transport defects have been observed in the Wobbler mouse (Mitsumoto et al., 1990; Mitsumoto and Gambetti,

1986), pmn mutant mouse (Sagot et al., 1998), hNF-H overexpressing mice (Collard et al., 1995) and the Paraplegin knockout mouse (Ferreirinha et al., 2004).

It has been hypothesised that NF accumulations in transport-deficient axons may impede further transport along microtubules, leading to cell death by ‘axonal strangulation’ (Williamson and Cleveland, 1999). Axonal transport defects have also been proposed to contribute to the dying back axonopathy observed in HSP due to impairment of efficient protein delivery to synaptic terminals. Furthermore, a lack of trophic support has been implicated in many forms of MND and this may also be caused by faulty transport mechanisms. Motor neurons are one of the largest cells in the nervous system (cell body diameter of 50-60 µm) with axonal processes of up to 1 metre in length and possess a high NF content due to their need for a robust cytoskeletal

42 network. Therefore, defects in axonal transport and abnormal NF accumulation could explain the selectivity of motor neuron death in MND. It is not known how the slowing of axonal transport in MND occurs, although several mechanisms have been proposed.

Role of hyperphosphorylation of NF subunits

Hyperphosphorylation of NF subunits has been proposed to play a role in deregulation of NF transport in MND. The mutations in NF-H that have been identified as a risk factor for SALS affect phosphorylation of side-arm domain Lys-Ser-Pro (KSP) repeat motifs (Al-Chalabi et al., 1999; Tomkins et al., 1998) and such KSP repeat phosphorylation is believed to be important in regulating NF transport. KSP repeats in

NF-M and NF-H are phosphorylated by multiple kinases such as glycogen synthase kinase-3α and -3β (Gsk3α/β), p42/p44 mitogen-activated protein kinases (p42/p44

MAPKs, also known as extracellular signal-related kinase (ERK2/ERK1), stress activated protein kinases (SAPKs) and the cyclin-dependent kinase-5/p35 complex

(Cdk5/p35) (Brownlees et al., 2000; Giasson and Mushynski, 1998; Guan et al., 1991;

Guidato et al., 1996; Sun et al., 1996; Veeranna et al., 1998). Indeed, studies in SOD1 mutant mice have shown that Cdk5 hyperphosphorylates NFs, leading to NF inclusions

(Nguyen et al., 2001) and SAPKs phosphorylate NF subunits in response to glutamate treatment, resulting in a slowing of NF transport, in cultured cortical neurons (Ackerley et al., 2000). However, recent studies using liquid chromatography tandem mass spectrometry (LC/MS/MS) have found that the phosphorylation of NF-H from sporadic

ALS spinal cord neurons does not differ from control samples (Strong et al., 2001), although it has been proposed that the levels of NF phosphorylation in the cell body compared with the axon may differ and this could affect the assembly and transport of

NFs (Miller et al., 2002; Strong et al., 2001).

43 Role of motor proteins

There is evidence to suggest that defects in the motor proteins which mediate axonal transport may be involved in the slowing of axonal transport in MND. Studies of green fluorescent protein (GFP)-tagged NF-M transport in cultured neurons have shown that transport is dependent on microtubules and the anterograde motor protein kinesin

(Koehnle and Brown, 1999; Yabe et al., 1999). Kinesin heavy chain mutations have been found to affect axonal transport in (Hurd and Saxton, 1996) and more recently, mice lacking the neuronal-specific kinesin heavy chain KIF5A showed a defect in NF transport, with NF subunits accumulating in neuronal cell bodies (Xia et al., 2003). Furthermore, the transport of tubulin is impaired and axonal levels of kinesin are reduced in SOD1G85R mice months before disease onset (Warita et al., 1999; Zhang et al., 1997), levels of cDNA encoding the kinesin-like protein KIF3B are reduced in the motor neuronal cell line ‘NSC34’ stably-transfected with mutant SOD1 (Kirby et al.,

2002) and mutations in KIF1Bβ have been found to cause Charcot Marie Tooth disease type 2A (CMT2A) (Zhao et al., 2001a). Kinesin dysfunction has also been implicated in

HSP (see below; Axonal transport defects in Hereditary spastic paraplegia (HSP)).

Additionally, components of the retrograde motor protein complex (dynein/dynactin) have been found to associate with NFs and catalyse their transport in vitro (Shah et al.,

2000) and mutations in dynactin (p150 subunit) have been found to cause an autosomal dominant form of lower motor neuron disease in humans (Puls et al., 2003). As dynein knockout mutations are lethal in both mouse and Drosophila models, a transgenic mouse model with a targeted disruption of the dynein/dynactin complex was recently engineered to investigate dynein involvement in MND. This was achieved by overexpression of the dynamitin subunit of dynactin (which disassembles dynactin) in postnatal motor neurons, resulting in the development of late-onset progressive motor

44 neuron degeneration and muscle atrophy, thus confirming the critical role of axonal transport in the pathogenesis of MND (LaMonte et al., 2002). Furthermore, mouse models have been utilised to demonstrate that motor neurons are uniquely sensitive to disruption of dynein function and retrograde transport. Mice heterozygous for either of two ENU-generated mutations, Loa and Cra1, demonstrate a motor neuron degenerative phenotype similar to the dynamitin transgenic mouse, and positional cloning has revealed that both mutations are in the cytoplasmic dynein heavy chain 1 gene

(Dnchc1). Neither mutation seems to affect the localisation of dynein, or its expression levels, but its function is subtly inhibited, affecting motor neurons alone. Retrograde transport of a fluorescently-labelled fragment of Tetanus toxin was found to be significantly reduced in these mice, whereas other functions of dynein such as nuclear motility during cell division, and formation and positioning of the Golgi apparatus, were normal (Hafezparast et al., 2003b; He et al., 2005; Kieran et al., 2005).

Axonal transport defects in hereditary spastic paraplegia (HSP)

One autosomal recessive form of HSP is caused by loss-of-function mutations in the

SPG7 gene, which encodes the protein Paraplegin (a member of the ‘ associated with a variety of cellular activities’ (AAA) family). Paraplegin knockout mice display an MND-like pathology and their neurons have large axonal swellings containing organelles and NFs, with impaired retrograde axonal transport (Ferreirinha et al., 2004). The mice also have abnormal mitochondria in spinal cord neurons, and this mitochondrial phenotype correlates with disease onset and neuronal degeneration, which has led to the proposal that mitochondrial dysfunction may underlie defective axonal transport. However, impairment of axonal transport in these mice is only seen after the onset of disease, which indicates that it may not be the primary cause of neurodegeneration (Ferreirinha et al., 2004). Mutations in the neuronal-specific kinesin

45 gene KIF5A have recently been found to cause a form of HSP (SPG10) in humans (Reid et al., 2002), which further implicates defective axonal transport in MND. SPG4/Spastin

(an AAA protein, but in a different subgroup to paraplegin) has been found to bind to microtubules, and has been proposed to act as a microtubule-severing enzyme (Charvin et al., 2003; Errico et al., 2002) which suggests that this form of HSP may involve cytoskeletal disassembly/disruption of transport. Interestingly, Spastin abnormally co- localises with kinesin when overexpressed in HEK cells and neurons (McDermott et al.,

2003) and recent studies using RNA interference (RNAi) in have revealed a role for Spastin in microtubule assembly in synaptic terminals (Trotta et al., 2004). A further microtubule-interacting protein, the novel GTPase Atlastin, is also mutated in juvenile-onset HSP (SPG3A) (Dalpozzo et al., 2003; Zhao et al., 2001b), although there is currently no evidence to suggest that this interaction has any effect on axonal transport. Atlastin shares homology with the family of large GTPases, which are involved in molecular trafficking events such as synaptic vesicle recycling and mitochondrial dispersion (Jones et al., 1998; Nicoziani et al., 2000; Smirnova et al.,

1998).

1.1.4.4 Glutamatergic excitotoxicity

A role for glutamatergic excitotoxicity (a mechanism whereby prolonged or excessive exposure to extracellular glutamate leads to cell death) in MND was first implicated by observations of elevated glutamate levels in the cerebrospinal fluid (CSF) of SALS patients (Plaitakis and Caroscio, 1987; Rothstein et al., 1990; Shaw et al., 1995a). The level of neuronal excitation by glutamate is regulated by a number of mechanisms. Two of relevance to MND are NMDA receptor and AMPA/kainate receptor inactivation and

Na+/K+-coupled glutamate reuptake by astrocytic transporter proteins (excitatory amino acid transporters; EAAT1-5, also known as glutamate transporters; GLT1-5). There is

46 extensive evidence to suggest an involvement of both mechansims in the pathogenesis of ALS, which is interesting as it highlights the importance of both motor neurons and astrocytes in the disease, as discussed in section 1.1.4. The precise molecular mechanisms that lead to excitotoxicity-mediated cell death are not known, although several pathways have been identified that contribute, including disruption of intracellular calcium homeostasis and production of free radicals (For a recent review see (Heath and Shaw, 2002)). Additionally, the motor neurons that are preferentially affected in MND are likely to be particularly sensitive to excitotoxic insults for several reasons, including a high expression of AMPA receptors lacking the GluR2 subunit

(this makes them highly calcium-permeable) and a low level of calcium-binding proteins such as calbindin D-28k and parvalbumin, compared with the neuronal populations that are frequently spared in MND such as the oculomotor, trochlear, abducens nerve and Onuf’s nucleus motor neurons (Alexianu et al., 1994; Elliott and

Snider, 1995; Ince et al., 1993; Reiner et al., 1995). Indeed, overexpression of parvalbumin in a transgenic mouse model of ALS delayed the disease onset (Beers et al., 2001). Together, these findings support a role for excitotoxicity and disrupted calcium homeostasis in MND.

Role of glutamate transporters

Astrocytes are the principle regulators of extracellular glutamate levels (Rothstein et al.,

1996), and deficiency in glutamate uptake by astrocytes appears to play a crucial role in the pathogenesis of ALS. The exact cause of glutamate transport deficiency in ALS is not known, although it has been postulated that the reduction in transport function is caused by a loss of the astrocytic glutamate transporter EAAT2 (also known as GLT1)

(Rothstein et al., 1992). In 60-70% of SALS cases, and in SOD1 mutant mouse models, a large reduction in EAAT2 expression (which is not purely due to cell death) is seen at

47 the end-stage of disease (Bristol and Rothstein, 1996; Bruijn et al., 1997b; Rothstein et al., 1995). EAAT2 reduction has also been observed at early stages of disease (before the onset of hind-limb paralysis) in a mutant SOD1G93A rat model and this loss is specific to areas of the spinal cord that contain motor neuron cell bodies (Howland et al., 2002). Furthermore, EAAT2 knock-down, obtained by administration of antisense oligonucleotides in vitro and in vivo, resulted in progressive hind-limb paralysis and motor neuron degeneration in rats (Rothstein et al., 1996). It has been suggested that abnormal mRNA splicing of EAAT2 could underlie reduced expression and function of the protein in ALS (Lin et al., 1998), although several other groups have failed to replicate these findings (Aoki et al., 1998; Flowers et al., 2001). However, no motor neuron loss is seen in EAAT2 knockout mice (although they develop hippocampal pathology and seizures, and most die as juveniles) (Tanaka et al., 1997). This suggests that an overall loss of EAAT2 function does not lead to MND pathology. Additionally, no reduction in EAAT2 expression has been observed in the brain or spinal cord of mice overexpressing SOD1G93A (Deitch et al., 2002). Several studies suggest that the reduction in EAAT2 function may be caused by biochemical changes in transporters already present rather than differences in transporter levels. For example, studies in

Xenopus oocytes have shown that the oxidative activity of mutant SOD1 leads to a reduction in EAAT2 function, and this effect can be blocked by the antioxidant

Mn(III)TBAP (Trotti et al., 1999). This suggests that EAAT2 may be susceptible to oxidative damage, resulting in decreased glutamate uptake at the synaptic cleft, and excitotoxic damage to motor neurons. Additionally, a mutation in the EAAT2 gene which results in impaired glutamate clearance capacity has also been reported in a single SALS case (Trotti et al., 2001).

48 Role of glutamate receptors

AMPA receptors are composed of different combinations of four subunits GluR1-4 (or

GluRA-D); the GluR2 (GluRB) subunit is of particular importance in determining the calcium permeability of the assembled receptor. An mRNA-editing defect of GluR2 in spinal cord motor neurons from 5 individual SALS cases has recently been reported, which confers increased calcium-permeability on AMPA receptors, leading to cell death

(Kawahara et al., 2004). Therefore, the presence/function of the GluR2 subunit and the calcium-permeability of AMPA receptors are believed to play a significant role in excitotoxic pathways in MND. The effect of a reduction in the calcium-permeability of

AMPA receptor in motor neurons has recently been investigated in a mutant SOD1 mouse model. This was achieved by crossing SOD1G93A overexpressing mice with a mouse line overexpressing GluR2 specifically in cholinergic neurons (resulting in a large reduction in the calcium-permeability of motor neuronal AMPA receptors). These mice displayed a delay of disease onset with a correlating delay in the formation of abnormal intracellular SOD1 aggregates, as compared with mice overexpressing

SOD1G93A alone (Tateno et al., 2004), suggesting that the calcium-permeability of

AMPA receptors specifically in motor neurons affects the formation of SOD1 aggregates, leading to MND, in this model. Additionally, the amount of carbonylated proteins (a marker of oxidative stress) in the spinal cord was also reduced/delayed in these mice, and it has been proposed therefore that the protein aggregates seen in the disease may be a consequence of ROS production (Tateno et al., 2004). Complimentary to these findings was a recent study in which a mouse line was created that overexpressed a functionally-modified GluR2 subunit (GluRB(N)), in which the subunit conferred calcium-permeability on its assembled AMPA receptor but the conductivity of the receptor was unaffected. These mice displayed a phenotype of progressive MND that closely resembles human SALS, and crossing with SOD1G93A transgenic mice

49 resulted in an acceleration of disease progression and decrease in survival, which further supports a role for an increase in calcium permeability caused by defective GluR2 mRNA editing in MND (Kuner et al., 2005).

Involvement of oxidative stress in excitotoxicity

Glutamate transporters are susceptible to oxidative damage, and oxidative modifications of transporter peptides have been reported both in ALS cases and in SOD1 mutant mice

(Liu et al., 2002; Pedersen et al., 1998), indicating that the reduction of transporter function in ALS may be caused by oxidative mechanisms. The presence of signs of mitochondrial degeneration in both human ALS cases and in SOD1 mutant mouse models (in which this precedes the onset of motor defects) provides a link between oxidative stress and excitotoxic mechanisms. It has been proposed that lack of mitochondrial function, coupled with the high energy demands of motor neurons, may result in a lowering of the neuron’s membrane potential, resulting in opening of glutamate receptors and influx of calcium, and less glutamate would then be needed to have an excitotoxic effect on the cell. However, a lack of involvement of oxidative stress in downregulation of EAAT2 has been reported in astrocytes (Tortarolo et al.,

2004).

Riluzole and evidence against the excitotoxicity hypothesis

In support of the glutamatergic excitotoxicity hypothesis, Riluzole, a drug that inhibits glutamatergic transmission, has been shown to delay disease progression in several forms of ALS (Bensimon et al., 1994; Lacomblez et al., 1996). However, the failure of several other antiglutamatergic agents, for example Gabapentin, in human ALS trials

(Miller et al., 1996; Miller et al., 2001) suggests that Riluzole may have a novel mode of action. Additionally, a recent study has demonstrated a lack of glutamatergic

50 involvement in ALS, using an in vivo rat spinal cord microdialysis model, and has postulated that large and long-lasting increases of glutamate in the spinal cord do not produce motor neuron hyperexcitation or degeneration (Corona and Tapia, 2004).

1.1.4.5 Neuroinflammation and autoimmunity

Neuroinflammation has been implicated in ALS due to several findings, including upregulation of the pro-inflammatory enzyme cyclooxygenase-2 (COX-2) mRNA and protein expression, and increased COX-2 activity (measured by increased prostaglandin

E2 levels) in SALS spinal cord (Maihofner et al., 2003; Yasojima et al., 2001) and in mutant SOD1 transgenic mice (Almer et al., 2001). Furthermore, a selective COX-2 inhibitor, Celecoxib, protects spinal cord motor neurons from mutant SOD1-mediated cell death by prolonging survival and protecting against microglial activation, astrogliosis and spinal cord neuron degeneration (Drachman et al., 2002; Pompl et al.,

2003).

Inflammatory responses also recruit immune mechanisms, the main effectors of which

(astrocytes and microglia) are activated in ALS (Hall et al., 1998; Kamo et al., 1987;

Kawamata et al., 1992; Schiffer et al., 1996). Indeed, microglial activation occurs before disease onset in mutant SOD1 transgenic mice (Alexianu et al., 2001) and the anti- inflammatory compound minocycline extends survival in mouse models of ALS (Kriz et al., 2002; Van Den Bosch et al., 2002; Zhu et al., 2002), which may occur by inhibiting microglial activation (Tikka et al., 2001; Tikka et al., 2002). Additionally, infiltration of macrophages, mast cells and T cell lymphocytes have been found in

SALS spinal cord (Graves et al., 2004; Hayashi et al., 2001; Kawamata et al., 1992;

Lampson et al., 1990), which suggests an involvement of both innate and acquired immune responses in MND. Autoantibodies against various components in the CNS

51 have been reported in ALS, although it is still unclear whether this represents an involvement of autoimmunity in the pathogenesis of MND, or if such antibodies have a beneficial effect in the disease. Examples of antibodies found in SALS sera include neuronal, NF subunit and calcium channel antibodies (Couratier et al., 1998; Kimura et al., 1994; Niebroj-Dobosz et al., 1999; Smith et al., 1992). Monoclonal immunoglobulin

G (IgG) is also detectable in SALS spinal cord, motor cortex and sera (Duarte et al.,

1991; Engelhardt and Appel, 1990). However, anti-inflammatory and immunosuppressive therapies have had little beneficial effect in ALS to date (Drachman et al., 1994).

1.1.4.6 Apoptosis

There is evidence to suggest that neuronal death in SOD1-mediated FALS occurs via apoptotic signalling mechanisms, which is implicated by the presence of DNA fragmentation, decreased expression of the anti-apoptotic protein Bcl-2, and increased expression of the pro-apoptotic protein Bax, in the spinal cords of ALS patients and of transgenic mutant SOD1 mice (Fujita et al., 2002; Martin, 1999; Vukosavic et al., 1999;

Yoshiyama et al., 1994). Additionally, expression of Bcl-2 in SOD1G93A mutant mice delays disease onset and increases survival (Kostic et al., 1997) and mutant SOD1 binds to Bcl-2 in mitochondria, which suggests that mutant SOD1 may sequester Bcl-2 into abnormal aggregates and thereby prevent its ability to participate in anti-apoptotic pathways (Pasinelli et al., 2004). Caspase inhibitors have also been effective in delaying onset and prolonging life (Li et al., 2000). Indeed, sequential activation of caspase-1 and caspase-3 (which belong to the family of apoptotic effector cysteine proteases) (Villa et al., 1997) in mutant SOD1 mouse spinal cord motor neurons and astrocytes coincides with onset of motor neuron loss (Li et al., 2000; Pasinelli et al., 2000; Vukosavic et al.,

1999).

52 Three major apoptotic pathways have been the most widely-studied; the mitochondrial

(intrinsic) pathway, the death receptor (extrinsic) pathway and the endoplasmic reticulum (ER) pathway. It has been shown that Cytochrome c translocation from mitochondria to the cytosol, and Bax translocation from the cytosol to mitochondria, which are crucial steps in the mitochondrial-dependent apoptotic pathway (Kroemer and

Reed, 2000), occurs in the spinal cord of SOD1G93A transgenic mice in parallel with neurodegeneration (Guegan et al., 2001). Additionally in these mice, activation of caspase-9 followed by caspase-7, and cleavage of the -linked inhibitor of apoptosis protein (XIAP) was reported, further supporting a role of the mitochondrial apoptotic pathway. Involvement of the death receptor pathway has also been identified in MND. Nitric oxide signalling via the death receptor Fas leads to activation of caspase-dependent specific death of embryonic spinal cord motor neurons in culture

(Raoul et al., 2002) and another death receptor p75NTR has also been implicated in cell death in mutant SOD1 mice (Kust et al., 2003; Turner et al., 2003).

Although it remains unknown how apoptosis may be triggered in MND, several hypotheses exist. Reactive astrocytes have been shown to promote apoptosis of motor neurons in culture via a nitric oxide and peroxynitrite-dependent mechanism (Cassina et al., 2002) and via nerve growth factor (NGF) in p75NTR-expressing neurons (Pehar et al., 2004). It is also widely believed that mitochondrial dysfunction plays a major role, particularly as mitochondrial abnormality, vacuolation and swelling is observed in ALS patients and FALS mouse models (Dal Canto and Gurney, 1995; Kong and Xu, 1998;

Wong et al., 1995), and mutant SOD1 localises to spinal cord and brain mitochondria

(Liu et al., 2004; Vijayvergiya et al., 2005), where it impairs Cytochrome c association with the inner mitochondrial membrane and leads to apoptosis (Kirkinezos et al., 2005).

53 1.1.4.7 Involvement of cell signalling pathways

Protein kinases

Abnormalities in the activity and/or expression of several kinases have been reported in

ALS post-mortem tissue, including protein kinase C (PKC), phosphatidylinositol 3- kinase (PI(3)-K), Cdk5 and SAPK (Bajaj et al., 1998; Krieger et al., 1996; Lanius et al.,

1995; Migheli et al., 1997; Nagao et al., 1998; Wagey et al., 1998). Furthermore a recent study, using a quantitative proteomics screening technique on human thoracic spinal cord samples from ALS patients and controls, has identified increased expression of several kinases in ALS (Hu et al., 2003b), including PKC, ERK2 and its putative downstream target 1 (RSK1), phosphorylated/activated p38 MAPK

(p38), protein kinase B (PKB; also known as Akt), SAPK and Cdk5.

Mutant SOD1 transgenic mice have been useful in studying Cdk5-mediated cell death in

MND. In SOD1G37R mice, mislocalisation and hyperactivation of Cdk5 has been observed, which correlates with increased production of a truncation product (p25) of its activator p35 (Nguyen et al., 2001). The ways in which Cdk5/p25 activity may cause neurodegeneration are unknown, although aberrant phosphorylation of substrates such as NFs and tau, leading to cytoskeletal alterations/defects in axonal transport, has been suggested as a potential event leading to cell death. Additionally, Cdk5 is involved in several cellular processes, including cell adhesion and synaptic signalling (Bibb et al.,

2001; Kwon et al., 2000), alterations in both of which could contribute to neurodegeneration. Interestingly, mice overexpressing Cdk5 and p35 do not show a profound disruption of the cytoskeleton (Van den Haute et al., 2001), unlike mice overexpressing p25 (Ahlijanian et al., 2000), which suggests that p25 is the toxic mediator of events leading to neurodegeneration. Oxidative stress has also been found to promote the generation of p25 from p35 through activation of the Ca2+-dependent

54 protease calpain, resulting in increased Cdk5 activity, increased phosphorylation of NF subunits, and inhibition of axonal transport (Lee et al., 2000; Shea et al., 2004).

However, transgenic mice overexpressing SOD1G93A in a p35-null background are not phenotypically different from mice overexpressing SOD1G93A alone (Takahashi and

Kulkarni, 2004), which suggests that the production of p25 from p35 and the activation of Cdk5 by p35 is not involved in the SOD1G93A-mediated disease seen in these mice.

Therefore, the precise role of Cdk5/p25 activity in ALS is currently unclear.

The activity of p38 is increased in SOD1G93A mouse spinal cord (neurons, astrocytes and microglia), and this occurs before disease onset and correlates with disease progression, although no alterations in its activity have been observed in human ALS tissue compared with controls (Hu et al., 2003a; Hu et al., 2003b; Tortarolo et al., 2003). p38 has been implicated in exitotoxicity, as inhibitors of p38 have been found to rescue cells from glutamate-induced cell death (Kawasaki et al., 1997), and it is also involved in phosphorylation of cytoskeletal proteins and modulating the expression of cytokines, nitric oxide and COX-2 (Ackerley et al., 2004; Guan et al., 1998; Mielke and Herdegen,

2001; Ono and Han, 2000). Although active p38 plays a role in certain apoptotic pathways (Kummer et al., 1997) the p38 activation seen in the SOD1G93A mouse is not thought to lead to apoptosis, due to the lack of ultrastructural apoptotic features in nuclei, DNA fragmentation and activated caspase-3 immunostaining in these mice

(Bendotti et al., 2001; Migheli et al., 1999; Tortarolo et al., 2003).

L1CAM

Neuronal cell adhesion molecule L1 (L1CAM) is mutated in an X-linked form of HSP

(SPG1) and plays an important role in cell recognition and signalling (Jouet et al.,

1994). The L1CAM protein is a member of the immunoglobulin superfamily of cell

55 adhesion molecules and is found primarily in the nervous system where it is involved in cellular processes such as neurite/growth cone guidance and neuronal migration during development, and cell survival (Castellani et al., 2000; Chen et al., 1999; Dahme et al.,

1997; Fransen et al., 1998).

ALS2

Amyotrophic lateral sclerosis 2 (ALS2)/alsin was first identified in 2001 by two independent research groups as a protein that is mutated in rare autosomal recessive juvenile forms of ALS (ALS2) and PLS (jPLS) (Hadano et al., 2001; Yang et al., 2001), and has subsequently been found to cause infantile-onset HSP (IAHSP) (Devon et al.,

2003; Eymard-Pierre et al., 2002; Gros-Louis et al., 2003b). Although the exact function of ALS2 is currently unknown, it shares homology with several GTPase-regulating proteins (guanine nucleotide exchange factors; GEFs). It is predicted, therefore, to be involved in various critical cellular processes such as signal transduction, regulation of the cytoskeleton and intracellular trafficking, as discussed below.

1.2 Alsin/ALS2

Mutations in the gene ALS2, on chromosome 2q33.2, have been found that cause forms of infantile/juvenile MND with similar clinical phenotypes involving UMN degeneration: ALS (juvenile ALS; ALS2) (Hadano et al., 2001; Yang et al., 2001), PLS

(juvenile PLS; jPLS) (Yang et al., 2001) and HSP (Infantile-onset ascending HSP;

IAHSP) (Devon et al., 2003; Eymard-Pierre et al., 2002; Gros-Louis et al., 2003b).

However, there are as yet no reports of pathology in the affected individuals. Of the nine ALS2 mutations described so far, eight are small deletions and one is a nonsense mutation. All mutations discovered so far are in coding regions, are inherited in an autosomal recessive pattern, and are predicted to result in premature truncation and a

56 loss of function (Figure 1.1). There is no apparent correlation between the site of mutation and the disease phenotype (Devon et al., 2003). It has been found that at least six of the disease mutant proteins are unstable and rapidly degraded, including the mutation predicted to result in a C-terminal truncation of only 28 amino acid residues

(4844delT), therefore it may be speculated that the disease is caused by a loss of the whole protein, and any of its functional domains could potentially regulate its normal function (Yamanaka et al., 2003).

Human ALS2 comprises 34 exons (33 of which contain coding sequence) over a region of 6.5 kb and encodes the novel 184 kDa protein ALS2 (also known as Alsin). An additional transcript of 2.6 kb (predicted to be derived from alternative splicing), known as ALS2 ‘short-form’, has been detected by Northern blot (Hadano et al., 2001), although as yet no 44 kDa protein corresponding to the short form has been detected either in mouse or human tissue (Otomo et al., 2003). Furthermore, short-form ALS2 was found to be unstable and rapidly degraded when over-expressed in a human cell line (Yamanaka et al., 2003). Recently, a novel ALS2 homologous gene ALS2 carboxyl- terminal like (ALS2CL) on human chromosome 3p21.3, which encodes a 108 kDa protein, has been described (Devon et al., 2005; Hadano et al., 2004).

The function of ALS2 is unknown, although sequence analysis has revealed 3 domains that are homologous to guanine nucleotide exchange factors (GEFs), which suggests involvement in Ras superfamily signalling pathways (see section 1.3). The 3 GEF domains include an amino (N)-terminal regulator of chromatin condensation (RCC1)- like domain (putative Ran GEF), a diffuse B-cell lymphoma (Dbl) homology (DH) domain followed by a pleckstrin homology (PH) domain (which is a hallmark of Rho

GEFs), and a carboxyl (C)-terminal VPS9 domain (putative Rab GEF). There are also 8

57 copies of a sequence motif called membrane occupation and recognition nexus

(MORN), which may be involved in recruitment to membranes (Takeshima et al.,

2000). ALS2CL protein shows high amino acid sequence similarity to the C-terminal region of ALS2, but lacks certain residues corresponding to the RCC1-like domain

(Hadano et al., 2004).

1.2.1 Expression of ALS2

ALS2 mRNA distribution in various human tissues has been analysed using RT-PCR

(reverse transcription-polymerase chain reaction) (Yang et al., 2001) and Northern blot

(Hadano et al., 2001). Both long- and short-form ALS2 mRNA has been detected in heart, placenta, lung, liver, spleen, skeletal muscle, kidney and pancreas, with the highest expression seen in brain and spinal cord. Analysis of different brain regions reveals highest expression in cerebellum and cerebral cortex (Hadano et al., 2001).

Immunoblotting and immunohistochemistry shows that the pattern of ALS2 long-form protein expression is similar to that of its mRNA, with the highest-expressing tissues being brain (particularly cerebellum and cerebral cortex), spinal cord and liver, in both mouse and human (Devon et al., 2005; Otomo et al., 2003; Yamanaka et al., 2003).

Additionally, the use of neuron- and glia-specific markers (Neu-N and glial fibrillary acidic protein (GFAP) respectively) has revealed that ALS2 is expressed in various neurons, but not in glial cells (Devon et al., 2005; Otomo et al., 2003). ALS2 is believed to be expressed at generally low levels, however, representing only approximately

0.0003% of the total detergent soluble fraction of mouse brain lysate (Yamanaka et al.,

2003).

58 Figure 1.1 Schematic of ALS2 (Long-form and Short-form) and predicted disease mutants Nine ALS2 mutations have been described to date, all of which result in a premature stop codon and a predicted truncated protein.

ALS2 Long-form

RCC1 DH PH MORN VPS9

1 59 167 218 576 627 690 885 1049 1244 1551 1657 169 525 578 9011657 1007 1656

ALS2 Short-form

Unique 24 residues

1 372 396

ALS2 disease mutants (predicted proteins) Disease and Reference

261delA ALS2 (Tunisia) Unique 3 (Hadano et al., residues 2001)

1 49 1130delTA IAHSP (Italy) (Eymard-Pierre et al., 2002) 1 335

1548delAG Unique 70 ALS2/jPLS (Kuwait); residues (Hadano et al., 2001; Yang et al., 2001)

1 475 545

1594delGTTTCCCCCA IAHSP (France) (Eymard-Pierre et al., 2002) 1 493

1990delCT Unique 25 jPLS (Saudi) residues (Yang et al., 2001)

1 620 645

2660delCT IAHSP (Italy) (Eymard-Pierre et al., 2002) 1 858

59 C3115T IAHSP (Israel) (Devon et al., 2003)

1 998

3742delA IAHSP (Algeria) (Eymard-Pierre et al., 2002)

1 1206

4844delT IAHSP (Pakistan) (Gros-Louis et al., 2003b)

1 1573

1.2.2 ALS2 exhibits GEF activity

ALS2 has an amino-terminal domain that is homologous to RCC1, a known GEF for the

GTPase Ran, and ALS2 was found to weakly stimulate GDP dissociation from Ran in an in vitro assay, although preliminary experiments with the isolated RCC1-like domain were contradictory (Otomo et al., 2003). It has been suggested that it is unlikely that

ALS2 acts as a Ran GEF in vivo (Topp et al., 2004), as although there are more than 90 proteins with RCC1 domains on the available databases (Bateman et al., 2002), only

RCC1 itself exhibits Ran GEF activity (Bischoff and Ponstingl, 1991). Thus, the Ran

GEF activity of ALS2 remains unclear at this stage. Aside from a role as a GEF, the

RCC1 domain may function as a protein-protein interaction domain, since it has the potential to form a seven-bladed beta-propeller structure (Topp et al., 2004).

The presence of a DH and C-terminally adjacent PH domain is an indication that ALS2 may function as a GEF for the Rho family of GTPases, as this tandem repeat is found in the majority of Rho family GEFs (Hart et al., 1991; Yaku et al., 1994) (see section

1.4.1). In vitro binding assays have shown that a fragment of ALS2 containing only the

DH/PH domain has the ability to bind to Rac1 but not to Rac3, RhoA or Cdc42 (Topp et al., 2004), which has led to the investigation of whether ALS2 can function as a GEF

60 for Rac1 in vivo. When overexpressed in S. frugiperda (sf9) cells with Rac1, ALS2 promoted an increase in the level of GTP-bound Rac1 compared to control conditions

(Topp et al., 2004). In NIH3T3 cells endogenous ALS2 co-localised with Rac1 at leading membrane edges, and overexpressed GFP-tagged ALS2 co-localised with actin in membrane ruffles and lamellipodia, although overexpression did not seem to stimulate these events (Topp et al., 2004), which further implies a role for endogenous

ALS2 as Rac GEF. Furthermore, overexpression of ALS2 was found to increase the activity of endogenous Rac in CHO cells whereas a construct in which the DH/PH was mutated did not (These studies and Kanekura et al 2005). In contrast, ALS2 was not found to act as a Rac GEF in in vitro GDP dissociation assays (Otomo et al., 2003).

ALS2 contains a C-terminal VPS9 domain, which is homologous to Rab5 GEF domains. All VPS9 domain-containing proteins that have been studied so far exhibit

Rab5 GEF activity (Hama et al., 1999; Horiuchi et al., 1997; Saito et al., 2002; Tall et al., 2001) and this includes ALS2 (Otomo et al., 2003; Topp et al., 2004). In vitro GDP dissociation assays were used to test the Rab5 GEF activity of ALS2, and to identify the regions of the protein necessary for this activity (Otomo et al., 2003; Topp et al., 2004).

It was found that ALS2 activates Rab5a, Rab5b and Rab5c but not other Rab family members (Otomo et al., 2003), and that the minimal region necessary for this activity is the N-terminal region containing the VPS9 domain and the MORN motifs. In support of this data, in vitro binding assays using ALS2 immunoprecipitated from both CHO and

SH-SY5Y cells have shown that ALS2 binds to Rab5a (Otomo et al., 2003; Topp et al.,

2004). Furthermore, yeast 2-hybrid analysis using the VPS9 domain of ALS2 fused to the Gal4 activation domain (as ‘prey’) with various Rab5 LexA DNA binding domain fusions (‘baits’), revealed that ALS2 interacts with GDP-bound or nucleotide-free Rab5

(S34N; a mutant that cannot bind GTP) but not wild-type (GTP-bound) Rab5 (Topp et

61 al., 2004). ALS2 (endogenous and transfected) has been found to form homo-oligomers

(presumably octamers, based on data from gel exclusion assays) in COS-7 cells, and mutants which have lost the ability to form such oligomers also display a loss of Rab5

GEF activity in GDP dissociation assays (Kunita et al., 2004).

1.2.3 ALS2 regulates endosomal morphology

Endogenous ALS2 has been shown to localise to cytoplasmic punctuate membrane structures in rat embryonic hippocampal neurons (Topp et al., 2004) and biochemical analysis of rat cortex has revealed that endogenous ALS2 is found in the cytosolic and membrane fractions (co-precipitating with the recycling endosome marker transferrin receptor and the early endosome marker early endosome antigen 1 (EEA1)), where it is peripherally associated with endosomal membranes (Devon et al., 2005; Topp et al.,

2004; Yamanaka et al., 2003). Immunocytochemistry has also shown that overexpressed

ALS2 co-localises with transfected Rab5A and endogenous EEA1 in HeLa cells and cultured rat embryonic cortical neurons (Otomo et al., 2003). The localisation of ALS2 demonstrated in the above research is in agreement with the findings that ALS2 can act as a GEF for Rab5, as this GTPase is involved in protein trafficking through early endosomes (see section 1.3.2). Indeed, co-transfection of ALS2 (or the VPS9 domain of ALS2) and Rab5A in cortical neurons (or NIH3T3 cells) resulted in enlarged endosomes, which is similar to the effect seen upon overexpression of a constitutively active Rab5 (Stenmark et al., 1994), whereas constructs with mutations in the VPS9

(Rab5GEF) domain did not, implying a role for ALS2 in Rab5-mediated promotion of endosome fusion/enlargement (Otomo et al., 2003; Topp et al., 2004). Furthermore, a construct encompassing the DH/PH, MORN and VPS9 domains (ALS2660-1657) was found to act in a constitutively active manner by causing dramatic endosome enlargement and this was dependent on its ability to form homo-oligomers (Kunita et

62 al., 2004; Otomo et al., 2003).

Interestingly, Spartin, a protein that is mutated in a form of complicated HSP (Troyer syndrome), may be involved in endosome morphology and molecular trafficking events

(Patel et al., 2002; Bakowska et al., 2005), and mutations have recently been discovered in vesicle-associated membrane protein (VAMP)/synaptobrevin-associated protein B

(VAPB), which is involved in intracellular membrane trafficking, in seven families with diagnoses ranging from ALS (ALS8) to SMA (Nishimura et al., 2004).

1.2.4 ALS2 binds to mutant SOD1 and displays neuroprotective activity

The relationship between ALS2 and the first protein to be identified as causative for

MND, SOD1, has been investigated in the hope of identifying common mechanisms in disease pathogenesis. Overexpression of ALS2 long-form (but not ALS2 short-form) in the motor neuron-like cell line NSC34 delays cell death caused by SOD1A4T, SOD1G85R or SOD1G93R (as observed by trypan blue exclusion assay), and this neuroprotection is specific for mutant SOD1-mediated cell death, as ALS2 has no effect on toxicity caused by mutants of presenilin1 and 2 (PS1M146L and PS2N141I), amyloid beta precursor protein

(AβPPV642I) or α-synucleinA53T (proteins involved in familial Alzheimer’s disease and familial Parkinson’s disease) (Kanekura et al., 2004). Interestingly, ALS2 long-form has the ability to bind specifically to the aforementioned SOD1 disease mutants but not to wild-type SOD1, whereas ALS2 short-form can bind to wild-type SOD1 but not to the

SOD1 mutants (Kanekura et al., 2004). The region responsible for both the neuroprotective activity of ALS2 long-form and its binding to mutant SOD1 has been found to reside in the DH/PH (Rho family GEF) domain (Kanekura et al., 2004).

Moreover, the neuroprotective activity of ALS2 can be blocked by reduction of Rac1 expression (using small interfering RNA (siRNA) for Rac1) which further supports the

63 hypothesis that ALS2 functions as a Rac GEF in this pathway. Rac has previously been shown to promote cell survival in COS-7 cells by activation of PI(3)K and Akt (Murga et al., 2002), and it has been shown that PI(3)K and Akt family proteins may be involved in ALS2-mediated neuroprotection against mutant SOD1. Indeed, treatment with wortmannin (a PI(3)K inhibitor) and reduction of Akt3 expression (using siRNA) independently abolished ALS2 neuroprotection (Kanekura et al., 2005). Additionally, overexpression of ALS2 but not a mutant without Rac GEF activity (ALS2T701A) resulted in increased phosphorylation of Akt3. Therefore, these findings suggest that

ALS2 mediates neuroprotection specifically against mutant SOD1 by sequential activation of Rac1, PI(3)K and Akt3, although whether this function of ALS2 occurs in neuronal cells and by the endogenous protein remains to be discovered.

1.3 The Ras Superfamily of GTPases

Members of the Ras superfamily of small guanosine triphosphatases (small GTPases; also known as ‘small G proteins’) are monomeric proteins that bind and hydrolyse guanine nucleotides (Bourne et al., 1991). By cycling between

(GTP)-bound (active) and guanosine diphosphate (GDP)-bound (inactive) conformations, Ras GTPases act as molecular switches, controlling a wide range of intracellular signalling pathways in all . To date, more than 80 Ras superfamily members have been identified in mammals (Scita et al., 2000), and are divided into five families: Ras, Rho, Rab, Arf and Ran (see Table 1.4), all of which possess a common structurally-conserved GTP-binding region known as the ‘G domain’

(Bourne et al., 1990; Valencia et al., 1991).

64 Table 1.4 Mammalian Ras GTPase Superfamily

RAB FAMILY RAS FAMILY RHO FAMILY ARF FAMILY RAN FAMILY Rab1A, B Ha-Ras RhoA, B, C, D Arf1- 6 Ran Rab2A, B Ki-Ras RhoE//Rho8 Sar1a, b Rab3A, B, C, D N-Ras RhoG Arl1- 7 Rab4A, B, C R-Ras RhoH/TTF Ard1 Rab5A, B, C M-Ras Rac1, 2, 3 Rab6A, B, C RalA, B Cdc42 Rab7 Rap1A, B /Rho6 Rab8A, B Rap2A, B /Rho7 Rab9A, B, C TC21 TC10 Rab10 Rit TCL Rab11A, B Rin Rab12-15 Rad Chp/Wrch2 Rab17, 18 Kir/ Gem Rif Rab20, 21 Rheb RhoBTB1, 2 Rab22A, B, C κB-Ras1, 2 MIRO-1, 2 Rab23 Rab25, 26 Rab27A, B Rab28A, B Rab30 Rab32A, B Rab34-39 Rab40A, B, C Rab41-43

(Burridge and Wennerberg, 2004; Stenmark and Olkkonen, 2001; Takai et al., 2001; Wherlock and Mellor, 2002)

Ras itself regulates cell growth and differentiation by activation of the MAPK cascade, whereas the other superfamily members are involved in a variety of cellular events. The

Rho (in particular Rho, Rac and Cdc42) and Rab (in particular Rab5) family GTPases, which are implicated in ALS2 function, will be discussed below.

65 The activity (nucleotide binding state) of GTPases is highly regulated. In contrast to the moderate hydrophobicity of other Ras superfamily members, Rho and Rab family

GTPases contain highly hydrophobic geranylgeranyl moieties and therefore the association of guanine nucleotide dissociation inhibitors (GDIs) is needed to stabilise

GDP-binding and sequester the inactive GTPase to the cytoplasm. Indeed, the majority of GTPase in the cell is found in the inactive GDP-bound form, associated with GDIs in the cytoplasm. An additional regulating protein known as a GDI-displacement factor

(GDF) is required for the correct membrane localisation of Rab GTPases. The active state is promoted by guanine nucleotide exchange factors (GEFs) which tether the

GTPase to a particular subcellular localisation and promote GDP release (this results in

GTP binding due to a higher intracellular level of GTP than GDP). The active state is then negatively regulated by GTPase activating proteins (GAPs), which catalyse the intrinsic ability of the GTPase to hydrolyse GTP to GDP (Figure 1.2) (Moon and

Zheng, 2003; Schmidt and Hall, 2002; Zheng, 2001)

Figure 1.2 Schematic diagram of the small GTPase activation cycle

Activation Signal

GDP GTP

GDI GEF

GDP GTP

GTPase GTPase

GDI Effector GAP GDI Cytoskeletal organisation; Apoptosis; ; Cell cycle Pi progression; Membrane traffic; ; Vesicular & nuclear transport

66 1.3.1 Rho family GTPases

Rho family GTPases are ubiquitously expressed and so far 21 genes have been identified in humans, encoding at least 23 signalling proteins (Wherlock and Mellor,

2002). More than 40 effectors, 60 GEFs and 40 GAPs have been described for the mammalian Rho family (Raftopoulou and Hall, 2004). Rho family GTPases are key regulators of multiple cellular activities including regulation of the cytoskeleton and cell adhesion, cell polarity, cell cycle progression, apoptosis, neuronal axon guidance, differentiation, oncogenesis and gene transcription.

The best characterised function of Rho family GTPases is the control of signal transduction pathways linking membrane receptors to regulation of the actin cytoskeleton (for reviews see (Van Aelst and D'Souza-Schorey, 1997; Hall, 1998)).

Such research has focused on the roles of three Rho GTPase family members, Rho (A,

B and C), Rac (1 and 2) and Cdc42, in the spatiotemporal control of cytoskeletal dynamics. Although structurally related, Rho, Rac and Cdc42 have been shown to have distinct effects on the actin cytoskeleton. In classic studies using quiescent Swiss 3T3 fibroblasts (a cell line in which serum starvation results in a very low background of filamentous (F)-actin), Rho induces stress fiber formation and clustering of into focal adhesion complexes (Ridley and Hall, 1992), Rac promotes actin polymerisation at the cell periphery to form lamellipodia and membrane ruffles (Ridley et al., 1992), and Cdc42 induces the formation of microspikes or filopodia (Nobes and

Hall, 1995). Furthermore, Rho GTPases have been found to play essential roles in a variety of cellular events which involve rearrangements of the actin cytoskeleton including cell migration, phagocytosis, axonal and dendritic growth and guidance, and synaptogenesis (Allen et al., 2000; Bradke and Dotti, 1999; Jin and Strittmatter, 1997;

May and Machesky, 2001; Murphy and Montell, 1996; Santos et al., 1997; Tashiro and

67 Yuste, 2004; Thies and Davenport, 2003; Yamashita et al., 1999; Zipkin et al., 1997). It is now generally accepted that Rho exerts its effects on the actin cytoskeleton by means of actin- contractility, although the precise molecular mechanisms remain to be identified, whereas Rac and Cdc42 have been shown to activate members of the

Wiskott-Aldrich syndrome protein/WASP Verprolin-homologous protein

(WASP/WAVE) family, which stimulate the actin-related protein 2/3 (Arp2/3) complex to promote actin polymerisation (Machesky and Insall, 1999; Miki et al., 2000).

Activation of Rho family GTPases can be mediated by activation of numerous receptors including growth factor receptors (Hall, 1998), cell adhesion receptors such as

(DeMali et al., 2003; Price et al., 1998), cadherin (Braga, 2002) and immunoglobulin superfamily members (Thompson et al., 2002), and G-protein-coupled receptors

(Collins et al., 1996; Coso et al., 1996).

Rho family effectors

Each Rho family GTPase interacts with multiple effectors and several effectors are recognised by multiple family members. A multitude of effectors have been identified, reflecting the diversity and complexity of Rho family GTPase signalling. A selection of

Rho family effectors are shown in Table 1.5.

A motif consisting of 16 amino acids (ISXPXXXXFXHXXHVG) known as the Cdc42/

Rac Interactive Binding (CRIB) domain (also known as p21 binding domain; PBD) has been identified as a region required for binding to Rac and/or Cdc42 in their GTP- bound (active) form, and was first identified in p65PAK (also known as p21 activated kinase/PAK1-6) (Burbelo et al., 1995). Subsequently, more than 25 CRIB domain- containing proteins (from a wide variety of eukaryotic species) have been identified

68 from GenBank database searches, including the mixed lineage kinases (MLK), WAVE, the activated Cdc42-associated tyrosine kinase (ACK) family of non-receptor tyrosine kinases, and marrow stromal/endothelial cell protein (MSE55) (see Table 1.5) (Burbelo et al., 1995). PAK has been implicated in Rac/Cdc42-mediated cytoskeletal rearrangements, promoting formation of polarised filopodia and membrane ruffles, and increasing motility, in fibroblasts (Sells, 1999; Sells et al., 1997), and leading to the loss of stress fibers and focal adhesions in HeLa and Swiss 3T3 cells (Manser et al., 1997).

However, such observations are complex, and often cell-type specific (for a recent review see (Bokoch, 2003)). For example, Y40C mutants of Rac and Cdc42, which are unable to bind to CRIB domain-containing proteins, are still able to induce actin reorganisation, suggesting that Rac/Cdc42 signalling to the actin cytoskeleton may be mediated by proteins other than PAK (Westwick et al., 1997), such as Partner of Rac1

(POR1) (Ishizaki et al., 1996; Van Aelst et al., 1996).

Table 1.5 Selected effectors of the Rho GTPase family

GTPase Effectors

Rho PI(3)K; PI(4,5)K; phospholipase D; rhophilin; kinectin; rhotekin; DGKζ, PRK1/PKN, PRK2, MBS, ROK/ROCK/Rho kinase; Bni1; Bnr1; PKC; p140mDia; Fks1; Fks2 ACK; PAK; PI(3)K; WASP; N-WASP; S6-kinase; MLK2; MLK3; IQGAP; MRCK; Cdc42 MSE55; Skm1; Gic1; Gic2; Borg; Bni1; Ste20; Cla4; IRSp53 NADPH oxidase; PAK; PI(3)K; PI(4,5)K; DGK; POSH; IQGAP; MLK2; MLK3; Rac MSE55; POR1; Sra-1; S6-kinase; IRSp53 (Aspenstrom, 1999; Bourne et al., 1990)

Rho, Rac and Cdc42 regulate gene expression in mammalian cells, for example by activating the serum response factor (SRF) (Hill et al., 1995). This activation is believed to be due to Rho family member-mediated changes in actin dynamics (Gineitis and

Treisman, 2001; Mack et al., 2001; Sotiropoulos et al., 1999). Rac and Cdc42 have also been shown to activate the SAPK/Jun N-terminal kinase (JNK) and p38 pathways,

69 which modulate processes such as gene transcription, apoptosis, development, transformation, immune activation and inflammation in response to toxins, physical stress and inflammatory cytokines (Coso et al., 1995; Minden et al., 1995). For a SAPK review see (Tibbles and Woodgett, 1999). Several studies suggest that binding/activation of PAK is necessary for activation of the SAPK pathway by

Rac/Cdc42 (Bagrodia et al., 1995; Joneson et al., 1996; Lamarche et al., 1996; Zhang et al., 1995). However there is also evidence suggesting that Rac can activate the SAPK pathway independently of PAK (Teramoto et al., 1996b; Westwick et al., 1997). In addition to PAK, two targets of Rac that have been reported to contribute to SAPK activation are MLK3 (Teramoto et al., 1996a) and the scaffold protein plenty of SH3s

(POSH) (Tapon et al 1998).

Crosstalk between Rho family members

Cross-talk between Rho GTPase family members (Rho, Rac, and Cdc42) is well- documented and known to play an important role in modulating and coordinating downstream cellular responses resulting from Rho GTPase signalling. In quiescent

Swiss 3T3 fibroblasts, a hierarchical linear cascade of activation has been identified, in which Cdc42 activates Rac and Rac activates Rho, resulting in the successive formation of filopodia, membrane ruffles and stress fibers (Etienne-Manneville and Hall, 2002;

Nobes and Hall, 1995). However, in other cell systems different Rho family members have been found to antagonise each other. For example, in N1E-115 neuroblastoma cells, Rho results in neurite retraction and cell rounding whereas Rac promotes cell spreading and neurite outgrowth (Leeuwen et al., 1997), and preferential activation of either Rac or Rho inhibits the opposite phenotype. Similarly, inhibition of Rho in Swiss

3T3 or Rat1 fibroblasts induces Rac-associated phenotypes (small focal contacts, cell spreading and motility) and vice versa with inhibition of Rac (Arthur and Burridge,

70 2001; Rottner et al., 1999). The exact mechanism(s) by which Rac and Rho pathways antagonistically interact with each other are unknown, although it has been proposed to involve either cross-regulation of downstream GTPase effector pathways (Sanders et al.,

1999; van Leeuwen et al., 1999), or direct disruption of GTPase conformation (resulting in modulation of its ‘switch’ mechanism of activation) (Caron, 2003; Nimnual et al.,

2003). The downstream Rac effector PAK1, for example, has been shown to inhibit Rho signalling (actin-myosin contractility) through its phosphorylation of myosin-II heavy chain and myosin light chain kinase (Sanders et al., 1999; van Leeuwen et al., 1999), and it has recently been reported to negatively-regulate the activity of Net1 (a Rho GEF) resulting in a downregulation of Rho activity in vitro (Alberts et al., 2005). A novel mechanism has also been described recently, in which production of ROS via Rac- dependent activation of NADPH oxidase (which is already known to contribute to host defence/microbial killing, cell cycling, transformation and gene transcription) (Bokoch and Diebold, 2002; Joneson and Bar-Sagi, 1998; Kheradmand et al., 1998; Suh et al.,

1999) causes an inhibition of the low molecular weight protein tyrosine phosphatase

(LMW-PTP) leading to activation of p190RhoGAP (which inactivates Rho), and this mechanism has been shown to be both necessary and sufficient for Rho activity downregulation in fibroblasts (Nimnual et al., 2003). Interestingly, the generation of

ROS by Rac is negatively regulated by Cdc42, although the significance of this finding in relation to Rho inhibition has not yet been investigated (Diebold et al., 2004).

Another Rho family member, RhoG, has been found to activate both Rac and Cdc42 in independent mechanisms (Gauthier-Rouviere et al., 1998), and a signalling pathway involved in RhoG-mediated activation of Rac has recently been described, in which activated RhoG binds engulfment and cell motility 1 (ELMO1) and this activates the

Rac-GEF dedicator of cytokinesis 180 () leading to Rac activation (Katoh

71 and Negishi, 2003), although the significance of this finding is controversial (Prieto-

Sanchez and Bustelo, 2003; Wennerberg et al., 2002).

1.3.2 Rab family GTPases

Rab GTPases are key regulators of vesicular transport in eukaryotic cells, and are involved in most if not all aspects of transport including vesicle formation, targeting and docking; and membrane remodelling and fusion. Most Rab proteins are ubiquitously expressed but some are tissue specific, for example Rab3 expression is restricted to neurons. Over 60 Rab proteins have been discovered in humans (Bock et al., 2001) and several of these are known to participate in more than one intracellular transport step, including Rab1, Rab5 and Rab11. For example, Rab5 has important roles in vesicle formation at the plasma membrane (McLauchlan et al., 1998), microtubule-dependent vesicle motility (Nielsen et al., 1999), membrane remodelling (through interaction with

PI(3)K) (Christoforidis et al., 1999b) and endosome fusion (Gorvel et al., 1991;

Stenmark et al., 1994).

Rab5 effectors

A variety of effectors have been described for Rab5, which reflects its role in numerous stages of vesicular transport steps (see above). 22 proteins that specifically bind to GTP- bound Rab5 were identified from bovine brain (Christoforidis et al., 1999a) and several of these are now known to be Rab5 effectors including Rabaptin-5, EEA1, Rabenosyn,

PI(3)K, Rabkinesin-6, tail interacting protein 47 (TIP47) and yeast Vac1 (Segev, 2001).

Interestingly, Rabaptin-5 also binds to both Rab4 and the Rab3 effector Rabphilin3, which suggests crosstalk exists between endocytic and exocytic pathways.

72 Crosstalk between Rab5 and Rho family members

Several observations have suggested that there could be crosstalk between Rho family

GTPases and Rab5. Rab5 has been reported to affect Rho family GTPase-dependent cytoskeletal changes. For example, dominant negative Rab5 inhibits reconstruction of the actin cytoskeleton after phorbol myristate acetate (PMA) treatment and constitutively active Rab5 induces actin remodelling; both processes are dependent on

Rac and Rho (Imamura et al., 1998; Spaargaren and Bos, 1999). Furthermore, Rab5 is essential for a form of receptor tyrosine kinase (RTK)-induced actin remodelling called

‘circular ruffling’ (a membrane ruffling event distinct from RTK-Ras-Rac-mediated lamellipodia formation, that is important in macropinocytosis and 3-dimensional migration, and is also dependent on activation of PI(3)K and Rac), which further suggests an involvement of Rab5 in Rho family GTPase signalling pathways (Lanzetti et al., 2004). Crosstalk between Rho family GTPase-mediated signalling to the actin cytoskeleton and Rab5-mediated endocytic events is also suggested by increasing evidence of involvement of actin reorganisation in endocytosis (Qualmann et al., 2000), and the involvement of several Rho family GEFs in various endocytic trafficking events. For example, activated Rac or Rho inhibits transferrin receptor endocytosis in

HeLa cells (Lamaze et al., 1996), RhoD localises to early endosomes and alters their distribution and motility (Murphy et al., 1996), RhoB localises and activates its effector protein kinase C-related kinase 1 (PRK1; also known as protein kinase N (PKN)) on endosomes and this retards trafficking of internalised EGF receptor from endosomes to a pre-lysosomal compartment (Gampel et al., 1999; Mellor et al., 1998), Rac1 can directly interact with Synaptojanin 2 (a polyphosphoinositide phosphatase implicated in the uncoating of clathrin-coated vesicles) (Malecz et al., 2000), and Intersectin 1 (a scaffolding protein regulating the formation of clathrin-coated vesicles in endocytosis)

73 (Guipponi et al., 1998) has been recently shown to exhibit Cdc42 GEF activity in cultured cells (Hussain et al., 2001).

1.4 Guanine nucleotide exchange factors (GEFs)

Small GTPases are highly regulated proteins and guanine nucleotide exchange factors

(GEFs) are the principle mediators of their activation. They achieve this in 2 ways: by destabilising GDP-GTPase interaction leading to GDP release, and by stabilising this nucleotide-depleted transition state, enabling GTP (which is at an approximately 10-fold higher concentration in the cell than GDP) to bind to the GTPase.

1.4.1 DH/PH (Dbl-family) GEFs

Most GEFs responsible for activating Rho family GTPases share two common motifs, the diffuse b-cell lymphoma (Dbl) homology (DH) and a C-terminally adjacent pleckstrin homology (PH) domain (together known as the DH/PH domain). To date, 6

DH/PH-containing GEFs have been identified in Saccharomyces cerevisiae, 18 in

Caenorhabditis elegans, 23 in Drosophila melanogaster and over 60 in humans

(Schmidt and Hall, 2002; Venter et al., 2001). The DH domain (first identified in the oncogenic product Dbl as the minimal region exhibiting Cdc42 GEF activity) is the catalytic domain necessary for GEF activity (Hart et al., 1994) whereas the PH domain has been proposed to function in localisation to the plasma membrane, as PH domains are known to bind to both phosphorylated phosphoinositides (PIPs) and proteins

(Lemmon and Ferguson, 2000; Rebecchi and Scarlata, 1998). This function is supported by the finding that the PH domain can be substituted with a membrane-targeting signal in certain GEFs such as Lymphoid blast crisis (Lbc)’s first cousin (Lfc) (Whitehead et al., 1995). However, the role of PH domains in Dbl-family GEFs remains controversial, as they have also been reported to participate in GEF activity/GTPase binding in Dbl’s

74 big sister (Dbs) and Trio (Liu et al., 1998b; Rossman et al., 2002) and in contrast, to inhibit GEF activity by masking access to the DH domain in several GEFs including

Vav and 1 (Sos1) (Han et al., 1998; Nimnual et al., 1998).

Furthermore, some GEFs such as T-lymphoma invasion and metastasis inducing protein

1 (Tiam1) and RasGRF contain a second (N-terminal) PH domain, and it is this PH domain that is required for membrane localisation rather than the PH of the DH/PH motif (Buchsbaum et al., 1996; Michiels et al., 1997; Stam et al., 1997).

Rho family GEFs without the classic DH/PH tandem motif have also recently been described. Examples include the novel Rac-specific GEF SWAP-70 which has a PH domain that is N-terminal to the DH domain (a PH/DH domain) (Shinohara et al.,

2002), and the Rac GEF DOCK180 which does not contain a DH/PH domain

(Kiyokawa et al., 1998). DOCK180 displays Rac GEF activity which has been mapped to a region called the ‘Docker/Dedicator of cytokinesis’ domain (Brugnera et al., 2002) and it requires a cofactor protein (ELMO) for its function (Brugnera et al., 2002).

Additionally, over 10 members of this unconventional GEF family have been identified

(Cote and Vuori, 2002; Meller et al., 2002).

Specificity of Rho family GEF activity

Several GEFs show specificity towards a particular Rho family member, for example

Faciogenital dysplasia protein (Fgd1) acts as a Cdc42 GEF and p115RhoGEF is a Rho

GEF, whereas others have the ability to activate several different family members, such as Dbl, Vav, and Epithelial cell transforming sequence 2 (Ect2). Interestingly, GEF specificity often varies according to the experimental conditions, for example Tiam1 is a GEF for Rho, Rac and Cdc42 in vitro but only for Rac in vivo (Michiels et al., 1995).

Furthermore, certain GEFs contain more than one DH/PH domain, for example

75 mammalian and Drosophila Trio, its mammalian orthologue Kalirin, its C. elegans orthologue Uncoordinated locomotion (UNC-73), each contain two DH/PH domains, one preferentially activating Rac1/RhoG and the other preferentially activating Rho

(Bellanger et al., 1998; Debant et al., 1996; Penzes et al., 2000; Penzes et al., 2001;

Steven et al., 1998).

Most GEFs contain diverse functional domains in addition to the DH/PH module, which may be involved in their regulation, coupling to upstream receptors or signalling proteins, or coupling to effectors; this also suggests that many GEFs may be multifunctional by nature and participate in and/or integrate several signalling pathways. Interestingly, both alpha-PAK-interacting exchange factor (α-PIX) and Fgd1 have been shown to possess functions that are independent from their GEF activity

(Daniels et al., 1999; Nagata et al., 1998). Additional functional domains identified in

Rho family GEFs so far include Src homology 2 (SH2), Src homology 3 (SH3), Ser/Thr kinase, Tyr kinase, Ras-GEF, Rho-GAP, Ran-GEF, PSD-95/DlgA/ZO-1 (PDZ),

Regulator of G-protein signalling (RGS), and additional PH domains. ALS2 is the only

Dbl-family GEF containing an RCC1-like domain, a VPS9 (Rab5 GEF) domain and

MORN motifs, that has been described to date.

Regulation of Rho GEF activity

Like their GTPase targets, GEFs are also highly regulated proteins, and there are numerous ways in which this regulation is achieved. GEFs are sometimes negatively- regulated by intramolecular interactions between the PH domain or other regulatory domains, which mask the catalytic DH domain, preventing binding and activation of

GTPases. This is seen in several GEFs including Vav, Sos1, Dbl, Lbc and p115RhoGEF. This allosteric inhibition can be relieved by several mechanisms

76 including binding of PtdIns(3,4,5)P3 to the PH domain, or tyrosine phosphorylation

(both of which have been shown to stimulate Vav activity) (Aghazadeh et al., 2000;

Bustelo, 2000; Han et al., 1998). In contrast however, binding of either PtdIns(3,4,5)P3 or PtdIns(4,5)P2 to the PH domain of Dbl inhibits its Cdc42 GEF activity (Russo et al.,

2001). Numerous GEFs can also be activated by subunits, and this binding is presumed to activate the GEF by relieving autoinhibition. For example, binding of

Gα13 to the RGS domain of p115RhoGEF activates its Rho GEF activity (Hart et al.,

1998). Binding to certain signalling or adapter proteins are also thought to have various effects on GEF activity. For example, a novel protein has recently been discovered,

TRIPalpha, which inhibits the Rho-specific GEF domain of Trio (Schmidt et al., 2002).

1.4.2 Rab GEFs

Several mammalian Rab GEFs (and one putative Rab GEF; ALS2) have been described to date (although more GEFs for the Ypt GTPases, which are yeast Rab homologues, are known, one of which is Vps9). Examples of Rab GEFs are Rabex-5 (also known as

RabGEF1; a GEF for Rab5) (Horiuchi et al., 1997), MAP kinase activating death domain (Madd; also known as Rab3 guanine nucleotide exchange protein (Rab3GEP); a

GEF for Rab3A) (Wada et al., 1997) and RIN1,2,3 (Kajiho et al., 2003; Tall et al.,

2001). The majority of Ypt/Rab GEFs are part of large protein complexes within the cell, for example Rabex-5 forms a complex with Rabaptin-5 (a Rab5 effector), although the significance of such interactions is unknown (Horiuchi et al., 1997).

77

CHAPTER 2: MATERIALS AND METHODS

78 Unless stated otherwise, all molecular biology and cell culture reagents were purchased from Invitrogen Ltd. (UK) and all other chemicals were purchased from Sigma-Aldrich

Company Ltd. (UK). Stock solutions and buffers were prepared using ultrapure H2O from an Elga Maxima purification system. When required, solutions were sterilised either by autoclaving for 15 minutes at 15 lb/square inch or by filtration through a 0.2

µm pore filter (BDH).

2.1 Materials

2.1.1 Stock solutions

Acrylamide-bis-acrylamide (37:5:1 stabilised solution; National Diagnostics, USA)

Ampicillin (Amp; 100 mg/ml, filter sterilised)

Ammonium persulphate (APS; 20% w/v)

Adenosine 5′-triphosphate (ATP; 10 mM; Pharmacia, UK)

5-Bromo-4-chloro-3-indolyl-β-D-galactoside (X-gal; 20 mg/ml in N, N-dimethyl- formamide)

Chloroform (BDH): iso-amyl alcohol (IAA, Fisons) (24:1 v/v)

4′, 6-Diamidino-2-phenylindole (DAPI; 0.5 mg/ml in Dimethyl sulphoxide (DMSO))

Dithiothreitol (DTT; 1 M) dNTPs (25mM of each, 100 mM final; Pharmacia, UK)

Ethanol (70% v/v)

Ethylenediaminetetraacetic acid (EDTA; 0.5 M)

Ethylene glycol-bis(2 aminoethylether)-N,N,N’,N’-tetraacetic acid (EGTA; 0.5 M)

Ethidium bromide (10 mg/ml)

Glucose (20% w/v, filter-sterilised)

Glycerol (50% v/v, autoclaved)

79 HEPES (pH 7.0, 0.5 M)

Isopropyl β-D-thiogalactoside (IPTG; 100 mM)

Kanamycin (Kan; 10 mg/ml, filter-sterilised)

Magnesium chloride (MgCl2; 0.5 M)

Magnesium sulphate (MgSO4; 1 M, autoclaved)

Paraformaldehyde (PFA; 4% w/v in PBS)

Phenol (Fisons): chloroform: iso-amyl alcohol (25:24:1 v/v/v)

Phenylmethysulphonyl fluoride (PMSF; 100 mM)

Phosphate-buffered saline (PBS; Potassium phosphate 1.5 mM, Sodium phosphate 8.1 mM, NaCl 140 mM, Potassium chloride 2.7 mM)

Potassium acetate (C2H3KO2; 5 M)

RNase A (20 mg/ml)

Sodium azide (2% w/v)

Sodium dodecyl sulphate (SDS; 10% w/v)

Sodium chloride (NaCl; 4 M)

Sodium fluoride (NaF; 1 M)

Sodium hydroxide (NaOH; 10 M)

Sodium orthovanadate (Na2VO4; 20 mM, activated, pH 10.0)

Tris (hydroxymethyl) amino methane (Tris-HCl, buffered with HCl or NaOH to pH range 6.8-8.8)

Tris-buffered saline (TBS, pH 7.6; Tris hydroxymethyl amino methane 100 mM, NaCl

150 mM, buffered with HCl to pH 7.6)

Tris-EDTA (TE; Tris 10 mM, EDTA 1 mM; buffered with NaOH to pH 8.0)

Trypsin-EDTA (1x liquid; 0.25% in 1 mM EDTA)

Urea (6 M in PBS)

80 2.1.2 General molecular biology reagents

2.1.2.1 Plasmids

See Tables 2.1 and 2.2 below.

Table 2.1 Vectors

Vector Vector use Manufacturer

pBluescript SK+ Cloning vector Stratagene pCIneo Mammalian expression vector Promega pGEX-5X-1 Bacterial GST fusion protein expression vector Pharmacia

Table 2.2 Mammalian expression plasmids

Protein expressed Vector backbone Reference or source of plasmid Chloramphenicol pCIneo (Irving and Miller, 1997) acetyl transferase (CAT) Enhanced green fluorescent pEGFP.C1 Clontech protein (GFP) Myc-tagged ALS2 pCIneo N. J. O. Jacobsen (IOP) ALS2ala pCIneo M. S. Perkinton (IOP) DH/PH region of ALS2 pRK5myc A. Schmidt (University College London) (ALS2(DH/PH)) Rac1 pRK5myc A. Schmidt (University College London) N17 Rac1 pRK5myc A. Schmidt (University College London) L61 Rac1 pRK5myc A. Schmidt (University College London) RhoA pRK5myc A. Schmidt (University College London) Cdc42 pRK5myc A. Schmidt (University College London) N39 Rab5 pRK5myc A. Schmidt (University College London) VavΔN pRK5myc A. Schmidt (University College London) Net1ΔN pRK5myc A. Schmidt (University College London) Fgd1(DH/PH) pRK5myc A. Schmidt (University College London) p21 associated kinase-1 pCMV6myc S. Kesavapany (NIH, Bethesda, USA) (PAK1)

2.1.2.2 Primers

Unless indicated otherwise, all primers were obtained from Oswel DNA service,

University of Southampton. Sequencing of inserts in pBluescript SK+ was carried out

81 using the T3 and T7 primers:

T3 primer: 5'-AATTAACCCTCACTAAAGGG-3'

T7 primer: 5'-CGGGATATCACTCAGCATAATG-3'

Novel primers used for mutagenesis, sequencing and generating PCR products are described in the relevant sections below.

2.1.2.3 Growth of E. coli for DNA purification: media

Luria Bertani (LB) broth (powder; Invitrogen), 20 g/L sterilised by autoclaving

LB agar (powder; Invitrogen), 32 g/L sterilised by autoclaving

LB-amp broth (LB broth; ampicillin 100 µg/ml)

LB-amp agar (LB agar; ampicillin 100 µg/ml added when agar is less than ~55°C)

LB-kan broth (LB broth; kanamycin 50 µg/ml)

LB-kan agar (LB agar; kanamycin 50 µg/ml added when agar is less than ~55°C)

Bacteria containing plasmids of interest were stored at –70°C in sterile 25% glycerol

(v/v) solution in LB-amp/kan broth.

2.1.2.4 Plasmid DNA preparation from E. coli

Alkaline lysis solutions for small-scale purification of plasmid DNA from E. coli were used as previously described (Sambrook et al., 1989).

Solution 1 (sterilised by autoclaving)

50 mM Glucose

25 mM Tris pH 8.0

10 mM EDTA

82 Solution 2 (freshly-prepared from stock solutions)

0.2 M Sodium hydroxide

1% SDS (w/v)

Solution 3

5 M Potassium acetate

Glacial acetic acid

(final solution is 3 M with respect to potassium and 5 M with respect to acetate)

2.1.2.5 Agarose gel electrophoresis of nucleic acids

Solution for loading DNA samples in agarose gels

0.25% Bromophenol blue (w/v)

40% Sucrose (w/v)

Tris-Acetate-EDTA (TAE)

40 mM Tris-acetate pH 8.0

2 mM EDTA

Nucleic acid size markers

Phage λ DNA digested with Hind III (sizes in base pairs): 23,131; 9,419; 6,434; 4,335;

2,322; 2,023; 564; 125.

ΦX174 digested with Hae III (sizes in base pairs): 1,353; 1,078; 872; 603; 310; 281;

271; 234; 194; 118; 72.

2.1.2.6 Polymerase chain reaction (PCR) enzymes

Pyrococcus furiosus (Pfu) DNA polymerase (Stratagene)

83 Thermophilus aquaticus (Taq) DNA polymerase (Boehringer Mannheim)

2.1.3 Site-directed mutagenesis

ExSite™ PCR-based site-directed mutagenesis kit was purchased from Stratagene.

XL1-Blue supercompetent cells were provided with the kit. Details of specific primers used during work towards this thesis are given in the relevant results sections.

The following materials are supplied with the ExSite™ PCR-based site-directed mutagenesis kit:

Oligonucleotide control primer #1 (5’ phosphorylated; 75 ng/µl):

5’-CGCGCTTGGCGTAATCATGGTCAT -3’

Oligonucleotide control primer #2 (75 ng/µl):

5’-AGTACTCAATTAACCCTCACTAAAGGGAAC-3’ pWhitescript™ 4.5 kb control template (1.5 µg/µl) pUC18 transformation control plasmid (0.1 ng/µl in TE buffer)

10x mutagenesis buffer

200mM Tris-HCl (pH 8.8)

100mM KCl

100mM (NH4)2SO4

20mM MgSO4

1% Triton X-100 (v/v)

1 mg/ml BSA

Nucleotide mix

25 mM dATP

84 25 mM dCTP

25 mM dGTP

25 mM dTTP

2.1.4 Purification of GST fusion proteins

Glutathione elution buffer

10 mM reduced glutathione

50 mM Tris-HCl pH 8.0

2.1.5 Protein analysis

2.1.5.1 Protein sample preparation

Cell lysis buffer

50 mM Tris-HCl pH 7.6

150 mM NaCl

1 mM EDTA pH 8.0

1% Triton® X-100

Complete protease inhibitor cocktail (Roche)

Tissue homogenisation buffer

25 mM Tris-HCl pH 7.5

50 mM NaCl

5 mM MgCl2

1% Nonidet P40 (NP-40) (v/v)

5% Sucrose (w/v)

Complete protease inhibitor cocktail (Roche)

1 mM DTT

85 20 mM NaF

0.5 mM Na orthovanadate

1 mM PMSF

Magnesium lysis buffer

25 mM HEPES pH 7.5

150 mM NaCl

1% NP-40

10 mM MgCl2

1 mM EDTA

10% glycerol

1 mM NaF

1 mM Na orthovanadate

EDTA-free Complete protease inhibitor cocktail (Roche)

Cell fractionation homogenisation buffer

10 mM Tris pH 7.4

10 mM NaCl

3 mM MgCl2

1 mM EDTA

1 mM EGTA

1 mM PMSF

Complete protease inhibitor cocktail (Roche)

Cell fractionation wash buffer

0.1% NP-40

86 10 mM Tris pH 7.4

10 mM NaCl

3 mM MgCl2

1 mM EDTA

1 mM EGTA

1 mM PMSF

Complete protease inhibitor cocktail (Roche)

2x SDS protein sample buffer

125 mM Tris-HCl pH 6.8

2% SDS (w/v)

20% Glycerol (v/v)

0.005% Bromophenol blue (w/v)

1 M DTT stock solution was added just prior to use to a final concentration of 100 mM.

5x SDS protein sample buffer

125 mM Tris-HCl pH 6.8

10% SDS (w/v)

20% Glycerol (v/v)

0.005% Bromophenol blue (w/v)

1 M DTT stock solution was added just prior to use to a final concentration of 100 mM

2.1.5.2 Sodium dodecyl sulphate-polyacrylamide gel electrophoresis (SDS-PAGE)

Resolving gel

375 mM Tris-HCl pH 8.8

0.1% SDS (w/v)

87 8-15% Acrylamide (v/v)

0.05% TEMED (v/v)

0.075% Ammonium persulphate (w/v)

Stacking gel

50 mM Tris-HCl pH 6.8

0.1% SDS (w/v)

3.75% Acrylamide (v/v)

0.15% TEMED (v/v)

0.075% Ammonium persulphate (w/v)

Running buffer

25 mM Tris-HCl pH 8.3

192 mM Glycine

0.1% SDS (w/v)

Electrophoresis markers

Rainbow™ coloured protein molecular weight markers (Amersham).

High molecular weight, sizes in kDa: 220, 97.4, 66, 46, 30, 21.5, 14.3.

Benchmark™ pre-stained protein ladder (Invitrogen); consists of 10 pre-stained protein bands in the range of 10-200 kDa (sizes vary with batch; for exact sizes see the relevant results sections).

88 2.1.5.3 Immunoblotting (Western blotting) solutions

Blotting buffer

39 mM Glycine

48 mM Trizma® base (Tris(Hydroxymethyl)aminomethane)

0.037% SDS (electrophoresis grade) (w/v)

20% Methanol (v/v)

TBS-Tween (Tris-buffered saline; Tween-20)

TBS pH 7.6

0.1% Tween-20 (v/v)

Blocking buffer

TBS-Tween

5% Dried skimmed milk (w/v)

Antibody incubation buffer

TBS-Tween

3% Dried skimmed milk (w/v)

2.1.5.4 Protein staining

Coomassie Blue

10% Glacial acetic acid (v/v)

45% Methanol (v/v)

0.05% Coomassie Brilliant Blue R (w/v)

89 Coomassie Blue destain

5% Glacial acetic acid (v/v)

40% Methanol (v/v)

Ponceau S red solution

7% Glacial acetic acid (v/v)

0.2% Ponceau S (w/v)

2.1.5.5 Treatment of protein samples with λ protein phosphatase

10x λ protein phosphatase buffer (supplied by manufacturer, NEB, with λ protein phosphatase)

50 mM Tris-HCl

0.1 mM Na2EDTA

5 mM DTT

0.01% Brij 35

2 mM MnCl2

(pH 7.5 at 25 °C)

2.1.6 Antibodies

2.1.6.1 Primary antibodies

For a list of primary antibodies used in these studies see Table 2.3.

2.1.6.2 Secondary antibodies

Immunoblotting: Sheep anti-mouse Ig and donkey anti-rabbit Ig conjugated to horseradish peroxidase (HRP; Amersham); working concentration 1:5000.

90 Immunofluorescence: Goat anti-mouse and goat anti-rabbit Igs coupled to Alexa Fluor

350, Alexa Fluor 488 or Alexa Fluor 546 (Invitrogen); working concentration 1:400.

Immunohistochemistry: Goat biotinylated anti-rabbit IgG (Vector laboratories); working concentration 1:1000.

Table 2.3 Primary antibodies

Antibody Supplier Monoclonal/Polyclonal, Species Working Immunogen dilution Monoclonal, KLH-conjugated 9B11 synthetic peptide 1:1000(IF) New England (myc-tag) (EQKLISEEDL) corresponding Mouse 1:5000 (IB) Biolabs to residues 410-419 of human c- 1.5µg (IP) Myc c-jun (H-79) Santa Cruz Polyclonal Rabbit 1:1000 (IB) Polyclonal, Full length GST from GST Sigma Rabbit 1:5000 (IB) Schistosoma japonicum Polyclonal, synthetic peptide New England PAK1 derived from the N-terminal Rabbit 1:1000 (IB) Biolabs sequence of human Pak1 Polyclonal, KLH-conjugated synthetic peptide Rho A,B,C Upstate (SYPDTDVILMCFSIDSPDSLE Rabbit 1:1000 (IB) N-KK) corresponding to amino acids 73-94 of RhoA,B,C Monoclonal, clone 23A8, 1:2000 (IB) Rac Upstate recombinant protein containing Mouse 1:200 (IF) the full length human Rac Monoclonal, fusion protein Cdc42 Upstate corresponding to residues 1-191 Mouse 1:1000 (IB) of full-length human Cdc42 BD Monoclonal, generated from Rab5 Mouse 1:300 (IF) Biosciences human Rab5 Monoclonal, clone DM1A, DM1A 1:10,000 (IB) Sigma microtubules from chicken Mouse (α-tubulin) 1:1000 (IF) embryo Phallotoxin isolated from Amanita AlexaFluor phalloides mushroom and labelled 5-10 Units/ml 568-Phalloidin Invitrogen N/A with red-orange-fluorescent Alexa (IF) (F-actin) Fluor 568 conjugate Polyclonal, peptide corresponding presenilin 1 D. Asuni to residues 1-25 of human Rabbit 1:1000 (IB) (PS1) (IOP) presenilin 1 Monoclonal, recognises MPM-2 Upstate phosphorylated serine/threonine- Mouse 1:1000 (IB) proline residues

Abbreviations: IOP= Institute of Psychiatry; IF= immunofluorescence; IB= immunoblotting; IP= immunoprecipitation.

91 2.1.7 Mammalian cell culture

2.1.7.1 Chinese Hamster Ovary (CHO) cell culture

CHO cell medium

Nutrient Mixture Ham F-12 (HAM) with glutamine

10% Fetal bovine serum (FBS) (v/v)

100 U/ml Penicillin

100 µg/ml Streptomycin

1x Trypsin-EDTA solution

0.05% Trypsin (w/v)

0.53 mM EDTA

2.1.7.2 Primary neuronal culture

Supplemented Neurobasal medium

Neurobasal medium

2% B27 supplement (v/v)

100 U/ml Penicillin

100 µg/ml Streptomycin

2 mM L-Glutamine

Poly-D-lysine (PDL)

20 µg/ml poly-D-lysine (PDL) in ddH2O

Trypsin-EDTA (TE) solution

0.05% Trypsin

0.53 mM EDTA

92 HBSS without Ca2+ and Mg2+

Triturating solution

10 mg/ml Albumax

500 µg/ml Trypsin inhibitor

10 µg/ml DNAase 1

2.1.8 In vitro protein kinase assays

Lysis buffer

20 mM HEPES pH 7.4

2 mM EGTA

1% Triton X-100

10% glycerol

1 mM DTT

1 mM Na orthovanadate

50 mM β-glycerophosphate

1 mM PMSF

Complete protease inhibitor (Roche)

Wash buffer I

100 mM Tris-HCl pH 7.6

500 mM Lithium chloride

0.1% Triton X-100

1 mM DTT

93 Wash buffer II

25 mM HEPES pH 7.5

0.2% Triton X-100

1 mM EDTA

Kinase assay buffer

25 mM HEPES pH 7.5

20 mM MgCl2

20 mM β-glycerophosphate

20 mM p-nitrophenylphosphate

0.1 mM Na orthovanadate

2 mM DTT

2.1.9 Immunohistochemistry

TBSA (Tris-buffered saline; Sodium azide)

TBS pH 7.6

0.05% Sodium azide

TBS-Triton (Tris-buffered saline; Triton X-100)

TBS pH 7.6

0.3% Triton X-100 (v/v)

94 2.2 Methods

2.2.1 General molecular biology methods

2.2.1.1 Quantitation of nucleic acids

Spectrophotometric quantitation of nucleic acids was performed using an Ultrospec

3000 spectrophotometer (Pharmacia Biotech). The absorbency of samples at 260 nm and 280 nm was recorded. An optical density reading (at a wavelength of 260 nm;

OD260) of 1 corresponds to a concentration of ~50 µg/ml for double stranded DNA and

40 µg/ml for RNA. The OD260/280 ratio gives an indication of purity of the sample. Pure

DNA has a value of 1.8 and pure RNA has a value of 2.0 (Sambrook et al., 1989). In the case of low yield DNA, for example gel purified fragments, DNA concentration was also roughly quantified using ethidium bromide fluorescent quantitation. This method relies on the ability of ethidium bromide to bind to the DNA helix via intercalation. UV- induced fluorescence emitted by DNA-linked ethidium bromide is proportional to the total mass of DNA. The amount of DNA in a sample could then be quantified by visual comparison with the UV-induced fluorescence of a known quantity of a DNA standard or series of standards. This technique has a detection limit of as little as 5 ng of DNA.

2.2.1.2 Restriction enzyme digestion of DNA

Plasmid DNA was digested using the appropriate restriction enzyme and corresponding buffer according to the manufacturers’ instructions (Invitrogen and New England

Biolabs). In a typical reaction, 5-10 units of enzyme were used per microgram of DNA

(1 unit is usually the amount of enzyme required to cleave 1 µg of DNA in 1 hour at

37°C in the appropriate buffer). The volume of enzyme never exceeded 10% of final reaction volume and the incubation time with the enzyme was typically 1-2 hours at

37°C, depending on the quantity of DNA being digested.

95 2.2.1.3 Alkaline phosphatase treatment

Calf intestinal alkaline phosphatase (CIAP, Pharmacia) was used to remove 5′- phosphate groups from digested, double-stranded DNA fragments to prevent self- ligation. Typically 0.05 units of CIAP (1 µl of a 1-in-10 dilution) were used to dephosphorylate 10 µg of linearised DNA in the buffer supplied (One-Phor-All buffer,

Pharmacia). The reaction mixture was incubated for 30 minutes at 37°C. The phosphatase was then inactivated by incubating the reaction at 85°C for 15 minutes.

2.2.1.4 Agarose gel electrophoresis of DNA

Agarose (Ultra pure, electrophoresis grade) was dissolved in boiling 1x TAE buffer and cast on a gel bed with a suitable comb using a horizontal gel apparatus (Hybaid). On setting, gels were placed in an electrophoresis tank containing 1x TAE buffer to a level just above the gel surface. DNA samples containing DNA loading buffer were loaded into the sample wells and were run at 2-10 Volts/cm. Agarose concentrations of 0.6-

1.5% were used, depending on the size of the DNA fragment of interest. Gels were stained with 10 µg/ml ethidium bromide to visualise DNA. They were then placed on a

3UV™ transilluminator emitting ultra violet light (λ = 302 nm), visualised on a Sony video monitor and videographed on a Sony video graphic printer.

2.2.1.5 Recovery of DNA from agarose gels

Ethidium-illuminated DNA bands were excised from TAE/agarose gels using a clean scalpel blade, placed in Spin-X centrifuge tube filters (Costar), and centrifuged for 5 minutes at 7,000xg to recover the DNA from the gel. 50 µl ddH2O was added to the filter and the columns centrifuged again to recover residual DNA.

96 2.2.1.6 Ethanol precipitation of double-stranded DNA

DNA was precipitated from aqueous solution by the addition of NaCl to a final concentration of 150 mM. Two volumes of ethanol (100%) were added to the mixture and the solution was chilled on ice for 30 minutes. DNA was pelleted by centrifugation at 14,000xg for 15 minutes. The DNA pellet was then washed with 70% ethanol to remove excess salt. After air-drying for 5 minutes the pellet was resuspended in ddH2O.

2.2.1.7 Purification of nucleic acids

Removal of contaminants, such as proteins, from aqueous solutions of nucleic acids was achieved by extraction with phenol:chloroform:isoamyl alcohol (25:24:1 v/v/v) followed by chloroform:isoamyl alcohol (24:1 v/v). An equal volume of phenol:chloroform:isoamyl alcohol was added to the aqueous solution of DNA, and the contents of the tube vortexed for up to one minute to form an emulsion. The emulsion was centrifuged at 14,000xg for 10 minutes and the aqueous phase collected. This procedure was repeated at least one more time, or until no contaminants were visible at the interface of the organic and aqueous phases. An equal volume of chloroform: isoamyl alcohol was then added and the contents of the tube were vortexed and centrifuged as before. The nucleic acids were recovered by precipitation with ethanol as described in Section 2.2.1.6.

2.2.1.8 Ligation of prepared vectors and DNA fragments

Linearised, dephosphorylated (if necessary) vector DNA and purified insert DNA concentrations were determined by gel electrophoresis alongside standards of known concentration. Ligations were carried out as previously described (Sambrook et al.,

1989). The ratio of picomole ends for vector:insert was 1:3 and the total quantity of

DNA usually used was approximately 80 ng. The number of picomole ends was

97 calculated using the following formula:

6 (2 x 10 )/ (660 x number of base pairs) = picomole ends/µg double stranded DNA

A typical reaction contained the vector and insert DNA plus 1 unit of T4 DNA ligase in the supplied buffer and 1 mM ATP in a final reaction volume of 10 µl. Ligations were performed overnight at 12-16°C. Appropriate control ligations lacking either vector or insert DNA were carried out simultaneously.

2.2.1.9 Preparation of electrocompetent bacteria

The E. coli strain DH5α (Hanahan, 1983) was used as a host for plasmid amplification, unless otherwise specified in the text. Preparation of electrocompetent DH5α was as follows. A single bacterial colony, picked from a streak-out onto an LB-agar plate, was used to inoculate 5 ml of LB-broth and grown overnight at 37°C in a shaking incubator.

2 ml of this culture was then diluted 100 fold in fresh LB-broth and grown at 37°C with shaking until an OD600 of 0.4 was reached (approximately 2 hours); at this point the bacteria are in mid-log phase growth. The culture was chilled on ice for 10-30 minutes to halt growth and the bacteria were pelleted at 6,000xg for 15 minutes at 4°C in a

Beckman Avanti J25 refrigerated centrifuge. The bacterial pellet was washed with 40 ml ice-cold ddH2O and pelleted by centrifugation, as before. This was repeated to ensure that all traces of salts from the LB-broth were removed. The pellet was then washed in ice-cold, sterile 10% (v/v) glycerol solution. The final bacterial pellet was resuspended in ice-cold 10% (v/v) glycerol (0.3 ml 10% glycerol per 100 ml original culture) which acts as a cryoprotectant, and was then divided into 40 µl aliquots which were frozen on dry ice and stored at −70°C.

98 The transformation efficiency of electrocompetent cells was calculated by transforming an aliquot with 0.1 ng of control plasmid (pUC19, Invitrogen) and counting the resulting number of ampicillin resistant colonies. The calculated efficiency of the

7 electrocompetent bacteria was routinely 5-7 x 10 colonies/µg of plasmid DNA.

2.2.1.10 Electroporation of DH5α cells

Prior to electroporation, salt was removed from ligation reactions by microdialysis for

20 minutes on cellulose membrane filters (0.025 µm pore size, Millipore-UK Ltd). The dialysed ligation mixture was added to an aliquot of ice-thawed electrocompetent DH5α cells and mixed gently. This was quickly added to a dry, chilled, 0.2 cm electroporation cuvette (Bio-Rad) and electroporated (field strength 2.5 kV/cm, capacitance 25 µF, resistance 200 Ohms) using a Bio-Rad Gene Pulser according to the manufacturer’s instructions. Following electroporation, 1 ml of 37°C LB-broth was added to the cuvette. The solution was then transferred to a 1.5 ml Eppendorf tube and incubated at

37°C for 45 minutes to allow the antibiotic resistance gene on the plasmid construct to be sufficiently expressed. When colour selection was possible, plates were spread with

40 µl of X-gal (20 mg/ml prepared in DMSO) and 20 µl of 100 mM IPTG (prepared in ddH2O) and left for 30 minutes prior to plating. The culture was then spread onto LB- amp agar plates (200µl per plate) using a moulded, sterile Pasteur pipette, and left to allow absorption to occur. Plates were then inverted and incubated at 37°C for 14-16 hours.

2.2.1.11 Screening recombinant clones

Following overnight incubation, bacterial colonies present on the LB-amp plates were picked (when colour selection was used, only white colonies were picked) with sterile

99 inoculation loops and each was used to inoculate 5 ml LB-broth containing the appropriate antibiotic. These samples were placed overnight in a shaking incubator at

37°C. Plasmid DNA from these samples was then prepared by the method of alkaline lysis (see Section 2.1.2.4) which exploits differences in properties between plasmid and bacterial genomic DNA (Sambrook et al., 1989). Briefly, 1.5 ml of culture was pelleted in a Biofuge at 14,000xg for 2 minutes and the supernatant discarded. The pellet was resuspended in 100 µl ice-cold Solution 1 by vortexing. 200 µl freshly prepared

Solution 2 was added and mixed by inverting the tube several times, before placing the tubes on ice for no longer than 5 minutes. 150 µl Solution 3 was then added and the contents of the tube mixed by inversion. The resulting mixture was left to precipitate on ice for 5 minutes before centrifugation at 14,000xg for 10 minutes; this stage efficiently removes the precipitated genomic DNA. The supernatant, now containing plasmid DNA and bacterial RNA, was placed in a fresh tube and extracted once with phenol:chloroform:isoamyl alcohol (25:24:1 v/v/v) and once with chloroform:isoamyl alcohol (24:1 v/v). The plasmid DNA was precipitated by adding 2 volumes of 100% ethanol and leaving at room temperature for 5 minutes and the precipitate was pelleted by centrifugation at 14,000xg for 10 minutes. Pellets were washed in 70% ethanol, air- dried and then resuspended in 50 µl TE, pH 8.0 containing 20 µg/ml RNase A. An aliquot of the plasmid DNA was removed and used for restriction enzyme digestion to screen for the desired construct.

2.2.1.12 Large scale preparation of plasmid DNA

The QIAfilter Plasmid Maxi Kit (Qiagen Ltd.) was used according to manufacturer’s instructions. This kit is based on the alkaline lysis method of DNA recovery (Sambrook et al., 1989); typically 200 ml of culture was used, resulting in approximately 500 µg of pure plasmid DNA.

100 2.2.1.13 DNA sequencing

DNA was sent to MWG Biotech UK Ltd., Milton Keynes, for sequence analysis.

2.2.1.14 Polymerase chain reaction (PCR)

Two different thermostable enzymes were used for PCR. The first, Thermophilus aquaticus (Taq) DNA polymerase (Boehringer Mannheim) was used to carry out preliminary PCRs to determine optimal cycle number and annealing temperature for each primer. Taq DNA polymerase has high processivity although it does not have proofreading activity, so it is not ideal for cloning. For cloning PCRs, native Pfu DNA polymerase (Stratagene), isolated from the hyperthermophilic marine archaebacterium

Pyrococcus furiosus, was used. Pfu DNA polymerase possesses both 5' to 3' DNA polymerase and 3' to 5' exonuclease activities. The 3' to 5' proof-reading activity associated with this enzyme results in a 12-fold increase in fidelity of DNA synthesis over Taq DNA polymerase. When used for PCRs, Pfu DNA polymerase generates blunt-ended products which, once purified, can be cloned directly into blunt cloning sites such as those created by Sma1 in the pBluescript SK+ cloning vector. Reactions were performed in 0.2 ml 8-tube PCR strips (BDH), with the following additions: template DNA (100-500 ng), primers (25 pmol of each), 5 µl of 10x Taq buffer

(Boehringer Mannheim) or native Pfu buffer (Stratagene), 8 µl of 100 mM dNTPs, 2.5 units of Taq or Pfu DNA polymerase and ddH2O to 50 µl. Each reaction was mixed gently by pipetting. Reaction mixes were then placed in a T3 Thermocycler (Biometra) for the PCR reaction, as shown in Table 2.4. The conditions found to be most productive were used in subsequent PCR reactions for that particular template/primer combination. In general, the lowest number of amplification cycles was used to minimise the possibility of errors by the DNA polymerase, whilst at the same time generating enough amplified product for subsequent procedures, such as DNA

101 purification and cloning. A sample of each PCR product was analysed for size, quality and quantity by agarose gel electrophoresis alongside appropriate DNA markers

(Section 2.2.1.4).

Table 2.4 PCR Cycling Parameters

Step Cycles Temperature (°C) Time

1. Pre-incubation 1 95 5 minutes 95 1 minute 2. Melting, Primer annealing and DNA 10-30 40-70 1 minute polymerisation 1 minute (Taq) or 3 72 minutes (Pfu) 3. Extension 1 72 10 minutes

2.2.2 PCR-based site-directed mutagenesis

ExSite™ PCR-based site-directed mutagenesis kit was purchased from Stratagene. The

ExSite™ system allows site-specific mutation in double stranded plasmids, which eliminates the need for subcloning into M13-based bacteriophage vectors and for single- stranded DNA (ssDNA) rescue (Weiner et al., 1994). This system uses increased template concentration and reduced cycling number to reduce potential second-site mutations during the PCR and can be used for producing single site point-mutations, large or small deletions and/or 5’-end oligonucleotide-directed base insertions. It was used in the work presented here to delete a large sequence (237 bp) of ALS2 (see

Chapter 4). The mutagenic primers used are detailed in the relevant results sections. The protocol for the ExSite™ mutagenesis kit was followed according to the manufacturer’s instructions and is described in brief.

Control and experimental reactions were set up as shown below and subjected to the

PCR steps as shown in Table 2.5. Control reactions were carried out to monitor

102 mutation efficiency. The pWhitescript vector is the pBluescript vector with a stop codon

(TAA) within the β-galactosidase gene. Thus, even when plated onto LB-amp plates spread with IPTG and X-gal, bacteria transformed with pWhitescript appear white.

Annealing the oligonucleotide control primers to pWhitescript and following the

ExSite™ protocol converts the stop codon back to the Glu encoding codon seen in pBluescript (CAA), allowing functional β-galactosidase expression. Bacteria containing successfully mutated (β-gal+) plasmids then appear blue on LB-amp plates spread with

IPTG and X-gal.

Control reaction:

1 µl (1.5 µg; 0.5 pmol) of pWhitescript 4.5 kb control template DNA

2.5 µl of 10x mutagenesis buffer (Section 2.1.3.2)

1 µl of 25 mM dNTP mix

2 µl (15 pmol; 150 ng) of oligonucleotide control primer #1 (24 mer, phosphorylated)

2 µl (15 pmol; 150 ng) of oligonucleotide control primer #2 (30 mer)

15.5 µl ddH2O to final volume of 24 µl

1 µl (5 U/µl) ExSite DNA polymerase blend

Experimental reaction:

0.5 pmol of template DNA (0.5 pmol of template DNA = 0.33 µl/kb x size of the template, kb)

2.5 µl of 10x mutagenesis buffer

1 µl of 25 mM dNTP mix

15 pmol of each 5’-phosphorylated primer (15 pmol = 5 ng/base x size of primer, bases) ddH2O to a final volume of 24 µl

1 µl (5 U/µl) ExSite DNA polymerase blend

103 Table 2.5 ExSite™ Mutagenesis Cycling Parameters

Step Cycles Temperature (°C) Time 94 4 minutes 1 1 50 2 minutes 72 2 minutes/kb 94 1 minute 2 8 56 2 minutes 72 1 minute/kb 3 1 72 5 minutes

2.2.2.1 Digesting and polishing the PCR product

Following PCR, the reactions were placed on ice. 1 µl (10 U/µl) of the Dpn1 restriction enzyme and 0.5 µl (2.5 U/µl) of cloned Pfu DNA polymerase (both supplied with the kit) were added and reactions were incubated at 37°C for 30 minutes, followed by 72°C for an additional 30 minutes. The Dpn1 endonuclease (target sequence 5’-Gm6ATC-3’) is specific for methylated and hemimethylated DNA and is used to digest parental DNA and to select for mutation-containing amplified DNA (Nelson and McClelland, 1992).

Cloned Pfu DNA polymerase is used prior to end-to-end ligation of the linear template to remove any bases extended onto the 3’ ends of the product by ExSite DNA polymerase blend.

2.2.2.2 Ligating the PCR product

Following analysis of a small amount of the reaction samples by agarose gel electrophoresis (section 2.2.1.4), the following was added to each sample:

100 µl of ddH2O

10 µl of 10x mutagenesis buffer

5 µl of 10 mM ATP

104 10 µl of this mixture was transferred to a fresh microcentrifuge tube and incubated with

1 µl (4 U/µl) of T4 DNA ligase for 1 hour at 37°C.

2.2.2.3 Transformation into XL1-Blue supercompetent cells

2 µl of the ligation reaction was added to 80 µl thawed XL1-Blue supercompetent cells in a prechilled Falcon® 2059 polypropylene tube and incubated on ice for 30 minutes.

The cells were then subjected to a heat pulse for 45 seconds at 42°C, and then placed on ice for a further 2 minutes. Immediately, the entire volume of the cells was plated onto

LB-amp plates. The transformed control cells were plated on an LB-amp agar plate pre- spread with 40 µl of X-gal (20 mg/ml prepared in DMSO) and 20 µl of IPTG (100 mM prepared in ddH2O) for colour selection. As a further control, 0.1 ng of pUC18 transformation control plasmid was added to an aliquot of cells and incubated on ice for

30 minutes, subjected to a heat pulse for 45 seconds at 42°C, placed on ice for a further

2 minutes (as above), then 5 µl of the cells was added to 80 µl of LB broth and this was plated on LB-amp agar plates pre-spread with X-gal and IPTG as described above. The plates were incubated overnight at 37°C. The mutagenized pWhitescript control colonies should appear as blue colonies, and the expected number of colonies for the experimental reaction should be greater than 100 colonies. The transformation efficiency for the pUC18 control plasmid should be greater than 250 colonies, with more than 98% displaying the blue phenotype.

2.2.2.4 Screening for mutated plasmids

Plasmid DNA was isolated as in Section 2.2.1.11 and then screened for the desired mutation using analytical restriction enzyme digests. Clones appearing to be positive for the mutation were then sequenced to ensure that no undesirable changes were introduced during the mutagenesis protocol.

105 2.2.3 Purification of GST fusion proteins

The glutathione S-transferase (GST) gene fusion system is an integrated system for the expression, purification and detection of fusion proteins produced in E. coli. The use of the BL21 strain of E. coli is generally recommended for protein production, as this strain lacks both the lon protease and the ompT outer membrane protease, which can degrade proteins during purification (Grodberg and Dunn, 1988). The system relies on the cloning of the cDNA of interest into a pGEX plasmid vector (Pharmacia). pGEX vectors are designed for inducible, high level intracellular expression of genes/gene fragments as fusions with Schistosoma japonica GST. Fusion proteins are purified from bacterial lysates by using glutathione sepharose 4B according to manufacturer's instructions (Pharmacia). Purified proteins can then be eluted from the glutathione sepharose 4B beads. The steps involved in the purification of GST fusion proteins from

E. coli are described in detail in the manufacturer's instructions (Pharmacia). Conditions were optimised as described below:

1. Preparation of stock glutathione sepharose 4B

Glutathione sepharose 4B beads are supplied as a 75% slurry in 20% ethanol. The beads were washed three times with ice-cold PBS to remove the ethanol and re-suspended in

PBS to produce a 50% slurry.

2. Preparation of bacterial sonicates

A single colony of E. coli harbouring a recombinant pGEX plasmid was used to inoculate 5 ml of LB-amp broth. This was incubated with shaking at 37°C for 12-15 hours. The resulting culture was diluted 1:100 into fresh, pre-warmed LB-amp broth, and grown at 37°C with shaking until the OD600 reached approximately 0.6-0.8. At this stage, 100 mM IPTG was added to a final concentration of 0.4 mM, to induce protein

106 expression, and incubation was continued for a further 3 hours. The culture was then centrifuged at 6,000xg for 15 minutes at 4°C, to sediment the cells, and the supernatant was discarded. The cell pellet was re-suspended using 50 µl ice-cold PBS with complete protease inhibitor cocktail (Roche) per 1 ml of culture. Alternatively, for recovery of insoluble proteins, the pellet was resuspended in 50 µl of 6 M urea per 1 ml of culture; see step 3 below. Suspended cells were disrupted by sonication performed on ice in 30 second bursts, 4-6 times. The resulting sonicate was centrifuged at 15,000xg for 20 minutes at 4°C, and the supernatant was used in step 4, below.

3. Dialysis of salts

As described above, proteins that were insoluble in PBS were solubilised in 6 M urea with complete protease inhibitor cocktail and rocked gently at 4°C for 20 minutes. This was centrifuged at 15,000xg for 15 minutes, and the resulting supernatant was dialysed against PBS (2 l PBS per 10 ml supernatant) overnight at 4°C. Centrifugation was repeated as above, to remove any protein that had come out of solution during dialysis, and the resulting supernatant was used in subsequent steps.

4. Batch purification of fusion proteins using glutathione sepharose 4B

2 ml of the 50% glutathione sepharose 4B slurry was added to 10 ml supernatant (i.e. 1 ml bed volume of beads per 200 ml bacterial grow). The mixture was incubated with gentle agitation at room temperature for 1 hour, or 4°C for 2 hours, after which the matrix with adsorbed fusion protein was sedimented by centrifugation at 500xg for 5 minutes. The supernatant was discarded and the glutathione sepharose 4B pellet washed three times with PBS. At this stage the bound fusion protein could be quantitated by direct comparison with standard amounts of bovine serum albumin (BSA) on

Coomassie Blue stained SDS-PAGE gels. Purified bound fusion protein samples were

107 stored at 4°C (in PBS as a 50% slurry) for future use or were eluted as described in step

5.

5. Batch elution

500 µl of reduced glutathione elution buffer was added to the sedimented beads (per 1 ml bed volume of beads). The re-suspended matrix was incubated with shaking at room temperature for 10 minutes to elute the fusion protein. The matrix was re-sedimented by centrifugation, and the supernatant stored on ice. Elution and centrifugation steps were repeated a further two times before pooling the three eluates, resulting in purified GST- fusion protein dissolved in 1.5 ml glutathione elution buffer. Samples were taken at various stages throughout the above procedures, boiled in 2x SDS protein sample buffer and stored at −20°C for future analyses.

2.2.4 Protein analysis

2.2.4.1 Protein concentration determination

Determination of protein concentration was carried out using the Bio-Rad protein assay according to the manufacturer’s instructions (Bio-Rad, UK). This is based on the

Bradford dye-binding procedure and is a simple colorimetric assay for measuring total protein concentration (Bradford, 1976). The assay is based on the observation that the maximum absorbance for an acidic solution of Coomassie Brilliant Blue G-250 shifts from 465 nm to 595 nm when binding to protein occurs. Absorbance readings were recorded at a wavelength of 595 nm using an Ultrospec 3000 spectrophotometer

(Pharmacia Biotech) and values were compared to a curve drawn from a set of BSA standards, freshly prepared each time the assay was performed.

108 2.2.4.2 SDS-PAGE of protein samples

For SDS-PAGE analyses of cell lysates, cells were washed in PBS, lysed in lysis buffer

(Section 2.1.5.1) for 20 minutes on ice, then boiled for 5 minutes in appropriate amounts of 2x or 5x SDS protein sample buffer (final concentration 1% SDS). The protein samples were then separated by SDS-PAGE using the Mini Protean II gel electrophoresis system (Bio-Rad) with a discontinuous buffer system. 8-12% acrylamide gels were run at a constant voltage of 120-180 V, until the dye front reached the bottom of the gel. Protein bands could be visualised at this stage by staining in

Coomassie Blue for 1 hour and then washing for 3 successive periods of 20 minutes in

Coomassie Blue destain.

2.2.4.3 Immunoblotting (Western blotting)

Transfer of proteins to nitrocellulose

After SDS-PAGE, proteins were transferred from gels to a 0.45 µm nitrocellulose support (Schleicher and Schuell) using the following filter sandwich:

Cathode

5 filters (5 x 8 cm, grade 1, Whatman) soaked in Blotting buffer

SDS-polyacrylamide gel

Nitrocellulose membrane pre-soaked in Blotting buffer (5 x 8 cm)

5 filters (5 x 8cm, grade 1, Whatman) soaked in Blotting buffer

Anode

This sandwich was assembled on a semi-dry transfer unit (Transblotter SD, Bio-Rad) at

17 V for 45-60 minutes. Nitrocellulose membranes with protein transferred onto them were referred to as "Blots". Blots were incubated with Ponceau S for 1-2 minutes to

109 determine the efficiency of protein transfer. Blots were then washed for 5 minutes with

TBS-Tween. This method of protein detection is compatible with all subsequent methods of immunological probing.

Antibody probing of membrane-bound proteins

Blots were incubated in blocking buffer (see Section 2.1.5.3) for up to 1 hour at 37°C to reduce non-specific binding. All incubations with antibodies were performed in antibody incubation buffer. Blots were incubated with an appropriate dilution of primary antibody overnight at 4°C (Table 2.3 gives the details of all primary antibodies used). The blots were then washed for 10 minutes in TBS-Tween; this was repeated a further four times, after which an appropriate dilution of secondary antibody was added for 1 hour at room temperature. Secondary antibodies used depended on the species in which the primary antibody was raised, for details see Section 2.1.6.2. Following a further four 15 minute washes in TBS-Tween, immunoreactive species were immediately visualised using enhanced chemiluminescence development reagents

(ECL; Amersham) according to the manufacturer's instructions, in conjunction with

Hyperfilm-ECL (Amersham). Film was developed using a Hyperprocesser (Amersham).

2.2.4.4 Treatment of samples with λ protein phosphatase

Immunoprecipitation of ALS2 from cultured cortical neurons was carried out with 500

µg protein as described in Section 2.2.8.1. The pelleted beads were washed 3 times with

1 ml lysis buffer without protease inhibitor, followed by 3 washes in ice-cold PBS. The beads were then resuspended in 20 µl of λ protein phosphatase buffer (Section 2.1.5.5) containing 1 µl (400 units) of λ protein phosphatase. The mixture was incubated at 30°C for 1 hour and 20 µl of 2x SDS protein sample buffer was then added before boiling for

5 minutes. An IP sample without λ protein phosphatase was also prepared as above, as a

110 negative control. The protein samples were then analysed by SDS-PAGE and immunoblotting (Sections 2.2.4.2-2.2.4.3).

2.2.5 Preparation of rabbit polyclonal antibodies

Rabbit polyclonal antibodies were prepared by immunising rabbits with GST fusion proteins. GST fusion protein production and immunisations were performed "in house".

Fusion proteins were purified and eluted as described in Section 2.2.3. Following a pre- immune test bleed, the rabbit was immunised with 100 µg of antigen (in 500 µl of glutathione elution buffer) emulsified with 500 µl of TiterMax Classic Adjuvant

(Sigma) according to the manufacturer's instructions. The rabbit was immunised with 4 x 0.2 ml antigen/adjuvant subcutaneously, at 0, 2, 4 and 6 weeks. At 8 weeks the rabbit underwent a termination bleed. Bleeds were obtained by trained staff, collected in glass bottles and left at room temperature for 1 hour to allow clotting. The clots were rimmed with a glass rod and allowed to retract overnight at 4°C. The serum was carefully removed, centrifuged at 2,000xg for 10 minutes at 4°C and then stored in aliquots at -

20°C. Once defrosted, sodium azide was added to 0.05% and aliquots were stored at

4°C for up to 6 months.

2.2.6 Mammalian cell culture and transfection

2.2.6.1 CHO cell culture

CHO cells were maintained in monolayer culture in Nutrient Mixture Ham F-12 (HAM) with glutamine, supplemented with fetal bovine serum, penicillin and streptomycin (see

Section 2.1.7.1) at 37°C under an atmosphere of 5% CO2. Cells were passaged when they formed an approximately 80% confluent layer in the culture vessel. Cell medium was aspirated and the cells washed twice with PBS. After removal of the second wash, enough trypsin-EDTA solution was added to cover the monolayer (e.g. 2 ml trypsin-

111 EDTA solution for a 175 cm2 flask). The vessel was rocked to evenly distribute the trypsin-EDTA solution before placing it in a 37°C incubator until the cells were just beginning to detach (1-2 minutes). The culture vessel was struck to dislodge remaining adherent cells. Cell medium (e.g. 8 ml for a 175 cm2 flask) was then added to inactivate the trypsin, and the cells were pipetted up and down to triturate the cells. Cell density was determined by counting a sample of the cell suspension using a haemocytometer under low power magnification before distribution of an appropriate volume into vessels for sub-culturing.

2.2.6.2 Primary embryonic rat cortical and hippocampal neuron culture

Cortical and hippocampal neurons were obtained from embryonic day 18 (E18) Sprague

Dawley rat embryos (Charles Rivers) and cultured on poly-D-lysine (PDL)-coated glass coverslips, or PDL-coated cell culture dishes in supplemented Neurobasal medium

(Section 2.1.7.2).

Dissection and preparation of neurons

A time-mated Sprague Dawley dam was sacrificed by cervical dislocation. The abdominal wall was cut through and the two horns of the uterus removed. The foetuses

(typically 10-12 per dam) were removed and dissected in HBSS without Ca2+/Mg2+. The brains were removed by cutting the skulls open with fine sprung scissors and gently scooping out using a spatula. The cortex and hippocampus were dissected under a dissecting microscope: The midbrain and cerebellum were removed using Dumont forceps and the meninges were removed to prevent contamination of cultures with fibroblasts; the cortex and hippocampus were then dissected and processed separately.

The isolated tissue was incubated with TE/HBSS solution for 20 minutes at 37°C, after which DNAse 1 solution (0.001% DNAse in HBSS) was added. The solution was

112 gently mixed; the tissue was then re-suspended in triturating solution, and triturated using glass Pasteur pipettes of decreasing bore size to produce a single-cell suspension, which was then diluted with supplemented Neurobasal medium. Cell density was determined using a haemocytometer. Various plating densities were used and cells were transfected and/or harvested at 1.5– 21 days in vitro (DIV).

2.2.6.3 Transient transfection

Lipid based transfection of CHO cell cultures

Lipofectamine™ reagent (Invitrogen) is a 3:1 (w/w) liposome formulation of the polycationic lipid 2,3-dioleyloxy-N-[2(spermine-carboxamido)ethyl]-N,N-dimethyl-1- propanaminiumtrifluoroacetate (DOPSA) and the neutral lipid dioleoyl phosphatidylethanolamine (DOPE) in membrane-filtered H2O. The positively charged and neutral lipids form liposomes that can complex with nucleic acids. When applied to cultured cells, the lipid-nucleic acid complex facilitates the uptake of nucleic acids into the cells. CHO cells were transfected essentially according to the manufacturer's instructions, using amounts of reagents as indicated in Table 2.6. The cells were plated out the day before the transfection experiment so that they reached 50-80% confluency on the day of transfection (half of this density was used for immunofluorescence), and were harvested for analysis 24 hrs post-transfection.

Table 2.6 Transient transfection of CHO cells with Lipofectamine™ Reagent

Diameter of dish Number of cells Amount of plasmid Amount of (cm) plated DNA (µg) Lipofectamine (µl) 3.5 2 x 105 1 4 6 6 x 105 3- 4 12 10 1.6 x 106 8- 10 32

113 Lipid-based transfection of primary cortical neuron cultures

Cortical neurons were transfected using Lipofectamine 2000™ reagent (Invitrogen) according to the manufacturer's instructions. Cells were cultured in 12-well plates (3.5 cm diameter) on PDL-coated glass coverslips, at a density of 3 x 105 cells per well in 1 ml of growth medium. 25 µl Optimem® reduced serum medium (Invitrogen) was mixed with 1 µl Lipofectamine 2000 and this was incubated for 5 minutes at room temperature. 1-4 µg DNA was mixed with 25 µl Optimem and this was added to the

Optimem/Lipofectamine 2000 and mixed thoroughly. After incubation at room temperature for a further 25 minutes, the mixture was added dropwise onto the cell culture medium, this was swirled gently and the cells were then returned to incubation at 37°C. Cells were harvested 4-24 hours later.

2.2.7 GTPase activation assays

Cellular Rho, Rac and Cdc42 activities were assayed using commercially available kits essentially according to the manufacturer’s instructions (Upstate). Active (GTP-bound)

Rho, Rac and Cdc42 were captured on GST-bait coupled to glutathione agarose beads.

Active Rho was captured using GST-Rhotekin Rho-binding domain (GST-RBD) and active Rac and Cdc42 captured using GST-PAK1 p21-binding domain (GST-PBD)

(Benard et al., 1999; Ren and Schwartz, 2000; Taylor and Shalloway, 1996).

CHO cells were cultured in 10 cm dishes and transfected using Lipofectamine™ reagent. 24 hours post-transfection and following 16 hours serum starvation, cells were washed twice with ice-cold PBS, harvested into 500 µl ice-cold magnesium lysis buffer

(MLB; Section 2.1.5.1), and cleared by centrifugation at 14,000xg for 5 minutes. 10 µl of each lysate sample was retained for analysis by SDS-PAGE and immunoblotting, and the remainder of each lysate was immediately incubated with 5-10 µg of either GST-

114 RBD (for Rho assay), GST-PBD (for Rac and Cdc42 assays) or GST (as a negative control) coupled to agarose beads for 45 minutes at 4°C. The beads were collected by centrifugation at 14,000xg for 5 minutes, the supernatant was discarded and the beads were resuspended in 20 µl of 2x SDS protein sample buffer. The samples were boiled at

95°C for 5 minutes and subjected to SDS-PAGE and immunoblotting.

Positive and negative controls for the assay were set up as follows: lysates of CHO cells co-transfected with either Rho, Rac or Cdc42 and pCIneo-CAT were cleared as above and EDTA was added to a final concentration of 1 mM. Each sample was divided into two tubes (one positive and one negative control): for the positive control, GTPγS was added to a final concentration of 100 µM and for the negative control GDP was added to a final concentration of 1 mM. The tubes were incubated at 30°C for 15 minutes with agitation. The reaction was terminated by placing tubes on ice and adding MgCl2 to a final concentration of 60 mM. Each sample was then incubated for a further 30 minutes with 5-10 µg of either GST-RBD (for Rho assay) or GST-PBD (for Rac and Cdc42 assays) at 4°C. The beads were collected by centrifugation at 14,000xg for 5 minutes, the supernatant was removed and the beads were resuspended in 20 µl of 2x SDS protein sample buffer. The samples were boiled at 95°C for 5 minutes and subjected to

SDS-PAGE and immunoblotting.

2.2.8 Immunoprecipitation and in vitro protein kinase assays

2.2.8.1 Immunoprecipitation from cell lysates

All steps of immunoprecipitation (IP) experiments were performed at 4°C or on ice.

Cells were washed twice in ice-cold PBS before being incubated in cell lysis buffer (e.g.

1 ml per 10 cm dish; see Section 2.1.5.1) for 20 minutes. The lysates were then scraped and pipetted into 1.5 ml tubes and centrifuged at 15,000xg for 10 minutes. The

115 supernatants were removed to fresh tubes and were then pre-cleared by incubation with

25 µl of Protein A-sepharose in cell lysis buffer (50% v/v) per 1 ml of lysate, for 1 hour at 4°C. Following centrifugation for 5 minutes at 5,000xg, lysates were transferred to fresh tubes and samples taken to check transfection and to quantitate total protein as in

Section 2.2.4.1. Protein concentration was adjusted to 1 µg/µl and equal amounts of protein (typically 500-1000 µg) for each sample were then used for IP. 100-500 µg

samples were placed into each of two fresh tubes. One of each duplicate sample was incubated with primary antibody and the other acted as a ‘no antibody’ control. Both samples were placed on a rotor for 2-16 hours after which 30 µl of protein A-sepharose beads (50% v/v in lysis buffer) was added. This was incubated for a further 1-2 hours on the rotor before centrifuging for 5 minutes at 5,000xg. The pelleted beads were retained and washed three times in 1 ml of lysis buffer. The beads were then either resuspended in the appropriate buffer or boiled in 30 µl of 2x sample buffer for 10 minutes and subjected to SDS-PAGE.

2.2.8.2 Immunoprecipitation from rat brain homogenate

Whole adult rat brain was homogenised in 5 ml of tissue homogenisation buffer, using a dounce tissue homogeniser. Homogenates were then passed through a 28-gauge needle three times and centrifuged at 15,000xg for 30 minutes. The supernatant was then precleared by incubation with 50 µl of protein A-sepharose (50% slurry in PBS) per 1 ml of homogenate, for 1 hour. Following centrifugation at 5000xg for 5 minutes, homogenates were transferred to a fresh tube and samples were taken for protein quantitation (as described in section 2.2.4.1). 2 mg of total protein in 1 ml of buffer was used per IP sample, and the IP was then carried out as for CHO cell lysates (see above).

116 2.2.8.3 In vitro kinase assays

Purification of kinases

Kinase-transfected cells were washed twice with ice-cold PBS and harvested by scraping into 1 ml kinase assay lysis buffer at 4°C (Section 2.1.8). The cell lysates were pre-cleared with 50 µl of protein A-sepharose beads (50% v/v in lysis buffer). The total protein was quantitated (section 2.2.4.1), 50 µg added to each of the three tubes and the volume made up to 1 ml with lysis buffer. The protein was then isolated by immunoprecipitation using 1.5 µg 9B11 antibody (see Section 2.2.8.1). The beads were then washed two times with 1 ml lysis buffer, then two times with 1 ml wash buffer I, and finally two times with 1 ml wash buffer II (see section 2.1.8). Negative controls were set up without immunoprecipitating antibody. “Reaction Mix” control was set up without lysate/kinase.

Kinase assays

5 µg Myelin basic protein (MBP) was used as a substrate in the PAK1 kinase assay.

Kinase reactions were set up as follows: 30 µl 20 µM ATP in kinase buffer (made from

10 mM stock of ATP), 0.259 MBq [γ-32P] ATP (0.7 ul of 370 MBq/µl; Amersham-

Pharmacia Biotech), 5 µg substrate and kinase buffer to a final volume of 35 µl. The kinase reaction mixture was then added to the immunoprecipitated kinases (or no kinase control or reaction mix only), still on beads. Samples were incubated for 20 minutes at

30°C and the reactions then terminated by the addition of 20 µl of 5x sample buffer.

The samples were analysed by performing SDS-PAGE, vacuum drying gels on a Slab

Gel Dryer vacuum pump (Savant) and exposing them to Hyperfilm-ECL (Amersham) for 2-72 hours.

117 2.2.9 CHO cell fractionation

CHO cells were cultured in 10 cm dishes and transfected with 10 µg ALS2 cDNA using

Lipofectamine™ reagent. 18 hours post-transfection cells were washed twice with ice- cold PBS and harvested into 1 ml of cell fractionation homogenisation buffer (Section

2.1.5.1). Cells were homogenised using a dounce homogeniser and centrifuged at 375xg for 5 minutes at 4°C. The pellet (nuclear fraction) was washed with cell fractionation wash buffer (Section 2.1.5.1) five times, resuspended in a final volume of 100 µl wash buffer and added to 100 µl 2x SDS protein sample buffer. The supernatant was centrifuged at 100,000xg for 30 minutes at 4°C. The resulting supernatant (cytoplasmic fraction) was added to 80 µl 2x SDS protein sample buffer and the pellet (membrane fraction) was washed with wash buffer 3 times, resuspended in 50 µl wash buffer and

50 µl 2x SDS protein sample buffer. Samples were subjected to SDS-PAGE and immunoblotting. The following amounts were used to obtain equal loading: 25 µl cytoplasmic fraction; 1.25 µl membrane fraction; 1.75 µl nuclear fraction. For tubulin immunoblots 1/10th of the above amounts were loaded on the gel.

2.2.10 Immunohistochemistry

The methods used in immunohistochemical staining are based on avidin-biotin and peroxidase methodologies (Hsu and Raine, 1981). The target protein (ALS2) was detected by a primary ALS2 antibody (described in Chapter 3) that in turn was detected with a biotinylated secondary antibody. The secondary antibody was then bound with

Avidin-Biotinylated enzyme Complex (ABC). Staining was completed by addition of the 3,3'-diaminobenzidine tetrahydrochloride (DAB) chromogen which precipitates at the antigen site, generating a brown stain.

118 Sectioning of frozen rat brain

Rat brain was fixed with 4% paraformaldehyde (PFA)/PBS (pH 7.4) for 48 hours at room temperature. Fixed brain specimens were washed once with PBS and then soaked in 50 ml of 30% sucrose in TBSA solution at 4°C (see Section 2.1.10). When the brain specimens descended to the bottom of the tube they were removed and stored at -70°C. A Leitz 1321 freezing microtome (Leitz, Japan) was used to prepare frozen brain sections. Free-floating sections in TBS were used for immunohistochemistry.

Avidin-Biotinylated enzyme Complex (ABC) staining

40 µm sections were used for staining. Endogenous peroxidase activity was blocked with 1% H2O2 in TBS for 15 minutes after which the sections were rinsed 3 times for 5 minutes in TBS. Non-specific protein binding was blocked with 15% normal goat serum (NGS; Vector Laboratories, UK) in TBS-Triton (see Section 2.1.10) for 40 minutes. Sections were then incubated overnight at 4°C, with constant gentle agitation, in primary antiserum diluted with 10% NGS/TBS-Triton. Primary antibody treated sections were rinsed 3 times for 5 minutes in TBS and incubated for 2 hours in 1:1000 diluted biotinylated secondary antiserum (Vector laboratories, USA) with 10%

NGS/TBS-Triton. Secondary antiserum treated sections were rinsed 3 times for 5 minutes in TBS and incubated for 2 hours in Vectastain® Elite® ABC solution (Vector laboratories, USA) prepared at 1:1000 dilution in TBS, 30 minutes in advance. ABC solution treated sections were rinsed 3 times for 5 minutes in TBS. Sections were then incubated in 0.05% DAB solution (preferably in darkness to reduce background staining) for 5-10 minutes. 0.05% DAB solution was prepared by crushing a 10 mg

DAB tablet (Sigma, UK) into 20 ml TBS with 6 μl of 30% H2O2 and used immediately.

Staining intensity was frequently checked during incubation. The reaction was terminated by transferring sections to ice-cold TBS. Sections were rinsed a further 3

119 times for 5 minutes in TBS. Sections were taken from the TBS, placed onto Superfrost- plus slides (BDH, UK) and air-dried overnight. Sections on slides were then left in xylene for at least 30 minutes, and mounted with a glass cover slip with a xylene-based mountant, e.g. DPX® (BDH, UK). Images were captured on a Zeiss Axioscope 2 MOT using Axiocam and Axiovision software (Zeiss, Germany).

2.2.11 Immunofluorescence

Cells on glass coverslips were washed twice in PBS at room temperature before being fixed in 4% PFA/PBS for 15 minutes, and permeabilised in 0.1% Triton X-100/PBS for

10 minutes. Cells were then incubated with 5% FBS/PBS blocking solution for 30-60 minutes to reduce non-specific binding. Incubation with primary antibodies in blocking solution was then carried out for 1 hour. The cells were washed for 10 minutes in PBS

5-6 times, before being incubated for 1 hour with 5% FBS/PBS containing goat anti- mouse and goat anti-rabbit Igs coupled to Alexa Fluor 350, Alexa Fluor 488 or Alexa

Fluor 546 (Invitrogen). Nucleic acids were stained with 50 µg 4′,6-Diamidino-2- phenylindole (DAPI; 0.5 mg/ml in DMSO) for 5 minutes, followed by a further 5-6 washes (10 minutes each wash) in PBS. Finally, coverslips were mounted onto slides using Vectashield mountant (Vector Laboratories). Conventional images were viewed using a Zeiss Axioscop microscope and captured using a charge coupled device (CCD) camera (Princeton Instruments), and confocal images captured using a Zeiss LSM 510

META confocal microscope at the appropriate excitation wavelengths.

120

CHAPTER 3: PREPARATION OF AN ALS2

ANTIBODY AND LOCALISATION OF ALS2 IN

NEURONAL TISSUES

121 3.1 Introduction

Alsin/ALS2 was first identified in 2001 by two independent research groups (Hadano et al., 2001; Yang et al., 2001) as a protein of unknown function that is mutated in rare autosomal recessive juvenile forms of MND. The identified mutations all cause a predicted loss of function of the ALS2 protein. This research project was started immediately after the publication of these findings, with the purpose of investigating the normal function of ALS2, so as to shed light on the pathological mechanisms involved in MND.

Sequence homology data suggests that ALS2 may function as a guanine nucleotide exchange factor (GEF) for the Ras superfamily of GTPases. ALS2 Long-form contains three putative GEF domains; an amino-terminal domain that displays homology to the

Ran GEF RCC1, a central region containing Dbl and pleckstrin homology domains

(DH/PH) that are found in GEFs for Rho family members, and a carboxyl-terminal vacuolar protein sorting 9 (VPS9) domain which is found in GEFs for Rab5. ALS2

Short-form has been identified at the mRNA level, although there is as yet no evidence of the presence of a short form protein. As this predicted truncated protein is not thought to contain any of the functionally homologous GEF domains an antibody was created specifically to ALS2 Long form (referred to simply as ALS2).

The study of the subcellular localisation of ALS2 could provide clues to its function within the cell. For example a cytoplasmic/intracellular or plasma membrane localisation would indicate function as a Rho or Rab GEF whereas localisation within the nucleus may suggest a function as a GEF for Ran (as Ran is activated in the nucleus)

(Hughes et al., 1998). The first step of this project was therefore the preparation and characterisation of an antibody to ALS2, in order to facilitate such studies.

122 3.2 Methods

3.2.1 PCR and cloning into pGEX expression vector

PCR reactions were carried out in a T3 Thermocycler (Biometra) in order to amplify a sequence of cDNA encoding a 651 bp fragment of ALS2 (corresponding to amino acid residues 452-668 of human ALS2). Pairs of synthetic oligonucleotide primers were designed to introduce an EcoR1 site (GAATTC) at each end of the construct (to facilitate cloning into the EcoR1 site of pGEX-5X-1), as follows:

Primer1: 5’-GCGGAATTCGAACAGGTTAAACAGGAATCAATGC-3’

Primer 2: 5’-GCGGAATTCTCCAAGCTTACTACAGGAGAGAAG -3’

Approximately 100 ng DNA was amplified in a mixture containing 25 pmol of each primer (Oswel DNA service, University of Southampton), 16 µl deoxynucleotide triphosphates (25 mM each), 1 µl Pfu DNA polymerase (2.5 units) and 10 µl 10x Pfu reaction buffer (Stratagene). The final reaction volume was 100 µl. Double-stranded

DNA was denatured at 95°C for 5 minutes, followed by a cycle of melting at 95ºC for 1 minute, primer annealing at 60°C for 1 minute, and finally DNA polymerisation at 72°C for 3 minutes. This was repeated for 12, 18, 24 or 30 cycles, and was followed by a final extension step of 72°C for 10 minutes. The PCR product (obtained from the minimum cycle number) was cloned blunt into the Sma1 site of pBluescript SK+ (Stratagene) and sequenced (MWG Biotech) to ensure no errors had occurred. The PCR product (insert) was then sub-cloned in-frame as an EcoR1 fragment into the pGEX-5X-1 GST vector

(see Section 2.2.1).

3.2.2 Preparation of GST fusion protein

The GST-ALS2452-668 construct was transformed into and expressed in BL21-

CodonPlus®-RIL (BL21-RIL; Stratagene) E. coli. BL21-RIL cells contain extra copies

123 of rarely-used E. coli tRNAs (argU, ileY and leuW) that are frequently found in A/T rich genomes. Empty pGEX-5X-1 was used to express GST alone in BL21-RIL cells as a control. A 1 ml sample was taken from each culture at 0 hrs (at OD600 0.6-0.8) and 1 hr, 2hr and 3hrs post-induction with 0.4 mM IPTG. Samples were resuspended in ice- cold PBS, lysed by sonication and separated by centrifugation (at 15,000xg for 10 minutes) into soluble (supernatant) and insoluble (pellet) fractions. 2x SDS protein sample buffer was added to each sample and the samples were boiled for 5 minutes prior to analysis by SDS-PAGE.

The GST-ALS2452-668 fusion protein was expressed predominantly in the insoluble fraction. The protein was therefore purified from the insoluble fraction by solubilising in 6 M urea followed by dialysing against PBS to a final concentration of approximately

30 mM urea. The protein was then captured using glutathione sepharose 4B beads, and eluted with reduced glutathione elution buffer.

3.2.3 Production of a polyclonal antibody to ALS2

100 µg GST-ALS2452-668 antigen (in glutathione elution buffer) was concentrated to 500

µl using Centricon® centrifugal filter units (Millipore) and was emulsified in 500 µl

TiterMax® classic adjuvant, according to the manufacturer’s instructions (Sigma-

Aldrich). An immunologically-naïve New Zealand white rabbit was injected into 4 subcutaneous sites with 200 µl antigen/adjuvant per site. This was repeated 3 more times at 10-15 day intervals until a terminal bleed was carried out (47 days after the first immunisation).

3.2.4 Affinity purification of ALS2 polyclonal antibody

Antibodies were affinity-purified from sera using immobilised antigen. Adsorbents were

124 prepared by spotting 500 µg antigen onto Immobilon PVDF membrane (Millipore). The protein was allowed to dry before blocking in 5% skimmed milk/PBS for 2 hours. 2 ml of the antiserum was incubated with the membranes and 10 ml 3% skimmed milk/PBS overnight. After incubation, the membranes were washed with PBS, and antibody was eluted in 500 µl Elution Buffer (0.5 M glycine, 0.5 M NaCl, pH 2.4) for 10 minutes.

The elution step was repeated three more times and the pooled eluate was neutralised to pH 8 before addition of sodium azide to 0.05%.

3.2.5 Cell culture and transfection

CHO cells were cultured and transiently-transfected as described in Section 2.2.6. A carboxyl-terminal myc-tagged human ALS2 cDNA was prepared by N.J.O. Jacobsen

(Institute of Psychiatry). Briefly, the 5’ end of ALS2 was amplified by PCR from a human brain cDNA library and ligated to a 5’-truncated partial ALS2 cDNA (clone

KIAA1563, obtained from Kazusa Research Institute, Japan) so as to create a full-length clone. A carboxyl-terminal myc tag was then added by PCR and the tagged full-length cDNA then cloned into pCIneo (Promega) as a Not1 fragment.

3.2.6 CHO cell fractionation

CHO cells were cultured in 10 cm dishes and transfected with 5 µg myc-tagged ALS2 cDNA using Lipofectamine™ (as described in Section 2.2.6.3). Cells were washed twice with ice-cold PBS, and incubated in 1 ml cell fractionation homogenisation buffer

(Section 2.1.5.1) on ice for 10 minutes, followed by homogenisation with a dounce homogeniser. Cell fractionation was performed as described in Section 2.2.9. Samples were boiled for 5 minutes in 2x SDS protein sample buffer, and subjected to SDS-

PAGE and immunoblotting.

125 3.2.7 Antibodies

Immunoblotting, immunohistochemistry and immunofluorescence were carried out as described in Sections 2.2.4.3, 2.2.10 and 2.2.11. The ALS2 antibody described in this chapter was used at varying concentrations as indicated in the relevant results sections.

For details and concentrations used of other antibodies in this chapter see Table 2.3.

Antibody preabsorption was carried out by incubation of ALS2 antibody with 10x molar excess of the purified GST-ALS2452-668 immunogen or with purified GST protein alone (quantified by SDS-PAGE and staining with Coomassie Blue, alongside BSA standards), prior to addition to the immunoblot.

3.3 Results

3.3.1 GST fusion protein preparation

A GST-ALS2 expression plasmid (GST-ALS2452-668) was created in order to purify a fragment of ALS2 to use as an antigen for the production of ALS2 antibodies. This involved isolating a 651 bp fragment of ALS2 corresponding to amino acid residues

452-668 of human ALS2 (Figure 3.1), which was generated by PCR as described above.

A sample of each PCR product was checked for size, quality and quantity by agarose gel electrophoresis, in order to identify the most productive conditions (Figure 3.2).

PCR reactions were initially carried out with Taq DNA polymerase in order to optimise the primer annealing temperature. Reactions were carried out with annealing temperatures of 55°C, 60°C, 65°C, and 70°C and samples from these reactions were analysed by agarose gel electrophoresis. 60°C was found to be the optimum temperature

(Figure 3.2 A). PCR reactions were then carried out with Pfu DNA polymerase at 60°C with different cycle numbers (12, 18, 24 and 30). 18 cycles was used as this was the lowest number of cycles (minimising the possibility of errors by the DNA polymerase)

126 that generated enough amplified product for the subsequent procedures (Data not shown). The PCR product was then purified from the TAE gel (see section 2.2.1.5) and cloned into the Sma1 cloning site of pBluescript SK+. Following colour selection and sequencing (MWG Biotech), a clone with the correct insert was identified, and the insert was then sub-cloned in-frame into pGEX-5X-1 as an EcoR1 fragment. Clones were screened using an analytical digest with EcoR1 and BamH1 and a clone containing the GST-ALS2452-668 insert was identified (referred to as GST-ALS2452-668).

3.3.2 GST fusion protein expression and purification

The bacterial expression plasmid GST-ALS2452-668 was transformed into BL21-RIL E. coli and the protein was expressed as described (Section 2.2.3). Samples taken during expression were analysed by SDS-PAGE followed by staining with Coomassie Blue and immunoblotting (Figure 3.3). The GST-ALS2452-668 fusion protein was identified as a band migrating at approximately 55 kDa (which is the predicted size) and the presence of this protein was confirmed by immunoblotting with a polyclonal antibody to GST

(Figure 3.3 A, B). The majority of this protein was found in the insoluble fraction.

Empty pGEX-5X-1 was expressed under the same conditions, as a control; the increase in expression with time of a band migrating at approximately 29 kDa, mainly in the soluble fraction, confirmed that GST was expressed in abundance under these conditions (Figure 3.3 C and D). The conditions were optimised for maximal expression of both the GST control and GST-ALS2452-668 by varying the amount of IPTG (0.1 mM,

0.25 mM and 0.4 mM) and the temperature during expression (30°C and 37°C).

However, the GST-ALS2452-668 fusion protein was expressed in the insoluble fraction in all of the above conditions, and the maximal expression was with a temperature of 37°C and induction with 0.4 mM IPTG. After capture with glutathione sepharose beads,

127 elution with reduced glutathione buffer and concentration, the GST-ALS2452-668 fusion protein was analysed and quantified by SDS-PAGE (Figure 3.4).

3.3.3 Characterisation of ALS2 antibody

Following production of the polyclonal ALS2 antibody, the antibody was characterised by immunoblotting and immunofluorescence. Myc-tagged ALS2 was expressed in CHO cells and 16 hours later cells were harvested in lysis buffer, boiled in 2x SDS protein sample buffer and subjected to SDS-PAGE and immunoblotting. Equal amounts of

CHO lysate were probed with varying amounts of ALS2 antibody in order to determine the specificity of the antibody and the optimal concentration for detection of ALS2 protein (Figure 3.5 A). A single species was detected in ALS2 transfected but not non- transfected CHO cells, with both the ALS2 and myc (9B11) antibodies (Figure 3.5 A).

This reactive species migrated at approximately 185 kDa, which is the predicted size for the myc-tagged ALS2 construct, and the optimal concentration of the ALS2 antibody was determined to be 1:1000. As negative controls, preimmune serum from the rabbit that produced the ALS2 antibody was used in place of a primary antibody on one blot, and primary antibody was omitted on another.

Immunoblotting of rat or mouse brain homogenates with the polyclonal ALS2 antiserum was initially unsuccessful. ALS2 is now known to be a low abundance protein, and is thought to represent only approximately 0.0003% of the total detergent- soluble fraction of mouse brain lysate (Yamanaka et al., 2003). The antibody was therefore affinity-purified with GST-ALS2452-668 antigen. The affinity-purified antibody resulted in lower non-specific background, allowing blots to be exposed for a longer period of time, which is more efficient for detecting low-abundance proteins. The affinity-purified ALS2 antibody was able to detect a band of the predicted molecular

128 mass of ALS2 in rat brain homogenate, co-migrating with the 185 kDa band in transfected CHO cell lysate (Figure 3.5 B). To further determine the specificity of the antibody, it was incubated with 10x molar excess of the antigen prior to incubation with the blot. This preabsorption with the antigen resulted in the loss of the 185 kDa band, whereas incubation with GST alone did not affect the intensity of the band. Therefore, the ALS2 sequence of antigen (rather than the GST sequence) was effectively

“competing out” the band, which indicates that the band detected is ALS2.

CHO cells transfected with myc-tagged ALS2 were also subjected to immunofluorescence; cells were incubated with the ALS2 antibody at a concentration of

1:100 and anti-myc (9B11) antibody. The pattern of staining obtained with ALS2 antibody was identical to that of 9B11 (Figure 3.5 C). The overexpressed ALS2 protein appeared to be localised to the cytoplasmic (particularly perinuclear) region of the cell, but excluded from the nucleus.

3.3.4 Localisation of ALS2 in brain and spinal cord

The ALS2 antibody was used to detect endogenous ALS2 in rat brain and spinal cord sections by immunohistochemistry (Figure 3.6). Frozen brain and spinal cord sections were incubated with ALS antibody at a concentration of 1:100 and staining was observed in a variety of neuronal populations. This staining was found to be specific for

ALS2 as preincubation of the antibody with GST-ALS2452-668 resulted in a lack of immunoreactivity (Figure 3.6). High magnification of the sections revealed that ALS2 staining within neurons appears to be in both the cell body and neuronal processes.

Interestingly, ALS2 was detected in large neurons in the spinal cord that are located in a motor neuron-rich area of the spinal cord. Therefore, ALS2 seems to be present in

129 motor neurons. Particularly intense staining was seen in neurons of the cerebellum, and

ALS2 staining was also observed in neurons of the cortex, hippocampus and brainstem.

The affinity-purified ALS2 antibody was then used to immunoprecipitate endogenous

ALS2 from rat brain. A single band was identified, co-migrating with ALS2 immunoprecipitated from transfected CHO cells using anti-myc antibody 9B11, at approximately 185 kDa (Figure 3.7). Thus, the ALS2 antibody produced in these studies successfully detects ALS2 by immunofluorescence, immunoblotting, and immunohistochemistry, and can immunoprecipitate ALS2 from rat brain.

3.3.5 Overexpressed ALS2 is localised in cytoplasmic and membranous fractions of mammalian cells

Differential centrifugation of ALS2-transfected CHO cell homogenate was used to determine the subcellular localisation of overexpressed myc-tagged ALS2. ALS2 was found to be localised to the cytoplasmic and membrane fractions but not the nuclear fraction (Figure 3.8). To check the purity of the prepared fractions, the samples were probed for endogenous proteins that are known to be specific to the cytoplasmic, membrane and nuclear fractions (α-tubulin, presenilin1 and c-jun respectively). This confirmed that the different fractions had been separated successfully.

3.4 Discussion

In these studies, a polyclonal antibody that is specific to endogenous ALS2 was prepared and used successfully for immunoblotting, immunofluorescence, immunohistochemistry and immunoprecipitation. The antibody detected a major species of 185 kDa, which is the predicted size of the long form of ALS2. Specificity of the antibody has been demonstrated by pre-absorption with the antigen. One further test for specificity would be to use on ALS2 mouse knockout tissue, or cells expressing an

130 ALS2 RNAi construct, although as yet neither of these have as yet been published.

Recently, the ALS2 C-terminal like (ALS2CL) gene has been identified as an ALS2 homolog, which encodes a 108 kDa protein ‘ALS2CL’ (Devon et al., 2005; Hadano et al., 2004). Although the antigen used to prepare the antibody in these studies shows approximately 92% sequence similarity to ALS2CL, no band migrating at 108 kDa was seen in immunoblots of rat brain homogenate or cultured rat embryonic cortical neurons, and therefore this antibody might not detect ALS2CL.

To investigate the subcellular localisation of ALS2, a myc-tagged ALS2 construct was overexpressed in CHO cells, and immunofluorescence and cellular fractionation (by differential centrifugation) studies were carried out. ALS2 was found to be mainly localised to the cytoplasm and membrane fractions and excluded from the nucleus, which is in agreement with studies of endogenous ALS2 in rodent brain and spinal cord

(Devon et al., 2005; Topp et al., 2004; Yamanaka et al., 2003). However, the cell fractionation studies reported here show a higher level of ALS2 in the cytoplasmic fraction compared with the membrane fraction. This is most likely due to these experiments being carried out in transfected CHO cells and this overexpression of ALS2 would result in higher levels in the cytoplasm. Immunohistochemical staining revealed a high abundance of ALS2 in cerebellar neurons, which is consistent with the finding that the cerebellum is the brain region with the highest ALS2 expression in immunoblots of mouse brain homogenates (Yamanaka et al., 2003). Most interestingly, ALS2 was detected in large cells in spinal cord which closely resemble motor neurons. Thus it appears that ALS2 is present in motor neurons, the cell type that is predominantly affected in MND.

131 Figure 3.1 Schematic of ALS2 Long-form, ALS2 Short-form and GST-ALS2452-668

Schematics of ALS2 Long-form, ALS2 Short-form and the predicted GST-ALS2452-668 fusion protein are shown as indicated. The ALS2 fragment corresponds to amino acids 452-668 of human ALS2, which encompasses sequences of the RCC1-like domain.

ALS2 Long-form RCC1 DH PH MORN VPS9

1 59 167 218 525 576 627 690 885 1049 1244 1551 1657 169 578 9011657 1007 1656

ALS2 Short-form Unique 24 residues

1 372 396

GST-ALS2452-668

452 668

132 Figure 3.2 PCR amplification of sequences encoding GST-ALS2452-668

GST-ALS2452-668 fusion protein was created by PCR and the products were subjected to agarose gel electrophoresis for analysis. PCR products resulting from reactions carried out with Taq DNA polymerase at different annealing temperatures (55°C, 60°C, 65°C and 70°C) were as shown. The PCR product ran at the predicted size of approximately 670 base pairs (bp). L is DNA ladder ΦX174HaeIII. Subsequent PCR reactions were carried out with Pfu DNA polymerase with an annealing temperature of 60°C. PCR products were cloned into pBluescript SK+, clones were identified by colour selection and restriction digest with EcoR1, and positive clones were sequenced (MWG Biotech). The insert from a chosen clone was then subcloned into pGEX-5X-1 as an EcoR1 fragment, and positive clones were identified by an analytical digest with EcoR1 and BamH1 followed by agarose gel electrophoresis.

L 55°C 60°C 65°C 70°C bp 1353 1078 872 603 310

133 Figure 3.3 GST-ALS2452-668 and GST protein expression

GST-ALS2452-668 fusion protein and GST were expressed in BL21-RIL E. coli and samples were taken when uninduced (0 hrs) and at hourly time points post-induction with 0.4 mM IPTG (1 hr, 2 hrs and 3 hrs). Samples were separated into total (T), soluble (S) and insoluble (I) fractions; proteins were separated by SDS-PAGE and either stained with Coomassie Blue (A, C) or immunoblotted and probed for GST-ALS2452-668 or GST using a rabbit polyclonal GST antibody (Sigma) (B, D). GST control was found to be expressed mainly in the soluble fraction (migrating at approximately 29 kDa) whereas

GST-ALS2452-668 was found mainly in the insoluble fraction. GST-ALS2452-668 was found to migrate at its predicted size of ~55 kDa and the presence of the fusion protein in the samples was confirmed by immunoblotting with anti-GST.

A 0 hrs 1 hr 2 hrs 3 hrs kDa T S I T S I T S I T S I 220 97

66 GST-ALS2452-668 45

30

B GST-ALS2452-668

C 0 hrs 1 hr 2 hrs 3 hrs

T S I T S I T S I T S I kDa

66 45

30 GST D GST

134 Figure 3.4 Purification of GST-ALS2452-668

GST-ALS2452-668 was expressed in BL21-RIL cells for 3 hrs, solubilised in 6 M urea, dialysed against PBS and then captured with glutathione sepharose 4B beads. The GST-

ALS2452-668 antigen was eluted by addition of reduced glutathione buffer, concentrated by centrifugal filtration and analysed by SDS-PAGE alongside BSA standards (stained with Coomassie Blue). The sample shown represents 1/60th of the yield from 200 ml of E. coli. The total yield was therefore approximately 150 µg per 100 ml of E. coli culture.

BSA standards (μg) GST- kDa 0.5 1 2 5 ALS2452-668 97

66

45

135 Figure 3.5 Characterisation of ALS2 antibody (A) non-transfected CHO cells (NT) and CHO cells transfected with myc-tagged ALS2 (ALS2) were subjected to SDS-PAGE and immunoblotting with varying concentrations of rabbit polyclonal ALS2 antisera (1:100, 1:500, 1:1000, 1:2500; 1:25,000) or with myc antibody (9B11). A single band was observed at the predicted size of approximately 185 kDa. Blots were incubated with pre-immune serum (PI) or without a primary antibody (No primary) as controls. (B) CHO lysates (NT and ALS2- transfected) and adult rat brain homogenate (RB) were subjected to SDS-PAGE and immunoblotting with affinity-purified ALS2 antibody at a concentration of 1:1000, either without or with (Preabsorbed GST-ALS2452-668) incubation with 10x molar excess of GST-ALS2452-668 antigen. Incubation of affinity-purified ALS2 antibody with GST protein was carried out as a control (Preabsorbed GST). Samples were also immunoblotted without primary antibody as a negative control (No antibody). The 185 kDa band was not detected upon preabsorption specifically with the GST-ALS2452-668 antigen. (C) immunofluorescence was carried out in CHO cells transfected with myc- tagged ALS2, co-staining with both rabbit polyclonal ALS2 antisera and mouse monoclonal myc (9B11) antibodies. Scale bar is 50 µm.

A PI 1:100 1:500 1:1000 1:2500 1:25,000 No primary 9B11 kDa NT ALS2 NT ALS2 NT ALS2 NT ALS2 NT ALS2 NT ALS2 NT ALS2 NT ALS2 220

97

66 45

B Preabsorbed Preabsorbed ALS2 antibody GST-ALS2452-668 GST No antibody NT ALS2 RB ALS2 RB ALS2 RB ALS2 RB kDa

220

97

C ALS2 Myc (9B11)

136 Figure 3.6 Localisation of ALS2 in adult rat brain and spinal cord sections ALS2 immunoreactivity is seen in various brain and spinal cord regions. (1) is Cortex incubated with the ALS2 antibody preabsorbed with the GST-ALS2452-668 immunogen. (2) is Cortex (without preincubation with immunogen). (3) is Brainstem, (4) is Cerebellum and (5) is Hippocampus. (6) is cervical spinal cord and (7) is high magnification of the area outlined in (6). Scale bars are 200 µm (1-5) and 100 µm (6,7).

137 Figure 3.7 Affinity-purified ALS2 antibody immunoprecipitates ALS2 from rat brain CHO cells were transfected with myc-tagged ALS2 and immunoprecipitation was carried out with (+) or without (-) 9B11 antibody to the myc-tag. Adult rat brain (RB) was homogenised and immunoprecipitation was carried out with (+) or without (-) affinity-purified ALS2 antibody. The samples were then probed on immunoblots with affinity-purified ALS2 antibody. An approximately 185 kDa species was immunoprecipitated, both from transfected CHO cells with 9B11 antibody and from rat brain with affinity-purified ALS2 antibody.

CHO RB kDa - + - +

182

120

88

138 Figure 3.8 ALS2 is present in the cytosolic and membrane fractions of transfected CHO cells CHO cells were transiently transfected with myc-tagged ALS2 (ALS2) and homogenised in buffer containing 1% Triton (see section 3.2.6). Homogenates were separated into cytosolic (C), membrane (M) and nuclear (N) fractions by differential centrifugation. Samples were then subjected to SDS-PAGE and immunoblotting using ALS2 antibody. NT and ALS2 are samples of non-transfected and ALS2-transfected CHO cell lysates. Samples were also probed for c-jun (nuclear-specific protein), presenilin 1 (PS1; membrane-specific protein) and α-tubulin (cytosol-specific protein) as fractionation controls.

NT ALS2 C M N

ALS2

c-jun

PS1

α-tubulin

139

CHAPTER 4: ALS2 ACTS AS GEF FOR RAC1 AND

ACTIVATES PAK1

140 4.1 Introduction

Although the function of ALS2 is as yet unclear, its amino acid sequence predicts that it functions as a GEF. GEFs positively regulate the activity of members of the Ras superfamily of GTPases. These GTPases cycle between inactive (GDP-bound) and active (GTP-bound) conformational states and GEFs stimulate GTP-binding so as to promote activation of the GTPase (For reviews see Etienne-Manneville and Hall, 2002;

Schmidt and Hall, 2002). ALS2 contains three putative GEF domains; an amino- terminal domain that displays homology to the Ran GEF RCC1, a central region containing Dbl and pleckstrin homology (DH/PH) domains that are found in GEFs for members of the Rho family of GTPases (for example Rho, Rac and Cdc42), and a carboxyl-terminal vacuolar protein sorting 9 (VPS9) domain which is found in GEFs for

Rab5 (Hadano et al., 2001; Yang et al., 2001). Indeed, there is now evidence that ALS2 functions as a GEF for Rab5, and this is via its VPS9 domain (Otomo et al., 2003; Topp et al., 2004).

To investigate the hypothesis that ALS2 functions as a GEF for the Rho family of

GTPases via its DH/PH domain, I have studied the level of activation of the three most characterised Rho family GTPases (Rho, Rac and Cdc42) in the presence of ALS2.

These studies were carried out in mammalian (CHO) cells. I have demonstrated that

ALS2 causes an increase in activation of Rac but not Rho or Cdc42. The catalytic activity of Dbl-family (DH/PH) GEFs is predicted to be mediated by the Dbl homology

(DH) domain, based on studies of the Dbl protein (Hart et al., 1994), and this has subsequently been confirmed for a number of GEFs (Aghazadeh et al., 1998; Kubiseski et al., 2003; Matsuo et al., 2002; Movilla and Bustelo, 1999; Whitehead et al., 1995).

Therefore, I have disrupted the DH domain of ALS2 by deletion of its key residues, and have found that an intact DH domain is necessary for its function as a Rac GEF.

141 Active (GTP-bound) Rac binds to a variety of effectors including members of the p21- activated kinase (PAK) family. In order to further elucidate the downstream components of ALS2-Rac1 signalling, I have investigated the activation of PAK1 by ALS2, using in vitro protein kinase assays. These studies have revealed that ALS2 overexpression results in activation of PAK1 and that this activation is dependent on ALS2 Rac GEF activity (as shown by the use of the ALS2 mutant described above). Thus, ALS2 has the ability to function as a Rac GEF and activate PAK1 in mammalian cells.

4.2 Methods

4.2.1 Plasmids

Plasmids used were as follows: RhoA, Rac1, Cdc42, L61Rac1, Net1ΔN, VavΔN and

Fgd1(DH/PH) (all in pRK5myc expression vector), myc-tagged ALS2 (in pCIneo), pCIneo-CAT and PAK1 (in pCMV6myc) were as described (Table 2.2). ALS2(DH/PH) was prepared by Anja Schmidt (University College London): Briefly, the DH/PH GEF domain (amino acids 681-1010) of ALS2 was generated by PCR and cloned as a

BamH1-EcoR1 fragment into pRK5myc.

A mutant ALS2 clone (ALS2ΔDH) in which the DH/PH GEF domain was disrupted was created, by deletion of sequences encoding the DH domain (residues 747-826), which has been identified as the region necessary for the catalytic GEF activity of the

Dbl-family GEFs. The mutagenesis was performed using an ExSite™ PCR-based site- directed mutagenesis kit as described in Section 2.2.2. The following 5’-phosphorylated primers were used (Oswel DNA service, University of Southampton):

Primer 1: 5’-ACTCAGTTGATGGAAATACTGAATAC-3’

Primer 2: 5’-ACACAGCTTGCTGAATCGGCTAGC-3’

142 PCR products were analysed by restriction enzyme digest and sent for sequencing to confirm that the mutagenesis had been successful (MWG Biotech).

4.2.2 Cell culture and transfection

Transfections were carried out with Lipofectamine™ reagent as described in Section

2.2.6.3. For GTPase activation assays, CHO cells were cultured in 10 cm dishes and transiently co-transfected with 2 µg of either myc-tagged RhoA, Rac1 or Cdc42 and 8

µg of ALS2 or pCIneo-CAT cDNA. As positive controls, CHO cells were co- transfected with RhoA, Rac1 or Cdc42 and constitutively active GEFs for Rho, Rac or

Cdc42 (Net1ΔN, VavΔN and Fgd1(DH/PH) respectively). As a negative control for the

Rac activation assay (and to investigate the ability of ALS2ΔDH, an ALS2 mutant in which the DH domain is disrupted, to act as a GEF for Rac) cells were co-transfected with 2 µg Rac and 8 µg ALS2ΔDH. Additionally, the ability of the isolated DH/PH domain to stimulate activation of Rac was investigated and so CHO cells were transfected with Rac1 and myc-tagged ALS2(DH/PH). For PAK1 kinase assays, CHO cells were cultured as above and co-transfected with 2 µg myc-tagged PAK1, 2 µg Rac1 and 6 µg of either pCIneo-CAT as a negative control, ALS2, ALS2ΔDH or constitutively active Rac1 (L61Rac1) as a positive control.

4.2.3 GTPase activation assays

CHO cells were cultured in 10 cm dishes and co-transfected as described above. Cells were harvested 24 hours after transfection, following 16 hours of serum-starvation (in order to obtain a low basal level of Rho family GTPase activation (Jalink et al., 1993;

Jalink et al., 1994; Tigyi and Miledi, 1992)). Cellular Rho, Rac and Cdc42 activities were assayed using commercially available kits essentially according to the manufacturer’s instructions (Upstate) as described in Section 2.2.7. Captured Rho, Rac

143 and Cdc42 were detected on immunoblots using anti-myc antibody (9B11) and the relative amounts quantified by pixel densitometry using a Bio-Rad GS710 imaging densitometer and Quantity 1 software.

4.2.4 In vitro kinase assays

CHO cells were cultured in 10 cm dishes and co-transfected as described above. Cells were harvested 24 hours after transfection, following 16 hours of serum-starvation.

PAK1 was immunoprecipitated from the lysate using 1.5 µg of anti-myc epitope tag

(9B11) antibody, and in vitro kinase assays were carried out in kinase assay buffer

(Section 2.1.8) with 0.26 MBq γ[32P]-ATP, 20 μM ATP and 5 µg myelin basic protein

(MBP) as substrate, as described in Section 2.2.8.3. Samples were boiled in 2x SDS protein sample buffer and subjected to SDS-PAGE followed by staining with

Coomassie Blue and autoradiography, or immunoblotting.

4.3 Results

4.3.1 ALS2 acts as a GEF for Rac1 but not RhoA or Cdc42

ALS2 contains a central DH/PH domain which shows homology to GEFs that regulate the Rho family of GTPases. To determine whether ALS2 activates the 3 most widely- studied Rho family GTPases RhoA, Rac1 and Cdc42, in vivo pull-down assays were utilised to monitor the activities of these GTPases in transfected CHO cells. Active

(GTP-bound) Rho binds to the Rhotekin “Rho binding domain” (RBD; corresponding to residues 7-89 of mouse Rhotekin) whereas active Rac and Cdc42 both bind the PAK1

“p21 binding domain” (PBD; corresponding to residues 67-150 of human PAK1). GST-

RBD and GST-PBD “baits” can thus be used to isolate GTP-bound RhoA, Rac1 and

Cdc42 from experimentally manipulated cells; the amounts of these GTPases detected on immunoblots correlates with their activities.

144 Co-transfection of cells with ALS2 induced a significant increase in the amount of Rac1 but not Cdc42 pulled-down by GST-PBD (Figure 4.1B, C). ALS2 did not activate

RhoA (Figure 4.1A). Control samples, in which GST-PBD or GST-RBD was substituted with GST alone, did not result in activation of Rac, Rho or Cdc42. Similarly, pre-loading with GDP resulted in no active (GTP-bound) GTPase in the cells.

Transfection of dominant active GEF constructs, or pre-loading with non-hydrolysable

GTPγS, resulted in significant stimulation of GTPase activity in these assays (Figure

4.1). The isolated DH/PH GEF domain (residues 681-1010) did not activate Rac1 in these assays, which suggests that other regions of ALS2 may be necessary for control of its Rac1 GEF function (Figure 4.1 D). Thus, ALS2 stimulation of Rac1 activity requires the full-length ALS2 protein.

To confirm that the stimulatory effect on Rac activity was indeed mediated by the

DH/PH domain, an ALS2 mutant (ALS2ΔDH) was created in which the majority of the catalytic DH domain (residues 747-826) was deleted, using the ExSite™ PCR-based site-directed mutagenesis system (Figure 4.2 A, B, C). Rac activation assays were carried out as described above. ALS2ΔDH did not activate Rac1 (Figure 4.2 D), which shows that the ALS2 Rac1 GEF activity occurs via its catalytic DH domain.

4.3.2 ALS2 activates PAK1 and this is dependent on a functional DH domain

Immediate downstream targets for Rac include members of the p21-activated kinase

(PAK) family of serine/threonine kinases. PAK1 is a major neuronal member of the

PAK family. The ability of ALS2 to stimulate PAK1 activity was therefore investigated.

In vitro PAK1 kinase assays were performed from CHO cells co-transfected with PAK1

+/- ALS2 as described above, using MBP as a substrate. These experiments revealed that ALS2 stimulated PAK1 activity by approximately 1.5-fold, as measured by the

145 extent of phosphorylation of MBP. This stimulation was not seen in cells co-transfected with the non-functional mutant ALS2ΔDH, which indicates that this activity is dependent on a functional DH/PH domain and implicates Rac activity in ALS2 activation of PAK1 (Figure 4.3). Thus, ALS2 functions as a GEF to regulate Rac1-

PAK1 signalling.

4.4 Discussion

In these studies I have demonstrated that ALS2 functions as a Rac GEF via its DH domain, and have identified PAK1 as a downstream target of this activation. During the course of these studies (as discussed in Section 1.2.2), two research groups also reported that ALS2 functions as a Rac GEF. This was demonstrated by the use of similar pull- down assays utilising GST-PBD, in insect sf9 cells (Topp et al., 2004) and in CHO cells

(Kanekura et al., 2005). The level of Rac1 activation observed upon overexpression of

ALS2 was similar to that reported here. However, Kanekura et al. (2005) studied levels of endogenous Rac1 whereas in the studies reported here Rac1 was transfected into the

CHO cells. In agreement with the findings reported here, no activation of endogenous

Cdc42 or RhoA (pulled down with GST-RBD) was observed (Kanekura et al., 2005).

Furthermore, these authors used a mutant ALS2, in which a key threonine residue of the

DH domain had been mutated to alanine (residue 701) to demonstrate that a non- mutated DH domain was necessary for Rac activation (Kanekura et al., 2005), which confirms the findings reported here that the DH domain mediates ALS2 Rac GEF activity, as discussed below.

In addition, studies using in vitro GDP-dissociation assays have shown that ALS2 is a

GEF for Rab5 (Otomo et al., 2003; Topp et al., 2004). The stimulatory effect of ALS2 on Rab5 required its VPS9 domain (Otomo et al., 2003). The authors of this paper also

146 studied the effect of ALS2 on Rac activity but were unable to detect any stimulation of

Rac. This discrepancy is probably due to the different assays used. Topp et al. (2004),

Kanekura et al. (2005) and the studies reported in this thesis all utilised cellular Rac activation assays whereas Otomo et al. (2003) used in vitro exchange assays. Several

GEFs require specific components for their regulation that are only present in a cellular environment. For example the Rac GEF activity of SWAP-70 is minimal in in vitro

GDP dissociation assays but is greatly enhanced upon addition of phosphatidylinositol

3,4,5-triphosphate (PtdIns(3,4,5)P3) (Shinohara et al., 2002), and the activity of several other GEFs, including Vav, Sos and Tiam1, is enhanced by binding to phospholipids such as PtdIns(3,4,5)P3 (Das et al., 2000; Fleming et al., 2000; Han et al., 1998).

Therefore, it is possible that ALS2 may only act as a Rac GEF in vivo.

ALS2-dependent activation of Rac was found to require the DH domain, which has previously been identified as the region responsible for catalytic GEF activity in other

Rac GEFs such as Dbl (Hart et al., 1994). However, overexpression of the isolated

DH/PH region of ALS2 did not result in stimulation of Rac activity. This may be due to autoinhibition of the catalytic DH domain by residues of the PH domain. Indeed, this is known to occur for a number of GEFs including Sos1 and P-Rex1 (Nimnual et al.,

1998; Welch et al., 2002) although some isolated DH/PH constructs can act in a constitutively active manner; for example Fgd1(DH/PH) is a constitutively active Cdc42

GEF (Olson et al., 1996). Alternatively, regions of the full-length ALS2 protein other than the DH/PH domain may be necessary for regulation of the Rac GEF activity. For example, MORN motifs are implicated in association with the plasma membrane

(Takeshima et al., 2000) and this may be necessary for the Rac GEF activity of ALS2, as the activity of several Rac GEFs is dependent upon membrane-targeting (Whitehead et al., 1995; Whitehead et al., 1999). Although the regulatory regions of ALS2 are as yet

147 unidentified, this study shows that the full-length ALS2 protein is necessary for activation of Rac.

Finally, overexpression of ALS2 was found to activate PAK1, a major downstream effector of Rac1. The level of activation of PAK1 observed, as measured by phosphorylation of MBP, was approximately 1.5-fold. This is similar to the level of

PAK1 activation resulting from overexpression of the Rac1/Cdc42 GEF α-PIX in 293T human embryonic kidney cells (Yoshii et al., 2001). Thus, ALS2 functions as a Rac

GEF to activate PAK1. This activation requires a functional DH domain, as was demonstrated by overexpression of the non-functional ALS2ΔDH mutant.

Rac1 is involved in a variety of cellular activities including organisation of the actin cytoskeleton, regulation of cell proliferation, motility and survival, generation of ROS, and induction of both JNK and p38 signalling pathways (for reviews see Nobes and

Hall, 1995a; Van Aelst and D'Souza-Schorey, 1997). Of particular relevance to motor neuron disease is the role Rac plays in cell survival and death. Rac activates both apoptotic (through the activation of JNK) and anti-apoptotic (through direct interaction with PI(3)K and subsequent activation of the serine/threonine kinase Akt/PKB) pathways (Lee et al., 2004; Murga et al., 2002). Indeed, it has recently been reported that ALS2 mediates neuroprotection by sequential activation of Rac1, PI(3)K and Akt3 and this protection is specific against mutant SOD1 toxicity (Kanekura et al., 2005).

Six members of the PAK family (PAK1-6) have been identified; PAK1-3 are major downstream effectors of Rac and Cdc42, possessing a CRIB domain/p21 binding domain (PBD) which binds specifically to active Rac and Cdc42, whereas PAK4-6 differ significantly in their structure and regulation. For a review see (Bokoch, 2003).

148 Of the three serine/threonine kinases PAK1, PAK2 and PAK3, PAK1 has so far been the most widely studied, and is an important regulator of many cellular events including cytoskeletal dynamics and cell motility, transcription through SAPK/JNK and p38 kinase cascades, cell death and survival signalling and cell-cycle progression (Daniels et al., 1998; Frost et al., 1998; Frost et al., 2000; Frost et al., 1996; Robinson and Cobb,

1997; Schurmann et al., 2000; Sells et al., 1999; Sells et al., 1997; Zhang et al., 1995).

Several downstream effectors of PAK1 have been identified including Filamin,

Paxillin/PIX/PKL complex, the adaptor protein Nck, Merlin (a member of the

Ezrin/Radixin/Moesin family, which is thought to link actin filaments to the plasma membrane), Stathmin (a protein that regulates microtubule stability), and LIM kinase 1

(LIMK1) and myosin light chain kinase (MLCK), which control actin dynamics via their substrates Cofilin and myosin light chain, respectively (Chew et al., 1998; Manser et al., 1997). More recently, PAK1 has also been implicated in the regulation of microtubule dynamics by phosphorylating tubulin cofactor B (Vadlamudi et al., 2005).

Thus the combined studies presented here and in other reports show that ALS2 is a GEF for Rab5 and Rac and that ALS2-mediated Rac activity leads to activation of one of its downstream effectors PAK1 (Otomo et al., 2003; Topp et al., 2004; Kanekura et al.,

2005). Since ALS2 disease-causing mutations are all predicted to result in a loss of its function, it seems likely that perturbation of Rab5 and/or Rac/PAK signalling induces disease in the ALS2 kindreds.

149 Figure 4.1 ALS2 stimulates Rac activity GTPase activation assays were conducted in transfected CHO cells. (A, B, C) show representative assays for the three GTPases RhoA, Rac1 and Cdc42 respectively. Active RhoA was pulled-down from the lysates using GST-RBD and active Rac1 and Cdc42 were pulled-down with GST-PBD. The bound Rho, Rac and Cdc42 were then detected by immunoblotting (Active). The amounts of Rho, Rac and Cdc42 transfected into the cells were also detected by immunoblotting so as to demonstrate equal transfection efficiencies (Total). GST alone did not bind any GTPase in the transfected cells (No RBD/ No PBD). CHO cell samples were incubated with GDP as a negative control and with GTPγS as a positive control. CHO cells were transfected with the constitutively- active GEF mutants Net1ΔN, VavΔN and Fgd1DH/PH as positive controls for the Rho, Rac and Cdc42 assays respectively. (D) shows immunoblots from a Rac1 activation assay demonstrating that, unlike full-length ALS2, the isolated DH/PH domain of ALS2 does not activate Rac1. (E) shows histogram of fold-increases in GTPase activity, obtained from the average of 4 separate experiments for each GTPase. Error bars are SEM. One-Way ANOVA tests showed that ALS2 significantly increased Rac activity by 3.3 fold (p=<0.001) but not Rho (p=0.113) or Cdc42 (p=0.858) activities.

A CAT ALS2 Net1ΔN GDP GTPγS No RBD Active RhoA Total

B CAT ALS2 GDP GTPγS No PBD VavΔN Active Rac1 Total

C CAT ALS2 GTPγS GDP Fgd1DH/PH No PBD Active Cdc42 Total

D E CAT DH/PH ALS2 Active Rac1 Total

150 Figure 4.2 ALS2ΔDH does not activate Rac1 A mutant ALS2 (ALS2ΔDH), in which the central region of the Dbl-homology (DH) domain (residues 747-826) was deleted, was created by PCR-based site-directed mutagenesis. PCR products were analysed by agarose gel electrophoresis before (A; lane 1) and after (A; lane 2) digestion with Dpn1 to remove template DNA. The PCR products migrate at the predicted size of 10,185 bp. (A; lane 3) is linearised (by digestion with Xho1) 10,426 bp template DNA (pCIneo-ALS2). L is DNA ladder λHindIII. Following ligation, clones were isolated and subjected to restriction enzyme digest with Sal1 (B). Lane 1 is template (pCIneo-ALS2). Fragment sizes are 7,697 kb and 2,729 kb. Lane 2 is an incorrect clone, with fragments migrating at the same sizes as the template DNA. Lane 3 is a successfully-mutagenised clone (fragment sizes are 7,697 kb and 2,488 kb). L is DNA ladder λHindIII. The successfully-mutagenised clone (ALS2ΔDH) was transfected into CHO cells and the CHO cell lysate was subjected to SDS-PAGE and immunoblotting with anti-myc (9B11) antibody. (C) shows an immunoblot of lysates of CHO cells transfected with either ALS2 or ALS2ΔDH. ALS2ΔDH migrates slightly faster than ALS2. Rac activation assays were carried out in CHO cells co-transfected with Rac1 and either pCIneo-CAT (CAT) as a negative control, ALS2 or ALS2ΔDH. (D) shows an immunoblot of amounts of Rac1 bound to GST-PBD (Active). The amounts of Rac1 transfected into the cells were also detected by immunoblotting so as to demonstrate equal transfection efficiencies (Total). In contrast to ALS2, ALS2ΔDH did not activate Rac1. A B L 1 2 3 L 1 2 3 bp bp

23,130 23,130 9416 9416 6557 6557 4361 4361 2322 2027 2322 2027

C D kDa ALS2 ALS2ΔDH

CAT ALS2ΔDH ALS2 Active 181.8 Rac1 Total

115.5

151 Figure 4.3 ALS2 activates PAK1 PAK1 in vitro kinase assays were performed from CHO cells co-transfected with PAK1, Rac1 and either ALS2, ALS2ΔDH or vector pCIneo-CAT (CAT) as a negative control. For a positive control, cells were co-transfected with constitutively active Rac1 (L61Rac1). (-) and (+) refer to absence or presence of PAK1 immunoprecipitating antibody in the reactions; reaction mix (RM) contains no immunoprecipitation sample. (A) shows the Coomassie Blue-stained gel and (B) the corresponding autoradiograph of the samples. (C) is an immunoblot detecting the levels of PAK1 in the immunoprecipitation samples. The assays were repeated a further three times and one- way ANOVA tests revealed that the stimulatory effect of ALS2 and L61Rac1 on PAK1 activity was significant (p<0.01). (D) is a histogram showing fold increase in MBP phosphorylation with ALS2 and ALS2ΔDH compared with control (CAT). Data is pooled from five independent experiments. Error bars are SEM.

A RM CAT ALS2 ALS2ΔDH L61Rac kDa 115 - + - + - + - + - + 82 64 48.8 IgG 37.1 25.9 IgG 19.4 14.8 MBP

B Phospho-MBP

C PAK1

D

152

CHAPTER 5: ALS2 IS PRESENT IN NEURONAL

GROWTH CONES AND PROMOTES NEURITE

OUTGROWTH

153 5.1 Introduction

Rho family GTPases play a major role in organising the cytoskeleton and in particular, the actin cytoskeleton. In developing neurons, Rho, Rac and Cdc42 are present in growth cones of axons and dendrites where they regulate actin-myosin contractility to control neurite outgrowth and growth cone guidance (Li et al., 2000b; Threadgill et al.,

1997; Dickson, 2001; Nikolic, 2002). In some paradigms, Rac and Cdc42 promote whereas Rho inhibits neurite outgrowth. For example, in cell lines such as N1E-115 neuroblastoma cells and PC12 pheochromocytoma cells, and in primary cell systems such as chick retinal neurons and rat hippocampal neurons, Rac1 and Cdc42 promote neurite outgrowth whereas RhoA causes growth cone collapse and neurite retraction

(Albertinazzi et al., 1998; Da Silva et al., 2003; Jalink et al., 1994; Kita et al., 1998;

Schwamborn and Puschel, 2004; Sebok et al., 1999). However, a few studies using constitutively active (CA) and dominant negative (DN) mutants of Rac1 have revealed effects that contrast with those described above. For example, CA Rac1 has been found to decrease the length of the longest neurite in cultured rat cortical neurons (Kubo et al.,

2002) and DN Rac1 promotes neurite outgrowth in chick dorsal root ganglion (DRG) neurons (Fournier et al., 2003). These differential effects of GTPases on neurite development may be due to the type or age of the cells used in these studies and/or culture conditions. Additionally, it has been suggested that the varying results observed with the use of CA or DN GTPase mutants may reflect the importance of GTPase cycling between active and inactive states in the regulation of neuritogenesis. This idea is supported by studies in which both CA and DN mutants result in the same effect. For example, both CA and DN Rac1 mutants inhibit growth cone advance and neurite outgrowth in primary chick embryo motor neurons (Kuhn et al., 1998). Therefore, it is evident that tight regulation of GTPase activation (by GEFs and GAPs) is required for normal neuronal outgrowth.

154 A number of GEFs for Rho family GTPases have now been shown to be involved in the development of axons and dendrites including GEFT, Kalirin, Trio, Tiam1, Vav2, Vav3 and STEF (Aoki et al., 2005; Bryan et al., 2004; Chakrabarti et al., 2005; Estrach et al.,

2002; Kunda et al., 2001; Leeuwen et al., 1997; Matsuo et al., 2002; May et al., 2002;

Penzes et al., 2001). Therefore, since ALS2 functions as a Rac GEF (Chapter 4), its role in neurite initiation and outgrowth was investigated.

5.2 Materials and Methods

5.2.1 Antibodies and immunofluorescence microscopy

Affinity-purified polyclonal ALS2 antibody (see Chapter 3) was used at 1:100 dilution in 5% FBS/PBS for immunofluorescence. See Table 2.3 for details and working concentrations of Rac and α-tubulin antibodies. For staining of endogenous F-actin,

Alexa Fluor568-phalloidin was added to the coverslip (5 U/ml working concentration in

5%FBS/PBS) 30 minutes after incubation with ALS2 primary antibody and Alexa Fluor

488 secondary antibody. The coverslips were then washed in PBS for 3x 10 minutes and mounted in Vectashield. Conventional images were captured using a Zeiss Axioscop microscope and CCD camera (Princeton Instruments), and confocal images captured using a Zeiss LSM 510 META confocal microscope.

5.2.2 Preparation of mouse brain homogenates

Brain tissue samples from E15, E18, P1, P7, P11, P21 and 1yr old mice were homogenised in 1 ml of ice-cold tissue homogenisation buffer (see section 2.1.5.1) per

0.1 g of tissue using a dounce homogeniser, passed through a 28-gauge needle three times and centrifuged at 15,000xg for 30 minutes. Samples were taken for protein concentration determination following which a quarter volume of 5x SDS protein sample buffer was added to the supernatant and the samples were boiled for 10 minutes,

155 aliquoted and stored at -70°C until required. Protein samples were analysed as described in section 2.2.4.

5.2.3 Neurite outgrowth measurements

Cortical neurons grown on PDL-coated coverslips were co-transfected (using 1 µl

Lipofectamine 2000™ reagent per coverslip) at 2 DIV, with 1 µg of plasmid pEGFPC.1

(Clontech) expressing GFP plus 3µg of experimental or control plasmids (see section

2.2.6.3). Experimental and control plasmids included ALS2, ALS2ΔDH, dominant negative Rac1 (N17Rac), dominant negative Rab5 (N39Rab5), and pCIneo-CAT which was used to balance transfections so that all cells received the same numbers and amounts of plasmid. 24 hours after transfection, cells were fixed in 4% (w/v) paraformaldehyde (PFA) in TBS for 20 minutes, and processed for immunofluorescence

(to ensure all plasmids were expressing in the cells examined) as described in section

2.2.11. Images were analysed by counting the number of neurites, and measuring the length of the longest neurite per cell, using Metamorph® image analysis software.

Neurite lengths were determined as the distance from the edge of the cell body to the growth cone tip, as visualised by GFP fluorescence. Measurement of the longest neurite of each cell has been used in numerous studies on the effects of signalling cascades on neurite outgrowth (e.g. (Nikolic et al., 1996; Penzes et al., 2001)). Cells were analysed without knowledge of the transfected plasmids, and only healthy cells as judged by morphology (including nuclear staining with Hoechst 33258 (Sigma) to confirm that nuclei had a non-apoptotic appearance) were analysed.

156 5.2.4 Statistics

Statistical significance was determined using one-way ANOVA tests followed by Tukey post-hoc tests for pair-wise comparison. Differences were considered significant at p<0.05.

5.3 Results

5.3.1 Developmental expression of ALS2

The developmental expression of ALS2 was studied in mouse brain homogenates and in cultured cortical neurons using immunoblotting. Immunoblots of mouse brains aged from embryonic day 15 (E15) to 1 year revealed the presence of similar levels of ALS2

(Figure 5.1 A). Total protein levels in the samples were indicated by re-probing the immunoblots with α-tubulin antibody (DM1A). Others have also reported that ALS2 expression does not change markedly in the rodent brain post E10 (Devon et al., 2005).

However, in immunoblots of rat cortical neuron lysates there was a small but consistent decrease in ALS2 levels in >7 DIV cultures, even though levels of tubulin remained equal (Figure 5.1 B). This suggests that although levels of ALS2 do not vary in whole mouse brain samples, ALS2 may be developmentally regulated in specific neuronal populations such as cortical neurons.

5.3.2 ALS2 is present in growth cones of hippocampal neurons, where it is co- localised with Rac1 and F-actin

Application of the ALS2 antibody to 2 DIV rat hippocampal and cortical neurons produced prominent staining within cell bodies; labelling of punctate structures within neurites was also detected (Figure 5.2). These results are similar to those described by others in 7 DIV rat hippocampal neurons (Topp et al., 2004). However, in the 2 DIV

157 neurons, labelling within growth cones of both axons and dendrites was also detected

(Figure 5.2 and Figure 5.3).

Growth cones are highly specialised structures at the leading edge of developing neurites, which control the rate and direction of neurite outgrowth in response to environmental cues. Growth cones comprise two domains: the central (C) domain and the peripheral (P) domain. The C domain contains microtubules whereas the P domain is actin-rich containing the most motile structures, the lamellipodia and filopodia

(Letourneau, 1983; Tosney and Wessells, 1983). Co-staining for ALS2 and tubulin or actin revealed that ALS2 was present throughout the whole growth cone (Figure 5.3 B,

C). Double labelling for ALS2 and Rac1 was also performed, and this revealed a close overlap in the distributions of the two proteins in growth cones (Figure 5.3 A). The findings reported here complement the studies in Chapter 4 which demonstrate that

ALS2 is a Rac GEF and together they suggest that ALS2 may function in growth cone dynamics and neurite outgrowth.

5.3.3 Overexpression of ALS2 (but not ALS2ΔDH) promotes neurite outgrowth in rat embryonic cortical neurons via a Rac-dependent mechanism

A number of Rac GEFs are involved in various aspects of neuronal development, including neurite outgrowth and branching (Bryan et al., 2004; Leeuwen et al., 1997;

Lundquist et al., 2001; Penzes et al., 2001; Shin et al., 2002). As ALS2 has been shown to activate Rac (see Chapter 4), overexpression of myc-tagged ALS2 and ALS2ΔDH in

2-3 DIV cultured embryonic rat cortical neurons was utilised in order to investigate the involvement of ALS2 in neurite outgrowth and branching.

158 Prior to these experiments, the subcellular localisation of transfected ALS2 and

ALS2ΔDH was investigated in 3 DIV neurons (Figure 5.4). Both proteins showed a similar subcellular distribution to that of endogenous ALS2 in 2 DIV hippocampal neurons (Figure 5.2) where they localised to cell bodies, axons, dendrites and growth cones. Transfected ALS2 therefore localised to the appropriate cellular compartments where endogenous ALS2 resides.

In order to investigate the role of ALS2 in neurite outgrowth, ALS2 was overexpressed in 2 DIV embryonic rat cortical neurons together with GFP as a marker for neuronal cell shape. Neurite outgrowth was analysed 24 hours later, by comparison of the length of the longest neurite per cell with controls (cells that had been co-transfected with pCIneo-CAT and GFP). Comparison of the length of the longest neurite per cell has been used in several similar studies, for example (Nikolic et al., 1996) and (Penzes et al., 2001).

Overexpression of ALS2 in these neurons resulted in a significant (1.5-fold) increase in length of the longest neurite compared with control cells (Figure 5.5). This level of stimulation is similar to that observed by other Rac GEFs, for example GEFT (Bryan et al., 2004). As this stimulation of neurite outgrowth was hypothesised to be due to

ALS2’s Rac GEF activity, the mutant ALS2ΔDH, which is not functional as a Rac

GEF, was overexpressed in place of ALS2. In contrast to the effect seen with ALS2,

ALS2ΔDH did not stimulate neurite outgrowth (Figure 5.5). Therefore, ALS2 is thought to promote neurite outgrowth via activation of Rac. To pursue this hypothesis further, the experiments were repeated with co-expression of a dominant negative Rac1

(N17Rac1). This GTPase contains a substitution of Asn for Ser at position 17, and acts as a dominant negative protein by competing with endogenous Rac for binding to

159 upstream GEFs, but not interacting with downstream targets, thereby blocking Rac- dependent signalling pathways (Ridley et al., 1992). Such dominant negative GTPases have been used in numerous studies to dissect out pathways by which other GEFs stimulate neurite outgrowth (Bryan et al., 2004; Penzes et al., 2001). Co-expression of

N17Rac with ALS2 effectively blocked the ALS2-induced stimulation of neurite outgrowth (Figure 5.5), which supports the hypothesis that promotion of neurite outgrowth by ALS2 is via a Rac-dependent pathway.

ALS2 has also been shown to activate Rab5 via its VPS9 domain (Otomo et al., 2003).

To investigate whether stimulation of neurite outgrowth ALS2 could be linked to its ability to activate Rab5, a dominant negative Rab5 (N39Rab5) protein was co-expressed with ALS2 in these cells. Co-expression of N39Rab5 did not alter the stimulatory effect of ALS2 on neurite outgrowth (Figure 5.5). Thus, ALS2 stimulates neurite outgrowth via activation of Rac and independently of its function as a Rab5 GEF.

Expression of N17Rac alone had no effect on neurite outgrowth which is consistent with a number of other reports (Arakawa et al., 2003; Bryan et al., 2004). However, Rac has been shown to both promote and inhibit neurite outgrowth, and it is likely that such conflicting results are due to the different types of neurons, ages and culture conditions used for experimentation (e.g. (Jin and Strittmatter, 1997; Kita et al., 1998; Kuhn et al.,

1998; Lamoureux et al., 1997; Luo et al., 1994; Ruchhoeft et al., 1999)). Similarly, expression of N39Rab5 alone had no effect on neurite outgrowth.

160 5.3.4 Overexpression of ALS2 does not affect the number of neurites per cell, or the extent of neurite branching, in cultured cortical neurons

In some paradigms, Rac and PAK signalling has been shown to affect the initiation of neurite growth and branching (Allen et al., 2000; Hayashi et al., 2002; Leeuwen et al.,

1997; Threadgill et al., 1997). Indeed, several Rac GEFs have been found to affect the number of neurites per cell and/or the extent of neurite branching, including Tiam1

(Kunda et al., 2001) and GEFT (Bryan et al., 2004). Furthermore, expression of dominant negative Rac1 leads to a significant reduction in the number of primary dendrites in rat cortical neurons (Threadgill et al., 1997) and the extent of branching in

Xenopus retinal ganglion cells (Ruchhoeft et al., 1999), whereas Rac1 induces neuritogenesis and neurite branching in chick neural retinal cells (Albertinazzi et al.,

1998). Therefore, the effect of ALS2 on the number of neurites per cell and the extent of neuronal branching was investigated. This was measured by counting the total number of neurites emanating from the cell body, and by counting the number of neurite tips per cell, respectively. Overexpression of ALS2 did not affect the number of neurites, or the extent of branching (Figure 5.6). However, the results were only obtained from one timepoint during development (at 3 DIV) and therefore it remains possible that ALS2 may affect neurite promotion and/or branching at other stages in development.

5.4 Discussion

The studies presented here demonstrate that endogenous ALS2 is present in neuronal growth cones where it is co-localised with F-actin. This suggests that ALS2 may be involved in actin dynamics during neurite formation, outgrowth and/or pathfinding.

Furthermore, ALS2 co-localised with Rac1 which supports the hypothesis that ALS2 exerts its Rac GEF activity within the growth cone. Similarly, others have shown that

161 overexpressed ALS2 co-localises with Rac1 and with actin at leading membrane edges and membrane ruffles of NIH3T3 cells (Topp et al., 2004).

In the studies presented here, overexpression of ALS2 resulted in stimulation of neurite outgrowth (as measured by the length of the longest neurite per cell, compared with controls) and this effect was not seen with the non-functional mutant ALS2ΔDH.

Similarly, the promotion of outgrowth by ALS2 was not observed upon co-expression of dominant negative Rac1 (N17Rac1). Therefore, it can be concluded that ALS2 stimulates neurite outgrowth via its Rac GEF activity. In contrast, the effect of ALS2 on neurite length was not altered by co-expression of dominant negative Rab5 (N39Rab5), which suggests that the Rab5 GEF activity of ALS2 is not involved in this signalling pathway. Although Rac signalling has been shown to affect the initiation of neurite growth and branching, overexpression of ALS2 did not affect the number of neurites or extent of branching in these neurons. However, it is possible that ALS2 may be involved in such cellular events at a different stage in development.

In addition to Rac1, its downstream effector PAK1 is also present within the growth cone, where it too can function in neurite outgrowth, although (as in the case of Rac1) this may depend upon the stage of development and type of neuron (Daniels et al.,

1998). Rac1-PAK1 signalling has additionally been implicated in dendrite initiation in cultured embryonic mouse cortical neurons (Hayashi et al., 2002). ALS2 activates both

Rac1 and PAK1, and therefore the ALS2-mediated outgrowth observed in these experiments may involve PAK1 activation. This could be confirmed by use of dominant negative PAK1 proteins in the neurite outgrowth experiments to determine whether

PAK1 activation is essential for ALS2-mediated outgrowth to occur. Although overexpression of PAK1 has been reported to induce neurite outgrowth, it has been

162 suggested that this may be independent of its kinase activity (Daniels et al., 1998) and that PAK may be involved upstream of Rac1 (Obermeier et al., 1998). Similarly another

PAK subtype, PAK5, is highly expressed in brain, acts downstream of active Rac and

Cdc42, and leads to neurite outgrowth in a kinase activity-dependent manner (Dan et al.,

2002).

Molecular mechanisms of neurite outgrowth

Neurite (axon and dendrite) outgrowth defines neuronal shape, mediates neuronal pathfinding, and is essential for the establishment of synaptic connections during development, as well as being an important factor in neuronal regeneration following injury or neuropathological conditions. This complex process requires both co-ordinated cytoskeleton (actin and tubulin) remodelling and membrane expansion (mediated by vesicle targeting to the plasma membrane) in the growth cone.

Assembly and disassembly of polymeric actin filaments and polymerisation of tubulin into microtubules in the growth cone are essential for neurite extension. Several studies have demonstrated that polymerisation of actin occurs at the periphery of the growth cone, and depolymerisation occurs in the central domain, resulting in a retrograde flow of actin. Neurite elongation is driven by co-ordination of these actin forces with protrusion of microtubules into the central domain (Baas and Buster, 2004; Bradke and

Dotti, 1999; Li et al., 1994; Mallavarapu and Mitchison, 1999; Mitchison and

Kirschner, 1988; Suter and Forscher, 2000). Microtubules are transported into neurites and growth cones as short polymers, and live-cell imaging studies have shown that this transport is intermittent and infrequent, occurring at rates consistent with known proteins and in both anterograde and retrograde directions (Wang and

Brown, 2002). Several studies have recently highlighted the importance of molecular

163 motors, in particular cytoplasmic dynein and , in neurite outgrowth. It has been found that dynein transports microtubules from the centrosome into developing neurites

(Ahmad et al., 1998) as well as playing an important role in the retrograde transport of membranous vesicles, many of which contain growth factors (Karki and Holzbaur,

1999). have also been implicated in neurite outgrowth (Sharp et al., 1997).

Myosins have been found to fuel retrograde actin flow within growth cones and cause microtubules to move retrogradely down the neurite, leading to neurite retraction

(Ahmad et al., 2000; Schaefer et al., 2002), as well as transporting vesicular elements to the growing neurite tip (Evans and Bridgman, 1995). Therefore, it is believed that a balance between the forces mediated by dynein and kinesins on microtubules and actin- myosin contractility controls neurite outgrowth.

Rho family GTPases are key regulators of actin-myosin contractility and are thought to exert their role in neurite outgrowth by controlling the retrograde flow of actin within the growth cone. Downstream targets of Rho family GTPases have been identified which are involved in actin polymerisation (N-WASP, SCAR/WAVE and VASP/Ena), actin depolymerisation (PAK, ROCK and LIMK) and myosin activity (PAK, ROCK and MLCK). For a review see (Huber et al., 2003). Additionally, it has been suggested that Rac-PAK signalling may affect microtubule dynamics, as PAK phosphorylates and inhibits the microtubule stabilising protein Stathmin in response to Rac signalling (Daub et al., 2001), although the role of this signalling pathway in neurite outgrowth has not yet been investigated. The involvement of Rac in signalling to both actin and microtubules therefore raises the possibility that Rac is involved in co-ordination between the actin and microtubule arrays within the growth cone.

164 Evidence is emerging that neurite outgrowth is also dependent on exocytic addition of plasma membrane at the leading edge of the growth cone (Dai and Sheetz, 1995).

Inhibitors of vesicle fusion proteins such as tetanus neurotoxin-insensitive vesicle- associated membrane protein (TI-VAMP), SNAP-25 and Syntaxin inhibit neurite outgrowth, whereas overexpression of TI-VAMP or Syntaxin promotes neurite outgrowth (Grosse et al., 1999; Hirling et al., 2000; Igarashi et al., 1996; Martinez-Arca et al., 2001; Osen-Sand et al., 1996), which suggests that exocytosis is a rate-limiting step. The Rab family of GTPases are important regulators at various stages of membrane traffic, and antisense oligonucleotides of the Rab guanine nucleotide dissociation inhibitor (GDI) inhibit axonal outgrowth of hippocampal neurons

(D'Adamo et al., 1998), although so far the only Rab GTPase that has been implicated in neurite outgrowth is Rab8 (Huber et al., 1995).

Neurite outgrowth and MND

Several models of MND show impaired neurite outgrowth. For example, embryonic motor neurons isolated from pmn mutant mice show severely impaired axonal growth

(axonal length is reduced by more than 50% compared to controls) (Bommel et al.,

2002). As described in Section 1.1.3.1, these mice have a homozygous mutation in the

Tbce gene involved in tubulin assembly and develop a MND resembling spinal muscular atrophy (SMA) (Matsuo et al., 2002). Interestingly, the morphology of the growth cones of pmn mouse motor neurons appears normal, whilst axonal swellings are present with an irregular pattern of immunoreactivity for tubulin and tau (Jablonka et al., 2004). Additionally, knockdown of survival motor neuron protein (Smn), which is mutated in SMA, causes defects in motor axon outgrowth and pathfinding in zebrafish

(McWhorter et al., 2003).

165 Mice heterozygous for either of two ENU-generated mutations, Loa and Cra1, demonstrate a motor neuron degenerative phenotype, and positional cloning has revealed that both disease-causing mutations are in the cytoplasmic dynein heavy chain

1 gene (Dnchc1). Furthermore, mutations in dynactin (p150 subunit) have been found to cause an autosomal dominant form of lower motor neuron disease in humans (Puls et al., 2003). Aside from its role in retrograde transport of vesicles, dynein is also required for proper transport of microtubules into developing neurites (Ahmad et al., 1998).

Mutations in dynein may therefore cause motor neuron disease by perturbation of such functions. Studies of cultured spinal motor neurons from SOD1G93A mutant mice have also revealed neurite outgrowth defects in these cells (Azzouz et al., 2000). Similarly, motor neuron-neuroblastoma cells expressing SOD1G93A showed neurite outgrowth defects (Lee et al., 2002). Such defects were rescued by treatment with copper chelators

(Azzouz et al., 2000) and by expression of heat shock proteins HSP70 and HSP40 leading to reduced protein aggregation (Takeuchi et al., 2002b). Furthermore, the transport of tubulin is impaired and axonal levels of the motor protein kinesin are reduced in SOD1G85R mice months before disease onset (Warita et al., 1999; Zhang et al., 1997) which raises the possibility that microtubules may not be transported to the axonal tip during neuronal development in these mice.

Finally, mutations in L1CAM have been found to cause hereditary spastic paraplegia

(SPG1) (Jouet et al., 1994). L1CAM is an axonal glycoprotein involved in neuronal growth and guidance (Bixby et al., 1988; Castellani et al., 2000). Thus, aside from the findings described in this Chapter, a number of other studies have implicated defective neurite outgrowth in motor neuron disease.

166 Figure 5.1 Developmental expression of ALS2 in mouse brain and cultured rat embryonic cortical neurons (A) is an immunoblot to demonstrate developmental expression of ALS2 in mouse brain from embryonic day (E) 15 to post-natal (P) 1 year. (B) is a similar immunoblot of rat cortical neurons cultured for 2-21 DIV as indicated. The samples were also probed for tubulin to demonstrate equal protein loadings.

A E15 E18 P1 P7 P11 P21 1Yr

ALS2

tubulin

B 2 4 7 11 14 21 ALS2

tubulin

167 Figure 5.2 Subcellular localisation of ALS2 in cultured embryonic neurons Embryonic rat hippocampal neurons were cultured and at 2 DIV were subjected to immunofluorescence with ALS2 antibody (to detect endogenous ALS2) and Alexa Fluor 568-phalloidin (to detect filamentous (F)-actin). Pictures shown are confocal z- projections. ALS2 colocalises with F-actin (Overlay), and is present in both axonal and dendritic growth cones (arrows). Punctate ALS2 staining can be seen in the cell body and neurites. Scale bar is 10 µm.

168 Figure 5.3 Subcellular localisation of ALS2 in neuronal growth cones Endogenous ALS2 is co-localised with Rac (A) and F-actin (B) in the growth cone and with tubulin in the axon shaft (C) of 2 DIV rat embryonic hippocampal neurons. Overlays of co-staining are shown. A is a confocal slice; B and C are confocal z- projections. Scale bars are 5 µm.

A

B

C

169 Figure 5.4 Transfected ALS2 and ALS2ΔDH display identical subcellular localisation to endogenous ALS2 Embryonic rat cortical neurons were transfected with either myc-tagged ALS2 or ALS2ΔDH at 2 DIV and immunostained with the anti-myc antibody 9B11 24 hours later. Both of the transfected constructs localised to the cell body, processes and growth cones (arrows), which is an identical localisation to that of endogenous ALS2 (as seen in Figure 5.4). Scale bar is 10 µm.

170 Figure 5.5 ALS2 promotes neurite outgrowth in cultured rat cortical neurons 2 DIV rat cortical neurons were transfected with GFP+control or experimental plasmids and the length of the longest neurite measured using GFP as a marker. Transfections were balanced with pCIneo-CAT (CAT) so that all cells received the same numbers and amounts of plasmid. (A) shows histogram of mean neurite length for each transfection condition as indicated. Data was obtained from 40-50 cells per transfection and the experiments were repeated at least 3 times. ALS2 but not ALS2ΔDH stimulates neurite outgrowth by 1.5-fold (p=<0.001) and this effect is lost upon co-transfection with N17Rac but not N39Rab5. (B) shows representative images of cells in the different transfections: (a), GFP+CAT; (b), GFP+ALS2; (c), GFP+ALS2ΔDH; (d) GFP+ALS2+N17Rac; (e), GFP+ALS2+N39Rab5; (f), GFP+N17Rac; (g), GFP+N39Rab5. Scale bar is 10 µm.

A

B

171 Figure 5.6 Overexpression of ALS2 does not affect the number of neurites or the extent of neurite branching in cortical neurons (A) shows histogram of the mean number of primary neurites per cell for rat cortical neurons co-transfected with either GFP+pCIneo-CAT (CAT), GFP+ALS2 or GFP+ALS2ΔDH at 2 DIV. (B) shows histogram of the mean neurite tip number per cell, which is a measure of total neurite number and an indication of the extent of neurite branching, for the same transfection conditions as above. Data was obtained from 40-50 cells per transfection and the experiments were repeated three times to obtain mean values.

A

B

172

CHAPTER 6: ALS2 IS A PHOSPHOPROTEIN

173 6.1 Introduction

DH/PH (Dbl-family) GEFs activate Rho family GTPases (for example Rho, Rac and

Cdc42) by de-stabilising GDP interactions and stabilising the nucleotide-depleted

GTPase resulting in GTP binding (driven by GTP concentration being approximately

10-fold higher than GDP in the cell). GEFs are extremely important in the regulation and integration of Rho family GTPase signalling pathways and it is therefore not surprising that GEFs themselves are highly regulated. Several ways in which GEFs are positively regulated have been identified including activation by direct interaction with

G protein subunits. This activation has been found to occur in DH/PH GEFs that contain a regulator of G-protein signalling (RGS) domain, such as p115RhoGEF (Hart et al.,

1998). A second means by which GEFs are activated is by binding to phospholipids; for example the activities of Vav, Sos and Tiam1 are all enhanced by binding to

PtdIns(3,4,5)P3, which is thought to occur via the PH domain (Das et al., 2000; Fleming et al., 2000; Han et al., 1998). An additional and common form of positive regulation of the DH/PH GEFs is phosphorylation by protein kinases.

Many GEFs are phosphorylated and activated in vivo by tyrosine and/or serine- threonine protein kinases, for example Vav, p85 β-PIX and GEF-H1 (Das et al., 2000;

Shin et al., 2002; Zenke et al., 2004). Phosphorylation by protein kinases may elicit a variety of effects on the target protein and therefore could alter GEF activity, specificity, stability, association with other proteins or localisation within the cell. To begin to understand the upstream mechanisms that regulate ALS2 activity, studies were carried out to investigate whether ALS2 is phosphorylated in vivo.

174 6.2 Materials and Methods

6.2.1 Cell culture and transfection

CHO cells were cultured on 10 cm dishes and transfected using Lipofectamine™, as described in Section 2.2.6.3. For ALS2 immunoprecipitation, cells were transfected with 8 μg myc-tagged ALS2 cDNA. For Rac activation assays, cells were co- transfected with 2 µg of myc-tagged Rac1 and 8 µg of either ALS2, ALS2ala or pCIneo-CAT cDNA (see Table 2.2 for information about expression vectors used).

ALS2ala is a mutant ALS2 in which 5 identified serine and threonine phosphorylated residues were mutated to alanine to preclude phosphorylation. For PAK1 kinase assays, cells were co-transfected with 2 µg myc-tagged PAK1, 2 µg Rac1 and 6 µg of either pCIneo-CAT as a negative control, ALS2, ALS2ala or constitutively active Rac1

(L61Rac1) as a positive control.

Primary embryonic (E18) rat cortical cultures were prepared as described in Section

2.2.6.2 and cultured on PDL-coated 10 cm dishes. Cells were harvested for immunoprecipitation at 7 DIV in 1 ml of cell lysis buffer (see Section 2.1.5.1 and

2.2.8.1) and treated with λ protein phosphatase as described (Section 2.2.4.4). For neurite outgrowth assays, cells were transfected at 2 DIV with either ALS2, ALS2ala or pCIneo-CAT using Lipofectamine 2000™ and harvested at 3 DIV, essentially as described in Section 5.2.3.

6.2.2 Mass spectrometric sequencing of ALS2

ALS2 was immunoprecipitated from transfected CHO cells, as described in Section

2.2.8.1. Samples were subjected to SDS-PAGE and gels were stained with Colloidal

Brilliant Blue G (Sigma). The appropriate bands were excised from the gel and sequenced by mass spectrometry. Briefly, bands were reduced, alkylated and digested

175 with either trypsin, chymotrypsin or Asp-N (Roche Molecular Biochemical) and peptides extracted with two wash cycles of 50 mM NH4HCO3 and acetonitrile, and then lyophilised and resuspended in 20 μl of 50mM NH4HCO3. Peptide digests were analysed by on-line liquid chromatography tandem mass spectrometry (LC/MS/MS).

Peptides were ionised by electrospray ionisation using a Z-spray source fitted to a

QTof-micro (Micromass, UK). The instrument was set to run in automated switching mode, selecting precursor ions based on their intensity and charge state, for sequencing by collision-induced fragmentation. The MS/MS analyses were conducted using collision energy profiles that were chosen based on the m/z and the charge state of the peptide and optimized for phosphorylated peptides. The mass spectral data was processed into peak lists containing the mass/charge (m/z) value of each precursor ion and the corresponding fragment ion m/z values and intensities. Data was searched against a custom-built database containing the full-length sequence of ALS2 using the

Mascot searching algorithm (Matrix Science, UK). Peptides and phospho-peptides of

ALS2 were identified as described (Perkinton et al., 2004; Standen et al., 2003). Mass spectrometry studies were performed “in-house” by Helen Byers (see

Acknowledgements).

6.2.3 Rac activation assay

CHO cells were harvested in 1 ml MLB buffer 24 hours after transfection, following 16 hours of serum-starvation. Rac activation assays were carried out using GST-PBD, as described in Section 4.2.3. Immunoblotting revealed similar levels of ALS2ala and wild-type ALS2 in the cell lysates.

176 6.2.4 PAK1 kinase assay

CHO cells were harvested 24 hours after transfection, following 16 hours of serum- starvation and PAK1 in vitro kinase assays were carried out as described in Section

4.2.4.

6.2.5 Neurite outgrowth measurements

Neurite outgrowth measurements were carried out as described in Section 5.2.3.

Immunofluorescence revealed that the localisation of ALS2ala in cortical neurons was identical to that of wild-type ALS2.

6.3 Results

6.3.1 ALS2 is a phosphoprotein

To begin to understand the upstream mechanisms that regulate ALS2 activity, studies were carried out to investigate whether ALS2 is phosphorylated in vivo. Endogenous

ALS2 was immunoprecipitated from cultured embryonic rat cortical neurons at 7 DIV and samples were then incubated in buffer either with or without λ protein phosphatase

(Section 2.2.4.4). The samples were subjected to SDS-PAGE and immunoblotting with antibodies to ALS2 or to phospho-serine/threonine-proline (MPM-2). The ALS2 protein immunoprecipitated was found to be phosphorylated at serine/threonine-proline residues (Figure 6.1).

The ALS2 protein was then sequenced to identify phosphorylation sites. ALS2 is a particularly low abundance protein (Yamanaka et al., 2003) and so to obtain sufficient protein for sequencing, it was isolated from transfected CHO cells. Using a combination of trypsin, chymotrypsin and Asp-N protease digestion, 81 % sequence coverage was obtained. Serines 277, 492, 1335, 1464 and threonine 510 were all unambiguously

177 identified as phosphorylation sites but a number of other phosphopeptides were also detected although the responsible residues could not be identified (Figure 6.2). The identified residues all precede a proline making them candidates for phosphorylation by proline-directed kinases such as those of the MAP kinase superfamily, the cyclin- dependent kinases (Cdks) and glycogen synthase kinase-3α/β (GSK3α/β).

6.3.2 Mutation of serine/threonine-proline sites to alanine does not affect the activity of ALS2

To enquire whether phosphorylation of these sites influenced ALS2 regulation of Rac activity, mutants were constructed from myc-tagged ALS2 by Michael Perkinton

(Institute of Psychiatry) in which the identified serine/threonine residues were altered to alanine to preclude phosphorylation (ALS2ala). However, this mutant had no discernible effect on ALS2 Rac activity, PAK1 activity or neurite outgrowth (Figure

6.3).

6.4 Discussion

Several GEFs are known to be activated by phosphorylation. One way in which phosphorylation activates GEF activity is by inducing a change in conformation. For example, Vav is thought to be autoinhibited by binding of its N-terminal to the catalytic

DH domain and this intramolecular inhibition is relieved upon phosphorylation by Src family tyrosine kinases (Aghazadeh et al., 2000). Phosphorylation has also been shown to promote binding of proteins that recognise the phosphorylated epitope, for example

GEF-H1 is phosphorylated by the serine/threonine kinase PAK1, and this promotes binding to 14-3-3 adaptor protein (Zenke et al., 2004). Furthermore in some cases, for example Vav, phosphorylation is important in regulating GEF binding to its upstream regulators and/or adaptor proteins, thereby allowing further phosphorylation and

178 activation to occur (for a review see (Bustelo, 2000)). Therefore identification of ALS2 in vivo phosphorylation sites may provide important information on the upstream components of ALS2 signalling pathways.

Here, I have demonstrated that ALS2 is a phosphoprotein and report the identification of five cellular phosphorylation sites. All of these sites are ser/thr-pro motifs, which makes them targets for proline-directed kinases. Interestingly, aberrant activation of two proline directed kinases, p38 (a stress-activated kinase) and Cdk5/p35 (a neuronal cyclin-dependent kinase) have been described in human ALS cases and mutant SOD1 transgenic models of ALS (Ackerley et al., 2004; Hu et al., 2003a; Hu et al., 2003b;

Nguyen et al., 2001; Raoul et al., 2002; Tortarolo et al., 2003; Zhu et al., 2002).

However, mutation of the 5 ALS2 phosphorylation sites identified in the studies reported here did not noticeably alter the effect of ALS2 on Rac1 activity. Whether phosphorylation regulates the Rab5 GEF activity of ALS2 may be addressed in future studies and when the full complement of ALS2 phosphorylation sites have been identified.

179 Figure 6.1 ALS2 is phosphorylated on serine/threonine-proline residues in embryonic rat cortical neurons ALS2 was immunoprecipitated from rat cortical neurons (7 DIV) with affinity-purified ALS2 antibody and treated with λ protein phosphatase (λppase). Immunoprecipitations were carried out with (+) or without (-) the immunoprecipitating (IP) antibody and subjected to SDS-PAGE and immunoblotting with the phospho-serine/threonine-proline antibody MPM2 (Phospho SP/TP; upper panel). Blots were then re-probed for ALS2 to show the relative amounts of ALS2 in the immunoprecipitates (ALS2; lower panel).

IP antibody - + + λppase - - +

Phospho SP/TP

ALS2

180 Figure 6.2 Phosphorylation of ALS2 The amino acid sequence of ALS2 is shown; residues covered by mass spectrometry are shown in bold and identified phosphorylation sites are boxed. The Rac GEF domain including the pleckstrin homology region is underlined. Ser277 was covered by peptide DSHCCPLGVTLTEGQAENHASTALpSPSTETL; precursor ion [M+3H]3+, m/z=1131.82. Ser492 was covered by peptide RLSLPGLLSQVpSPR; precursor ion [M+2H]2+, m/z=841.89. Thr510 was covered by peptide TVVLpTPTYSGEADALLPSIR; precursor ion [M+2H]2+, m/z=1092.5. Ser1335 was covered by peptide QHRDpSPEILSR; precursor ion [M+3H]3+, m/z=472.2. Ser1464 was covered by peptide TGKSDSRSEpSPEPGYVVTSSGL; precursor ion [M+2H]2+, m/z=1160.61.

MDSKKRSSTE AEGSKERGLV HIWQAGSFPI TPERLPGWGG KTVLQAALGV KHGVLLTEDG EVYSFGTLLW RSGPVEICPS SPILENALVG QYVITVATGS FHSGAVTDNG VAYMWGENSA GQCAVANQQY VPEPNPVSIA DSEASPLLAV RILQLACGEE HTLALSISRE IWAWGTGCQL GLITTAFPVT KPQKVEHLAG RVVLQVACGA FHSLALVQCL PSQDLKPVPE RCNQCSQLLI TMTDKEDHVI ISDSHCCPLG VTLTESQAEN HASTALSPST ETLDRQEEVF ENTLVANDQS VATELNAVSA QITSSDAMSS QQNVMGTTEI SSARNIPSYP DTQAVNEYLR KLSDHSVRED SEHGEKPVPS QPLLEEAIPN LHSPPTTSTS ALNSLVVSCA SAVGVRVAAT YEAGALSLKK VMNFYSTTPC ETGAQAGSSA IGPEGLKDSR EEQVKQESMQ GKKSSSLVDI REEETEGGSR RLSLPGLLSQ VSPRLLRKAA RVKTRTVVLT PTYSGEADAL LPSLRTEVWT WGKGKEGQLG HGDVLPRLQP LCVKCLDGKE VIHLEAGGYH SL ALTAKSQV YSWGSNTFGQ LGHSDFPTTV PRLAKISSEN GVWSIAAGRD YSLFLVDTED FQPGLYYSGR QDPTEGDNLP ENHSGSKTPV LLSCSKLGYI SRVTAGKDSY LALVDKNIMG YIASLHELAT TERRFYSKLS DIKSQILRPL LSLENLGTTT TVQLLQEVAS RFSKLCYLIG QHGASLSSFL HGVKEARSLV ILKHSSLFLD SYTEYCTSIT NFLVMGGFQL LAKPAIDFLN KNQELLQDLS EVNDENTQLM EILNTLFFLP IRRLHNYAKV LLKLATCFEV ASPEYQKLQD SSSCYECLAL HLGRKRKEAE YTLGFWKTFP GKMTDSLRKP ERRLLCESSN RALSLQHAGR FSVNWFILFN DALVHAQFST HHVFPLATLW AEPLSEEAGG VNGLKITTPE EQFTLISSTP QEKTKWLRAI SQAVDQALRG MSDLPPYGSG SSVQRQEPPI SRSAKYTFYK DPRLKDATYD GRWLSGKPHG RGVLKWPDGK MYSGMFRNGL EDGYGEYRIP NKAMNKEDHY VGHWKEGKMC GQGVYSYASG EVFEGCFQDN MRHGHGLLRS GKLTSSSPSM FIGQWVMDKK AGYGVFDDIT RGEKYMGMWQ DDVCQGNGVV VTQFGLYYEG NFHLNKMMGN GVLLSEDDTI YEGEFSDDWT LSGKGTLTMP NGDYIEGYFS GEWGSGIKIT GTYFKPSLYE SDKDRPKVFR KLGNLAVPAD EKWKAVFDEC WRQLGCEGPG QGEVWKAWDN IAVALTTSRR QHRDSPEILS RSQTQTLESL EFIPQHVGAF SVEKYDDIRK YLIKACDTPL HPLGRLVETL VAVYRMTYVG VGANRRLLQE AVKEIKSYLK RIFQLVRFLF PELPEEGSTI PLSAPLPTER KSFCTGKSDS RSESPEPGYV VTSSGLLLPV LLPRLYPPLF MLYALDNDRE EDIYWECVLR LNKQPDIALL GFLGVQRKFW PATLSILGES KKVLPTTKDA CFASAVECLQ QISTTFTPSD KLKVIQQTFE EISQSVLASL HEDFLWSMDD LFPVFLYVVL RARIRNLGSE VHLIEDLMDP YLQHGEQGIM FTTLKACYYQ IQREKLN

181 Figure 6.3 Mutation of ALS2 phosphorylation sites does not affect ALS2 Rac GEF activity, PAK1 activation or neurite outgrowth Ser277, Ser492, Thr510, Ser1335 and Ser1464 in ALS2 were all mutated to alanine to preclude phosphorylation, generating the mutant “ALS2ala”. ALS2ala displayed the same functional properties as wild-type ALS2 in these studies: (A) shows Rac activation assays in CHO cells co-transfected with Rac1 and either pCIneo-CAT (CAT) as a negative control, ALS2 or ALS2ala. Active Rac was pulled down from lysates using GST-PBD and detected by immunoblotting (Active). The amounts of total Rac transfected into the cells were also detected by immunoblotting so as to demonstrate equal transfection efficiencies (Total). (B) shows PAK1 in vitro kinase assays that were performed from CHO cells co-transfected with PAK1, Rac1 and either CAT (as a negative control), ALS2 or ALS2ala. (-) and (+) refer to absence or presence of PAK1 immunoprecipitating antibody in the reactions; reaction mix (RM) contains no immunoprecipitation sample. “Total PAK1” is the relative amounts of PAK1 immunoprecipitated in the (+) samples. 2 DIV rat cortical neurons were transfected with either GFP+CAT (control condition), GFP+ALS2 or GFP+ALS2ala, and the length of the longest neurite was measured using GFP as a marker. (C) shows histogram of mean neurite length for each transfection condition as indicated. Data was obtained from 40- 50 cells per transfection and the experiments were repeated three times.

A B

C

182

CHAPTER 7: DISCUSSION & FUTURE

DIRECTIONS

183 7.1 Summary of findings

In the studies presented here, a polyclonal antibody to ALS2 was generated and used to investigate ALS2 localisation in the mammalian central nervous system. ALS2 was found to be present in a number of neuronal populations in the brain including neurons of the cortex, hippocampus and cerebellum, and in motor neurons of the spinal cord.

Since ALS2 contains a DH/PH domain, which is a hallmark of Rho family GEFs, assays were performed to investigate whether ALS2 activates the three most widely- studied Rho-family GTPases RhoA, Rac1 and Cdc42 in vivo. ALS2 was shown to activate Rac1 but not RhoA or Cdc42. Rac1 activation required the DH domain, as shown by use of an ALS2 mutant (ALS2ΔDH). Next, the activation of a major downstream effector of Rac1 (PAK1) by ALS2 was investigated. ALS2 was found to activate PAK1 in in vitro kinase assays. Rac1 is known to regulate neurite outgrowth during development. Experiments were therefore performed to investigate the role of

ALS2 in neurite outgrowth, and overexpression of ALS2 in rat embryonic cortical neurons was found to stimulate neurite outgrowth. Immunofluorescence showed that endogenous ALS2 is present in the cell body and processes, and is co-localised with F- actin and Rac in the growth cones, of rat embryonic cortical and hippocampal neurons, which further implicates a role for the protein in neurite outgrowth. Finally, the phosphorylation state of ALS2 was investigated, as the activity of GEFs is often regulated by means of phosphorylation. ALS2 was found to be a phosphoprotein in vivo and five serine/threonine phosphorylation sites were identified. Mutation of these sites to alanine did not affect the function of the ALS2 protein in Rac1 activation assays,

PAK1 kinase assays, or neurite outgrowth experiments. The findings reported in this thesis provide an insight into ALS2 function, which may contribute to our understanding of the molecular mechanisms involved in motor neuron disease.

184 7.2 The role of alsin/ALS2 in motor neuron disease

Nine disease-causing mutations in ALS2 have been described in nine different autosomal recessive kindreds. All of the affected individuals are homozygous for the mutation and develop a slowly progressive ascending upper motor neuron disorder which presents with a lower limb spasticity and can have onset in infancy, childhood or adolescence. All mutations result in premature translational termination and truncation of the full-length native protein (Devon et al., 2003; Eymard-Pierre et al., 2002; Gros-

Louis et al., 2003b; Hadano et al., 2001; Yang et al., 2001). The neuropathology has not been described but in a clinically similar genetic disorder, hereditary spastic paraplegia due to mutations in SPG4, upper motor neurons projecting in the corticospinal tract develop a dying-back axonopathy (Wharton et al., 2003). The recessive nature of ALS2 and truncation mutations suggest that the disorder is caused by a loss of normal ALS2 function, but the function of ALS2 is unclear and the precise mechanisms by which this leads to clinically selective motor neuron degeneration are unknown.

One possibility is that loss of ALS2 Rab5 GEF function perturbs membrane trafficking so as to induce disease. Indeed, disruptions to membrane trafficking and the Golgi apparatus are seen in mutant SOD1 transgenic mouse models of ALS and recently mutations in the vesicle-trafficking protein VAPB have been shown to cause late onset spinal muscular atrophy and ALS (Mourelatos et al., 1996; Nishimura et al., 2004).

There may even be mechanisms linking mutant SOD1 and ALS2 forms of ALS

(Kanekura et al., 2005; Kanekura et al., 2004).

Another possibility is that loss of ALS2 Rac1 GEF function compromises proper development of motor neurons making them more susceptible to later toxic insults.

185 Indeed, upper motor neurons are the largest in the central nervous system with the longest axons, and so any defect in axonal growth induced by loss of ALS2 function is likely to be most severe in these cells. Interestingly, Rho family GTPase and PAK family kinase signalling pathways have both been implicated in another nervous system disorder that is thought to result from disorganisation of neuronal network formation.

Mutations in Oligophrenin1 (which encodes p190 RhoGAP) and the Rac GEF α-PIX are both involved in X-linked mental retardation (MRX), which is classed as a developmental disorder, and is associated with an immature morphology of synaptic spines (Billuart et al., 1998; Xiao et al., 2003). MRX is also caused by point mutations in the brain-specific PAK isoform PAK3. One such mutation, R67C, lies in the conserved region reported to be critical for GTPase binding and PAK activation, which suggests that GTPase-dependent PAK activation is affected in the disease (Allen et al.,

1998; Bienvenu et al., 2000; Knaus et al., 1998). Furthermore, William's syndrome is a neurological condition characterized by mild mental retardation and defects in visuo- spatial cognition, and this syndrome results from deletions of the gene for LIM kinase.

LIM kinase acts downstream of both Rac-PAK and Rho-Rho kinase (ROCK/ROK) to phosphorylate Cofilin, regulating reorganization of the actin cytoskeleton within the growth cone (Frangiskakis et al., 1996; Kuhn et al., 2000; Maekawa et al., 1999).

Additionally, it is now known that endocytic trafficking and cytoskeletal dynamics are intimately linked and therefore it is possible that ALS2 may co-ordinate both Rac and

Rab5 activities to regulate actin remodelling in endocytic events in motor neurons

(Lanzetti et al., 2004; McPherson, 2002; Schafer, 2002). Such events may include endocytosis of AMPA receptors as this process is thought to involve actin reorganisation as well as endocytic trafficking (Zhou et al., 2001) and recently research was presented at the Society for Neuroscience 2004 meeting (Washington DC, USA)

186 stating that ALS2 interacts with GRIP1, an adaptor protein that specifically binds to the

C-terminals of AMPA receptor subtypes GluR2 and GluR3, via its RCC1 domain (Cai et al., 2004).

However, none of the above hypotheses are mutually exclusive. Whatever the mechanisms by which mutations in ALS2 induce disease, a proper understanding of

ALS2 function is likely to assist in unravelling the aberrant molecular processes by which motor neurons die in ALS, and therefore the work presented in this thesis provides a valuable contribution to this area of research.

7.3 Future directions

The results presented in this thesis provide novel information on the function of ALS2 and provide a starting point for several lines of study. Some suggestions for future research are given below.

1. All of the ALS2 disease-causing mutations described so far result in a predicted loss of function of the protein, as indicated by the recessive nature of inheritance and the finding that the mutant proteins are rapidly degraded when overexpressed in human cells (Yamanaka et al., 2003; Yang et al., 2001). Therefore, RNAi or knockout rodent models will be particularly useful tools in the analysis of the role of ALS2 in motor neuron disease. Although it has recently been reported that motor neurons of 20 month- old ALS2 knockout mice do not display any obvious morphological defects (Cai et al.,

2005), it would be interesting to investigate various aspects of neurogenesis in these cells, including growth cone morphology, neurite outgrowth and branching, and synapse formation.

187 2. The results in Chapter 5 show that overexpression of ALS2 enhances neurite outgrowth in a Rac1-dependent manner. It is possible that PAK1 is involved in this signalling pathway, especially as ALS2 activates PAK1 (as shown in Chapter 4). Co- expression of a dominant negative PAK1 construct with ALS2 in the neurite outgrowth experiments could be used to investigate this possibility further.

3. In Chapters 4 and 5 it was shown that ALS2 is a GEF for Rac1 and is present at all developmental stages in cultured cortical neurons, where it exhibits a punctate staining in the cell body, axon and dendrites. Furthermore, it is co-localised with Rac1 and F- actin in the neuronal growth cone, which raises the possibility that it is involved in regulation of actin assembly via activation of Rac1. Indeed as shown in Chapter 5, overexpression of ALS2 affects neurite outgrowth, and this is likely to be mediated by actin reorganisation. Dendritic spine formation and stability are also largely determined by the actin cytoskeleton, and numerous studies have implicated Rac1 in promotion of these events (Luo et al., 1996; Nakayama et al., 2000). Several Rac GEFs are enriched in spines, and their overexpression often results in an increase in the number and size of spines, for example Kalirin-7 (Penzes et al., 2001), GEFT (Bryan et al., 2004) and α-

PIX (Zhang et al., 2005). In contrast, the Rac GEF still life (SIF) localizes in the presynaptic periactive zone of the Drosophila neuromuscular junction; this suggests that the SIF-Rac pathway is involved in synaptic development (Sone et al., 2000).

Differential fractionation studies performed on extracts from rat cerebellum have revealed that a proportion of endogenous ALS2 protein is localised to the fraction containing synaptosomal membranes (Topp et al., 2004) and therefore it would be interesting to investigate whether ALS2 is present in dendritic spines or in the presynaptic terminal, and whether it functions to regulate actin-based events such as receptor endocytosis or dendritic spine morphogenesis.

188 4. The results in Chapter 6 show that ALS2 is a phosphoprotein in vivo, and five serine/threonine-proline phosphorylation sites were identified by mass spectrometry.

Although it was shown that mutation of these sites to alanine (to preclude phosphorylation) did not affect the ability of ALS2 to activate Rac1 or PAK1, or to enhance neurite outgrowth, there is still potential for these sites to be important in the regulation of ALS2 function. For example, phosphorylation of these sites may affect

ALS2 Rab5 GEF function or other cellular events mediated by its Rac GEF activity such as its recently discovered involvement in an anti-apoptotic pathway (Kanekura et al., 2005). These sites are phosphorylated in ALS2 expressed in CHO cells, however it would be interesting to identify further sites (or to confirm phosphorylation of these sites) in ALS2 obtained from either rodent brain or cultured neuronal cells.

Furthermore, the activity of many GEFs is regulated by tyrosine phosphorylation.

Interestingly, tyrosine phosphorylation activates several GEFs that are involved in growth cone motility and neurite outgrowth, for example Trio, Ephexin1 and p85 β-PIX

(Lanier and Gertler, 2000; Sahin et al., 2005; Shin et al., 2004). Although no phosphorylated tyrosines have as yet been identified in ALS2, the full-length protein contains 57 tyrosine residues. Therefore, it would be interesting to discover whether

ALS2 is also activated by tyrosine phosphorylation in vivo. Initial studies to investigate this possibility could include probing of immunoprecipitated ALS2 on immunoblots with antibodies that detect phosphotyrosine residues.

189 REFERENCES

Abalkhail, H., J. Mitchell, J. Habgood, R. Orrell, and J. De Belleroche. 2003. A New Familial Amyotrophic Lateral Sclerosis Locus on Chromosome 16q12.1-16q12.2. Am. J. Hum. Genet. 73:383-389.

Abe, K., L.H. Pan, M. Watanabe, T. Kato, and Y. Itoyama. 1995. Induction of nitrotyrosine-like immunoreactivity in the lower motor neuron of amyotrophic lateral sclerosis. Neurosci Lett. 199:152-154.

Ackerley, S., A.J. Grierson, S. Banner, M.S. Perkinton, J. Brownlees, H.L. Byers, M. Ward, P. Thornhill, K. Hussain, J.S. Waby, B.H. Anderton, J.D. Cooper, C. Dingwall, P.N. Leigh, C.E. Shaw, and C.C. Miller. 2004. p38alpha stress-activated protein kinase phosphorylates neurofilaments and is associated with neurofilament pathology in amyotrophic lateral sclerosis. Mol Cell Neurosci. 26:354-364.

Ackerley, S., A.J. Grierson, J. Brownlees, P. Thornhill, B.H. Anderton, P.N. Leigh, C.E. Shaw, and C.C.J. Miller. 2000. Glutamate slows axonal transport of neurofilaments in transfected neurons. J. Cell Biol. 150:165-175.

Aghazadeh, B., W.E. Lowry, X.Y. Huang, and M.K. Rosen. 2000. Structural basis for relief of autoinhibition of the Dbl homology domain of proto-oncogene Vav by tyrosine phosphorylation. Cell. 102:625-633.

Aghazadeh, B., K. Zhu, T.J. Kubiseski, G.A. Liu, T. Pawson, Y. Zheng, and M.K. Rosen. 1998. Structure and mutagenesis of the Dbl homology domain. Nat Struct Biol. 5:1098-1107.

Ahlijanian, M.K., N.X. Barrezueta, R.D. Williams, A. Jakowski, K.P. Kowsz, S. McCarthy, T. Coskran, A. Carlo, P.A. Seymour, J.E. Burkhardt, R.B. Nelson, and J.D. McNeish. 2000. Hyperphosphorylated tau and neurofilament and cytoskeletal disruptions in mice overexpressing human p25, an activator of cdk5. Proc. Natl. Acad. Sci. USA. 87:2910-2915.

190 Ahmad, F.J., C.J. Echeverri, R.B. Vallee, and P.W. Baas. 1998. Cytoplasmic dynein and dynactin are required for the transport of microtubules into the axon. J Cell Biol. 140:391-401.

Ahmad, F.J., J. Hughey, T. Wittmann, A. Hyman, M. Greaser, and P.W. Baas. 2000. Motor proteins regulate force interactions between microtubules and microfilaments in the axon. Nat Cell Biol. 2:276-280.

Albertinazzi, C., D. Gilardelli, S. Paris, R. Longhi, and I. de Curtis. 1998. Overexpression of a neural-specific rho family GTPase, cRac1B, selectively induces enhanced neuritogenesis and neurite branching in primary neurons. J Cell Biol. 142:815-825.

Alberts, A.S., H. Qin, H.S. Carr, and J.A. Frost. 2005. PAK1 negatively regulates the activity of the Rho exchange factor NET1. J Biol Chem. 280:12152-12161.

Al-Chalabi, A., P.M. Anderson, P. Nilsson, B. Chioza, J.L. Andersson, C.R. Russ, C.E. Shaw, J.F. Powell, and P.N. Leigh. 1999. Deletions of the heavy neurofilament subunit tail in amyotrophic lateral sclerosis. Hum. Mol. Genet. 8:157-164.

Alexianu, M.E., B.K. Ho, A.H. Mohamed, V. La Bella, R.G. Smith, and S.H. Appel. 1994. The role of calcium-binding proteins in selective motoneuron vulnerability in amyotrophic lateral sclerosis. Ann Neurol. 36:846-858.

Alexianu, M.E., M. Kozovska, and S.H. Appel. 2001. Immune reactivity in a mouse model of familial ALS correlates with disease progression. Neurology. 57:1282-1289.

Allen, K.M., J.G. Gleeson, S. Bagrodia, M.W. Partington, J.C. MacMillan, R.A. Cerione, J.C. Mulley, and C.A. Walsh. 1998. PAK3 mutation in nonsyndromic X-linked mental retardation. Nat Genet. 20:25-30.

Allen, M.J., X. Shan, and R.K. Murphey. 2000. A role for Drosophila Drac1 in neurite outgrowth and synaptogenesis in the giant fiber system. Mol Cell Neurosci. 16:754-765.

191 Almer, G., C. Guegan, P. Teismann, A. Naini, G. Rosoklija, A.P. Hays, C. Chen, and S. Przedborski. 2001. Increased expression of the pro-inflammatory enzyme cyclooxygenase-2 in amyotrophic lateral sclerosis. Ann Neurol. 49:176-185.

Andrus, P.K., T.J. Fleck, M.E. Gurney, and E.D. Hall. 1998. Protein oxidative damage in a transgenic mouse model of familial amyotrophic lateral sclerosis. J.Neurochem. 71:2041-2048.

Aoki, K., T. Nakamura, K. Fujikawa, and M. Matsuda. 2005. Local phosphatidylinositol 3,4,5-trisphosphate accumulation recruits Vav2 and Vav3 to activate Rac1/Cdc42 and initiate neurite outgrowth in nerve growth factor-stimulated PC12 cells. Mol Biol Cell. 16:2207-2217.

Aoki, M., C.L.G. Lin, J.D. Rothstein, B.A. Geller, B.A. Hosler, T.L. Munsat, H.R. Horvitz, and R.H. Brown, Jr. 1998. Mutations in the glutamate transporter EAAT2 gene do not cause abnormal EAAT2 transcripts in amyotrophic lateral sclerosis. Ann.Neurol. 43:645-653.

Arakawa, Y., H. Bito, T. Furuyashiki, T. Tsuji, S. Takemoto-Kimura, K. Kimura, K. Nozaki, N. Hashimoto, and S. Narumiya. 2003. Control of axon elongation via an SDF- 1alpha/Rho/mDia pathway in cultured cerebellar granule neurons. J Cell Biol. 161:381- 391.

Arthur, W.T., and K. Burridge. 2001. RhoA inactivation by p190RhoGAP regulates cell spreading and migration by promoting membrane protrusion and polarity. Mol Biol Cell. 12:2711-2720.

Ashley-Koch, A., E.R. Bonner, P.C. Gaskell, S.G. West, R. Tim, C.M. Wolpert, R. Jones, C.D. Farrell, M. Nance, I.K. Svenson, D.A. Marchuk, R.M. Boustany, J.M. Vance, W.K. Scott, and M.A. Pericak-Vance. 2001. Fine mapping and genetic heterogeneity in the pure form of autosomal dominant familial spastic paraplegia. Neurogenetics. 3:91-97.

192 Aspenstrom, P. 1999. Effectors for the Rho GTPases. Curr Opin Cell Biol. 11:95-102.

Azzouz, M., P. Poindron, S. Guettier, N. Leclerc, C. Andres, J.M. Warter, and J. Borg. 2000. Prevention of mutant SOD1 motoneuron degeneration by copper chelators in vitro. J.Neurobiol. 42:49-55.

Baas, P.W., and D.W. Buster. 2004. Slow axonal transport and the genesis of neuronal morphology. J. Neurobiol. 58:3-17.

Bagrodia, S., B. Derijard, R.J. Davis, and R.A. Cerione. 1995. Cdc42 and PAK- mediated signaling leads to Jun kinase and p38 mitogen-activated protein kinase activation. J Biol Chem. 270:27995-27998.

Bajaj, N.P.S., S.-T. Al-Sarraj, V. Anderson, M. Kibble, N. Leigh, and C.C.J. Miller. 1998. Cyclin-dependent kinase-5 is associated with lipofuscin in motor neurones in amyotrophic lateral sclerosis. Neurosci. Lett. 245:45-48.

Bakowska, J.C., Jenkins, R., Pendleton, J., and C. Blackstone. 2005. The Troyer syndrome (SPG20) protein spartin interacts with Eps15. Biochem Biophys Res Commun. 334:1042-1048

Barbeito, L.H., M. Pehar, P. Cassina, M.R. Vargas, H. Peluffo, L. Viera, A.G. Estevez, and J.S. Beckman. 2004. A role for astrocytes in motor neuron loss in amyotrophic lateral sclerosis. Brain Res Brain Res Rev. 47:263-274.

Bateman, A., E. Birney, L. Cerruti, R. Durbin, L. Etwiller, S.R. Eddy, S. Griffiths- Jones, K.L. Howe, M. Marshall, and E.L. Sonnhammer. 2002. The Pfam protein families database. Nucleic Acids Res. 30:276-80.

Beal, M.F., R.J. Ferrante, S.E. Browne, R.T. Matthews, N.W. Kowall, and R.H. Brown, Jr. 1997. Increased 3-nitrotyrosine in both sporadic and familial amyotrophic lateral sclerosis. Ann.Neurol. 42:644-654.

193 Beaulieu, J.M., and J.P. Julien. 2003. Peripherin-mediated death of motor neurons rescued by overexpression of neurofilament NF-H proteins. J. Neurochem. 85:248-256.

Beaulieu, J.M., M.D. Nguyen, and J.P. Julien. 1999. Late onset death of motor neurons in mice overexpressing wild-type peripherin. J.Cell Biol. 147:531-544.

Beckman, J.S., J. Chen, J.P. Crow, and Y.Z. Ye. 1994. Reactions of nitric oxide, superoxide and peroxynitrite with superoxide dismutase in neurodegeneration. Prog Brain Res. 103:371-380.

Beers, D.R., B.K. Ho, L. Siklos, M.E. Alexianu, D.R. Mosier, A.H. Mohamed, Y. Otsuka, M.E. Kozovska, R.E. McAlhany, R.G. Smith, and S.H. Appel. 2001. Parvalbumin overexpression alters immune-mediated increases in intracellular calcium, and delays disease onset in a transgenic model of familial amyotrophic lateral sclerosis. J Neurochem. 79:499-509.

Bellanger, J.M., J.B. Lazaro, S. Diriong, A. Fernandez, N. Lamb, and A. Debant. 1998. The two guanine nucleotide exchange factor domains of Trio link the Rac1 and the RhoA pathways in vivo. Oncogene. 16:147-152.

Ben Hamida, M., F. Hentati, and C. Ben Hamida. 1990. Hereditary motor system diseases (chronic juvenile amyotrophic lateral sclerosis). Conditions combining a bilateral pyramidal syndrome with limb and bulbar amyotrophy. Brain. 113 ( Pt 2):347- 363.

Benard, V., B.P. Bohl, and G.M. Bokoch. 1999. Characterization of and activation in chemoattractant-stimulated human neutrophils using a novel assay for active GTPases. J Biol Chem. 274:13198-13204.

Bendotti, C., N. Calvaresi, L. Chiveri, A. Prelle, M. Moggio, M. Braga, V. Silani, and S. De Biasi. 2001. Early vacuolization and mitochondrial damage in motor neurons of FALS mice are not associated with apoptosis or with changes in cytochrome oxidase histochemical reactivity. J Neurol Sci. 191:25-33.

194 Bensimon, G., L. Lacomblez, and V. Meininger. 1994. A controlled trial of riluzole in amyotrophic lateral sclerosis. ALS/Riluzole Study Group. N Engl J Med. 330:585-591.

Beretta, S., G. Sala, L. Mattavelli, C. Ceresa, A. Casciati, A. Ferri, M.T. Carri, and C. Ferrarese. 2003. Mitochondrial dysfunction due to mutant copper/zinc superoxide dismutase associated with amyotrophic lateral sclerosis is reversed by N-acetylcysteine. Neurobiol. Dis. 13:213-221.

Bergeron, C., K. Beric-Maskarel, S. Muntasser, L. Weyer, M.J. Somerville, and M.E. Percy. 1994. Neurofilament light and polyadenylated mRNA levels are decreased in amyotrophic lateral sclerosis motor neurons. J Neuropathol Exp Neurol. 53:221-230.

Biancalana, V., F. Serville, J. Pommier, J. Julien, A. Hanauer, and J.L. Mandel. 1992. Moderate instability of the trinucleotide repeat in spinobulbar muscular atrophy. Hum.Mol.Genet. 1:255-258.

Bibb, J.A., J.S. Chen, J.R. Taylor, P. Svenningsson, A. Nishi, G.L. Snyder, Z. Yan, Z.K. Sagawa, C.C. Ouimet, A.C. Nairn, E.J. Nestler, and P. Greengard. 2001. Effects of chronic exposure to cocaine are regulated by the neuronal protein Cdk5. Nature. 410:376-380.

Bienvenu, T., V. des Portes, N. McDonell, A. Carrie, R. Zemni, P. Couvert, H.H. Ropers, C. Moraine, H. van Bokhoven, J.P. Fryns, K. Allen, C.A. Walsh, J. Boue, A. Kahn, J. Chelly, and C. Beldjord. 2000. Missense mutation in PAK3, R67C, causes X- linked nonspecific mental retardation. Am J Med Genet. 93:294-298.

Billuart, P., T. Bienvenu, N. Ronce, V. des Portes, M.C. Vinet, R. Zemni, H. Roest Crollius, A. Carrie, F. Fauchereau, M. Cherry, S. Briault, B. Hamel, J.P. Fryns, C. Beldjord, A. Kahn, C. Moraine, and J. Chelly. 1998. Oligophrenin-1 encodes a rhoGAP protein involved in X-linked mental retardation. Nature. 392:923-926.

Bischoff, F.R., and H. Ponstingl. 1991. Catalysis of guanine nucleotide exchange on Ran by the mitotic regulator RCC1. Nature. 354:80-82.

195 Bixby, J.L., J. Lilien, and L.F. Reichardt. 1988. Identification of the major proteins that promote neuronal process outgrowth on Schwann cells in vitro. J Cell Biol. 107:353- 361.

Blair, I.P., C.L. Bennett, A. Abel, B.A. Rabin, J.W. Griffin, K.H. Fischbeck, D.R. Cornblath, and P.F. Chance. 2000. A gene for autosomal dominant juvenile amyotrophic lateral sclerosis (ALS4) localizes to a 500-kb interval on chromosome 9q34. Neurogenetics. 3:1-6.

Bock, J.B., H.T. Matern, A.A. Peden, and R.H. Scheller. 2001. A genomic perspective on membrane compartment organization. Nature. 409:839-841.

Bogdanov, M., R.H. Brown, Jr., W. Matson, R. Smart, D. Hayden, H. O'Donnell, M.F. Beal, and M. Cudkowicz. 2000. Increased oxidative damage to DNA in ALS patients. Free Radic.Biol.Med. 29:652-658.

Bokoch, G.M. 2003. Biology of the p21-activated kinases. Annu Rev Biochem. 72:743- 781.

Bokoch, G.M., and B.A. Diebold. 2002. Current molecular models for NADPH oxidase regulation by Rac GTPase. Blood. 100:2692-2696.

Bommel, H., G. Xie, W. Rossoll, S. Wiese, S. Jablonka, T. Boehm, and M. Sendtner. 2002. Missense mutation in the tubulin-specific chaperone E (Tbce) gene in the mouse mutant progressive motor neuronopathy, a model of human motoneuron disease. J. Cell Biol. 159:563-569.

Bonneau, D., J.M. Rozet, C. Bulteau, M. Berthier, R. Mettey, R. Gil, A. Munnich, and M. Le Merrer. 1993. X linked spastic paraplegia (SPG2): clinical heterogeneity at a single gene locus. J Med Genet. 30:381-384.

Borchelt, D.R., M.K. Lee, H.S. Slunt, M. Guarnieri, Z.-S. Xu, P.C. Wong, R.H. Brown Jr, D.L. Price, S.S. Sisodia, and D.L. Cleveland. 1994. Superoxide dismutase 1 with

196 mutations linked to familial amyotrophic lateral sclerosis possesses significant activity Proc.Natl.Acad.Sci.USA. 91:8292-8296.

Borchelt, D.R., P.C. Wong, M.W. Becher, C.A. Pardo, M.K. Lee, Z.S. Xu, G. Thinakaran, N.A. Jenkins, N.G. Copeland, S.S. Sisodia, D.W. Cleveland, D.L. Price, and P.N. Hoffman. 1998. Axonal transport of mutant superoxide dismutase 1 and focal axonal abnormalities in the proximal axons of transgenic mice. Neurobiol. Dis. 5:27-35.

Bourne, H.R., D.A. Sanders, and F. McCormick. 1990. The GTPase superfamily: a conserved switch for diverse cell functions. Nature. 348:125-132.

Bourne, H.R., D.A. Sanders, and F. McCormick. 1991. The GTPase superfamily: conserved structure and molecular mechanism. Nature. 349:117-127.

Bowling, A.C., J.B. Schulz, R.H. Brown, Jr., and M.F. Beal. 1993. Superoxide dismutase activity, oxidative damage, and mitochondrial energy metabolism in familial and sporadic amyotrophic lateral sclerosis. J Neurochem. 61:2322-2325.

Bradford, M.M. 1976. A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem. 72:248-254.

Bradke, F., and C.G. Dotti. 1999. The role of local actin instability in axon formation. Science. 283:1931-1934.

Braga, V.M. 2002. Cell-cell adhesion and signalling. Curr Opin Cell Biol. 14:546-556.

Bristol, L.A., and J.D. Rothstein. 1996. Glutamate transporter gene expression in amyotrophic lateral sclerosis motor cortex. Ann Neurol. 39:676-679.

Brown, S.D., and R. Balling. 2001. Systematic approaches to mouse mutagenesis. Curr Opin Genet Dev. 11:268-273.

197 Brownlees, J., A. Yates, N.P. Bajaj, D. Davis, B.H. Anderton, P.N. Leigh, C.E. Shaw, and C.C.J. Miller. 2000. Phosphorylation of neurofilament heavy chain side-arms by stress activated protein kinase-1b/Jun N-terminal kinase-3. J. Cell Sci. 113:401-407.

Bruening, W., J. Roy, B. Giasson, D.A. Figlewicz, W.E. Mushynski, and H.D. Durham. 1999. Up-regulation of protein chaperones preserves viability of cells expressing toxic Cu/Zn-superoxide dismutase mutants associated with amyotrophic lateral sclerosis. J.Neurochem. 72:693-699.

Brugnera, E., L. Haney, C. Grimsley, M. Lu, S.F. Walk, A.C. Tosello-Trampont, I.G. Macara, H. Madhani, G.R. Fink, and K.S. Ravichandran. 2002. Unconventional Rac- GEF activity is mediated through the Dock180-ELMO complex. Nat Cell Biol. 4:574- 582.

Bruijn, L.I., M.F. Beal, M.W. Becher, J.B. Schulz, P.C. Wong, D.L. Price, and D.W. Cleveland. 1997a. Elevated free nitrotyrosine levels, but not protein-bound nitrotyrosine or hydroxyl radicals, throughout amyotrophic lateral sclerosis (ALS)- like disease implicate tyrosine nitration as an aberrant in vivo property of one familial ALS-linked superoxide dismutase 1 mutant. Proc Natl Acad Sci U S A. 94:7606-7611.

Bruijn, L.I., M.W. Becher, M.K. Lee, K.L. Anderson, N.A. Jenkins, N.G. Copeland, S.S. Sisodia, J.D. Rothstein, D.R. Borchelt, D.L. Price, and D.W. Cleveland. 1997b. ALS-linked SOD1 mutant G85R mediates damage to astrocytes and promotes rapidly progressive disease with SOD1-containing inclusions. Neuron. 18:327-338.

Bruijn, L.I., M.K. Houseweart, S. Kato, K.L. Anderson, S.D. Anderson, E. Ohama, A.G. Reaume, R.W. Scott, and D.W. Cleveland. 1998. Aggregation and motor neuron toxicity of an ALS-linked SOD1 mutant independent from wild-type SOD1. Science. 281:1851-1854.

Brunialti, A.L., C. Poirier, H. Schmalbruch, and J.L. Guenet. 1995. The mouse mutation progressive motor neuronopathy (pmn) maps to chromosome 13. Genomics. 29:131- 135.

198 Bryan, B., V. Kumar, L.J. Stafford, Y. Cai, G. Wu, and M. Liu. 2004. GEFT, a Rho family guanine nucleotide exchange factor, regulates neurite outgrowth and dendritic spine formation. J Biol Chem. 279:45824-45832.

Buchsbaum, R., J.B. Telliez, S. Goonesekera, and L.A. Feig. 1996. The N-terminal pleckstrin, coiled-coil, and IQ domains of the exchange factor Ras-GRF act cooperatively to facilitate activation by calcium. Mol Cell Biol. 16:4888-4896.

Burbelo, P.D., D. Drechsel, and A. Hall. 1995. A conserved binding motif defines numerous candidate target proteins for both Cdc42 and Rac GTPases. J Biol Chem. 270:29071-29074.

Burridge, K., and K. Wennerberg. 2004. Rho and Rac take center stage. Cell. 116:167- 179.

Bustelo, X.R. 2000. Regulatory and signaling properties of the Vav family. Mol Cell Biol. 20:1461-1477.

Cai, H., H. Chiang, C. Xie, H. Shim, H. Wen, C.F. Coccia, M.R. Cookson, J. Rothstein, D.L. Price, and P.C. Wong. 2004. Characterisation of ALS2 knockout mice: the role of ALS2 in AMPA-mediated excitotoxicity. In Society for Neuroscience, Washington DC.

Cai, H., X. Lin, C. Xie, F.M. Laird, C. Lai, H. Wen, H.C. Chiang, H. Shim, M.H. Farah, A. Hoke, D.L. Price, and P.C. Wong. 2005. Loss of ALS2 function is insufficient to trigger motor neuron degeneration in knock-out mice but predisposes neurons to oxidative stress. J Neurosci. 25:7567-7574.

Caron, E. 2003. Rac signalling: a radical view. Nat Cell Biol. 5:185-187.

Carpenter, S. 1968. Proximal axonal enlargement in motor neuron disease. Neurology. 18:841-851.

Casari, G., M. De Fusco, S. Ciarmatori, M. Zeviani, M. Mora, P. Fernandez, G. De

199 Michele, A. Filla, S. Cocozza, R. Marconi, A. Durr, B. Fontaine, and A. Ballabio. 1998. Spastic paraplegia and OXPHOS impairment caused by mutations in paraplegin, a nuclear-encoded mitochondrial metalloprotease. Cell. 93:973-983.

Cassina, P., H. Peluffo, M. Pehar, L. Martinez-Palma, A. Ressia, J.S. Beckman, A.G. Estevez, and L. Barbeito. 2002. Peroxynitrite triggers a phenotypic transformation in spinal cord astrocytes that induces motor neuron apoptosis. J Neurosci Res. 67:21-29.

Castellani, V., A. Chedotal, M. Schachner, C. Faivre-Sarrailh, and G. Rougon. 2000. Analysis of the L1-deficient mouse phenotype reveals cross-talk between Sema3A and L1 signaling pathways in axonal guidance. Neuron. 27:237-249.

Cha, C.I., J.M. Kim, D.H. Shin, Y.S. Kim, J. Kim, M.E. Gurney, and K.W. Lee. 1998. Reactive astrocytes express nitric oxide synthase in the spinal cord of transgenic mice expressing a human Cu/Zn SOD mutation. Neuroreport. 9:1503-1506.

Chakrabarti, K., R. Lin, N.I. Schiller, Y. Wang, D. Koubi, Y.X. Fan, B.B. Rudkin, G.R. Johnson, and M.R. Schiller. 2005. Critical role for Kalirin in nerve growth factor signaling through TrkA. Mol Cell Biol. 25:5106-5118.

Charvin, D., C. Cifuentes-Diaz, N. Fonknechten, V. Joshi, J. Hazan, J. Melki, and S. Betuing. 2003. Mutations of SPG4 are responsible for a loss of function of spastin, an abundant neuronal protein localized in the nucleus. Hum Mol Genet. 12:71-78.

Chen, S., N. Mantei, L. Dong, and M. Schachner. 1999. Prevention of neuronal cell death by neural adhesion molecules L1 and CHL1. J Neurobiol. 38:428-439.

Chen, Y., D. Yee, K. Dains, A. Chatterjee, J. Cavalcoli, E. Schneider, J. Om, R.P. Woychik, and T. Magnuson. 2000. Genotype-based screen for ENU-induced mutations in mouse embryonic stem cells. Nat Genet. 24:314-317.

Chen, Y.Z., C.L. Bennett, H.M. Huynh, I.P. Blair, I. Puls, J. Irobi, I. Dierick, A. Abel, M.L. Kennerson, B.A. Rabin, G.A. Nicholson, M. Auer-Grumbach, K. Wagner, P. De

200 Jonghe, J.W. Griffin, K.H. Fischbeck, V. Timmerman, D.R. Cornblath, and P.F. Chance. 2004. DNA/RNA Gene Mutations in a Form of Juvenile Amyotrophic Lateral Sclerosis (ALS4). Am J Hum Genet. 74:1128-1135.

Chew, T.L., R.A. Masaracchia, Z.M. Goeckeler, and R.B. Wysolmerski. 1998. Phosphorylation of non-muscle myosin II regulatory light chain by p21-activated kinase (gamma-PAK). J Muscle Res Cell Motil. 19:839-854.

Ching, G.Y., C.L. Chien, R. Flores, and R.K.H. Liem. 1999. Overexpression of a- internexin causes abnormal neurofilamentous accumulations and motor coordination deficits in transgenic mice. J.Neurosci. 19:2974-2986.

Christoforidis, S., H.M. McBride, R.D. Burgoyne, and M. Zerial. 1999a. The Rab5 effector EEA1 is a core component of endosome docking. Nature. 397:621-625.

Christoforidis, S., M. Miaczynska, K. Ashman, M. Wilm, L. Zhao, S.C. Yip, M.D. Waterfield, J.M. Backer, and M. Zerial. 1999b. Phosphatidylinositol-3-OH kinases are Rab5 effectors. Nat Cell Biol. 1:249-252.

Clement, A.M., M.D. Nguyen, E.A. Roberts, M.L. Garcia, S. Boillee, M. Rule, A.P. McMahon, W. Doucette, D. Siwek, R.J. Ferrante, R.H. Brown, Jr., J.P. Julien, L.S. Goldstein, and D.W. Cleveland. 2003. Wild-type nonneuronal cells extend survival of SOD1 mutant motor neurons in ALS mice. Science. 302:113-117.

Cleveland, D.W., and J.D. Rothstein. 2001. From Charcot to Lou Gehrig: deciphering selective motor neuron death in ALS. Nat. Rev. Neurosci. 2:806-819.

Collard, J.-F., F. Cote, and J.-P. Julien. 1995. Defective axonal transport in a transgenic mouse model of amyotrophic lateral sclerosis. Nature. 375:61-64.

Collins, L.R., A. Minden, M. Karin, and J.H. Brown. 1996. Galpha12 stimulates c-Jun NH2-terminal kinase through the small G proteins Ras and Rac. J Biol Chem. 271:17349-17353.

201 Comi, G.P., A. Bordoni, S. Salani, L. Franceschina, M. Sciacco, A. Prelle, F. Fortunato, M. Zeviani, L. Napoli, N. Bresolin, M. Moggio, C.D. Ausenda, J.W. Taanman, and G. Scarlato. 1998. Cytochrome c oxidase subunit I microdeletion in a patient with motor neuron disease. Ann Neurol. 43:110-116.

Cookson, M.R., and P.J. Shaw. 1999. Oxidative stress and motor neurone disease. Brain Pathol. 9:165-186.

Corbo, M., and A.P. Hays. 1992. Peripherin and neurofilament protein coexist in spinal spheroids of motor neuron disease. J Neuropathol Exp Neurol. 51:531-537.

Corona, J.C., and R. Tapia. 2004. AMPA receptor activation, but not the accumulation of endogenous extracellular glutamate, induces paralysis and motor neuron death in rat spinal cord in vivo. J. Neurochem. 89:988-997.

Coso, O.A., M. Chiariello, J.C. Yu, H. Teramoto, P. Crespo, N. Xu, T. Miki, and J.S. Gutkind. 1995. The small GTP-binding proteins Rac1 and Cdc42 regulate the activity of the JNK/SAPK signaling pathway. Cell. 81:1137-1146.

Coso, O.A., H. Teramoto, W.F. Simonds, and J.S. Gutkind. 1996. Signaling from G protein-coupled receptors to c-Jun kinase involves beta gamma subunits of heterotrimeric G proteins acting on a Ras and Rac1-dependent pathway. J Biol Chem. 271:3963-3966.

Cote, F., J.-F. Collard, and J.-P. Julien. 1993. Progressive neuronopathy in transgenic mice expressing the human neurofilament heavy gene: a mouse model of amyotrophic lateral sclerosis. Cell. 73:35-46.

Cote, J.F., and K. Vuori. 2002. Identification of an evolutionarily conserved superfamily of DOCK180-related proteins with guanine nucleotide exchange activity. J Cell Sci. 115:4901-4913.

Couillard-Després, S., Q.Z. Zhu, P.C. Wong, D.L. Price, D.W. Cleveland, and J.P.

202 Julien. 1998. Protective effect of neurofilament heavy gene overexpression in motor neuron disease induced by mutant superoxide dismutase. Proc.Natl.Acad.Sci.USA. 95:9626-9630.

Couratier, P., F.H. Yi, J.L. Preud'homme, P. Clavelou, A. White, P. Sindou, J.M. Vallat, and M.O. Jauberteau. 1998. Serum autoantibodies to neurofilament proteins in sporadic amyotrophic lateral sclerosis. J.Neurol.Sci. 154:137-145.

Crow, J.P., J.B. Sampson, Y. Zhuang, J.A. Thompson, and J.S. Beckman. 1997. Decreased zinc affinity of amyotrophic lateral sclerosis-associated superoxide dismutase mutants leads to enhanced catalysis of tyrosine nitration by peroxynitrite. J Neurochem. 69:1936-1944.

Da Silva, J.S., M. Medina, C. Zuliani, A. Di Nardo, W. Witke, and C.G. Dotti. 2003. RhoA/ROCK regulation of neuritogenesis via profilin IIa-mediated control of actin stability. J Cell Biol. 162:1267-1279.

D'Adamo, P., A. Menegon, C. Lo Nigro, M. Grasso, M. Gulisano, F. Tamanini, T. Bienvenu, A.K. Gedeon, B. Oostra, S.K. Wu, A. Tandon, F. Valtorta, W.E. Balch, J. Chelly, and D. Toniolo. 1998. Mutations in GDI1 are responsible for X-linked non- specific mental retardation. Nat Genet. 19:134-139.

Dahme, M., U. Bartsch, R. Martini, B. Anliker, M. Schachner, and N. Mantei. 1997. Disruption of the mouse L1 gene leads to malformations of the nervous system. Nat Genet. 17:346-349.

Dai, J., and M.P. Sheetz. 1995. Axon membrane flows from the growth cone to the cell body. Cell. 83:693-701.

Dal Canto, M.C., and M.E. Gurney. 1995. Neuropathological changes in two lines of mice carrying a transgene for mutant human Cu,Zn SOD, and in mice overexpressing wild type human SOD: A model of familial amyotrophic lateral sclerosis (FALS). Brain Res. 676:25-40.

203 Dal Canto, M.C., and M.E. Gurney. 1997. A low expressor line of transgenic mice carrying a mutant human Cu,Zn superoxide dismutase (SOD1) gene develops pathological changes that most closely resemble those in human amyotrophic lateral sclerosis. Acta Neuropathol (Berl). 93:537-550.

Dalpozzo, F., M.G. Rossetto, F. Boaretto, E. Sartori, M.L. Mostacciuolo, A. Daga, M.T. Bassi, and A. Martinuzzi. 2003. Infancy onset hereditary spastic paraplegia associated with a novel atlastin mutation. Neurology. 61:580-581.

Dan, C., N. Nath, M. Liberto, and A. Minden. 2002. PAK5, a new brain-specific kinase, promotes neurite outgrowth in N1E-115 cells. Mol Cell Biol. 22:567-577.

Daniels, R.H., P.S. Hall, and G.M. Bokoch. 1998. Membrane targeting of p21-activated kinase 1 (PAK1) induces neurite outgrowth from PC12 cells. Embo J. 17:754-764.

Daniels, R.H., F.T. Zenke, and G.M. Bokoch. 1999. alphaPix stimulates p21-activated kinase activity through exchange factor-dependent and -independent mechanisms. J Biol Chem. 274:6047-6050.

Das, B., X. Shu, G.J. Day, J. Han, U.M. Krishna, J.R. Falck, and D. Broek. 2000. Control of intramolecular interactions between the pleckstrin homology and Dbl homology domains of Vav and Sos1 regulates Rac binding. J Biol Chem. 275:15074- 15081.

Daub, H., K. Gevaert, J. Vandekerckhove, A. Sobel, and A. Hall. 2001. Rac/Cdc42 and p65PAK regulate the microtubule-destabilizing protein stathmin through phosphorylation at serine 16. J Biol Chem. 276:1677-1680.

De Jonghe, P., L. Krols, A. Michalik, J. Hazan, G. Smeyers, A. Lofgren, J. Weissenbach, J.J. Martin, and C. Van Broeckhoven. 1996. Pure familial spastic paraplegia: clinical and genetic analysis of nine Belgian pedigrees. Eur J Hum Genet. 4:260-266.

204 De Jonghe, P., I. Mersivanova, E. Nelis, J. Del Favero, J.J. Martin, C. Van Broeckhoven, O.C. Evgrafov, and V. Timmerman. 2001. Further evidence that neurofilament light chain gene mutations can cause Charcot-Marie-Tooth disease type 2E. Ann. Neurol. 49:245-249.

Debant, A., C. Serra-Pages, K. Seipel, S. O'Brien, M. Tang, S.H. Park, and M. Streuli. 1996. The multidomain protein Trio binds the LAR transmembrane tyrosine phosphatase, contains a protein kinase domain, and has separate rac-specific and rho- specific guanine nucleotide exchange factor domains. Proc Natl Acad Sci U S A. 93:5466-5471.

Deitch, J.S., G.M. Alexander, L. Del Valle, and T.D. Heiman-Patterson. 2002. GLT-1 glutamate transporter levels are unchanged in mice expressing G93A human mutant SOD1. J Neurol Sci. 193:117-126.

DeMali, K.A., K. Wennerberg, and K. Burridge. 2003. Integrin signaling to the actin cytoskeleton. Curr Opin Cell Biol. 15:572-582.

Deng, H.X., A. Hentati, J.A. Tainer, Z. Iqbal, A. Cayabyab, W.Y. Hung, E.D. Getzoff, P. Hu, B. Herzfeldt, R.P. Roos, C. Warner, G. Deng, E. Soriano, C. Smyth, H.E. Parge, A. Ahmed, A.D. Roses, R.A. Hallewell, M.A. Pericakvance, and T. Siddique. 1993. Amyotrophic-Lateral-Sclerosis and Structural Defects in Cu,Zn Superoxide-Dismutase. Science. 261:1047-1051.

Devon, R.S., J.R. Helm, G.A. Rouleau, Y. Leitner, T. Lerman-Sagie, D. Lev, and M.R. Hayden. 2003. The first nonsense mutation in alsin results in a homogeneous phenotype of infantile-onset ascending spastic paralysis with bulbar involvement in two siblings. Clin Genet. 64:210-215.

Devon, R.S., C. Schwab, J.D. Topp, P.C. Orban, Y.Z. Yang, T.D. Pape, J.R. Helm, T.L. Davidson, D.A. Rogers, F. Gros-Louis, G. Rouleau, B.F. Horazdovsky, B.R. Leavitt, and M.R. Hayden. 2005. Cross-species characterization of the ALS2 gene and analysis of its pattern of expression in development and adulthood. Neurobiol Dis. 18:243-257.

205 Dickson, B.J. 2001. Rho GTPases in growth cone guidance. Curr Opin Neurobiol. 11:103-110.

Diebold, B.A., B. Fowler, J. Lu, M.C. Dinauer, and G.M. Bokoch. 2004. Antagonistic cross-talk between Rac and Cdc42 GTPases regulates generation of reactive oxygen species. J Biol Chem. 279:28136-28142.

Donaghy, M. 1999. Classification and clinical features of motor neurone diseases and motor neuropathies in adults. J Neurol. 246:331-333.

Drachman, D.B., V. Chaudhry, D. Cornblath, R.W. Kuncl, A. Pestronk, L. Clawson, E.D. Mellits, S. Quaskey, T. Quinn, A. Calkins, and et al. 1994. Trial of immunosuppression in amyotrophic lateral sclerosis using total lymphoid irradiation. Ann Neurol. 35:142-150.

Drachman, D.B., K. Frank, M. Dykes-Hoberg, P. Teismann, G. Almer, S. Przedborski, and J.D. Rothstein. 2002. Cyclooxygenase 2 inhibition protects motor neurons and prolongs survival in a transgenic mouse model of ALS. Ann. Neurol. 52:771-778.

Drory, V.E., M. Birnbaum, A.D. Korczyn, and J. Chapman. 2001. Association of APOE epsilon4 allele with survival in amyotrophic lateral sclerosis. J Neurol Sci. 190:17-20.

Duarte, F., S. Binet, L. Lacomblez, P. Bouche, J.L. Preud'homme, and V. Meininger. 1991. Quantitative analysis of monoclonal immunoglobulins in serum of patients with amyotrophic lateral sclerosis. J Neurol Sci. 104:88-91.

Duchen, L.W., and S.J. Strich. 1968. An hereditary motor neurone disease with progressive denervation of muscle in the mouse: the mutant 'wobbler'. J Neurol Neurosurg Psychiatry. 31:535-542.

Ekblom, J., S.S. Jossan, L. Oreland, E. Walum, and S.M. Aquilonius. 1994. Reactive gliosis and monoamine oxidase B. J Neural Transm Suppl. 41:253-258.

206 Elliott, J.L., and W.D. Snider. 1995. Parvalbumin is a marker of ALS-resistant motor neurons. Neuroreport. 6:449-452.

Engelhardt, J.I., and S.H. Appel. 1990. IgG reactivity in the spinal cord and motor cortex in amyotrophic lateral sclerosis. Arch Neurol. 47:1210-1216.

Errico, A., A. Ballabio, and E.I. Rugarli. 2002. Spastin, the protein mutated in autosomal dominant hereditary spastic paraplegia, is involved in microtubule dynamics. Hum Mol Genet. 11:153-163.

Escurat, M., K. Djabali, M. Gumpel, F. Gros, and M.M. Portier. 1990. Differential expression of two neuronal intermediate-filament proteins, peripherin and the low- molecular-mass neurofilament protein NF-L during the development of the rat. J. Neurosci. 10:764-784.

Estrach, S., S. Schmidt, S. Diriong, A. Penna, A. Blangy, P. Fort, and A. Debant. 2002. The Human Rho-GEF trio and its target GTPase RhoG are involved in the NGF pathway, leading to neurite outgrowth. Curr Biol. 12:307-312.

Etienne-Manneville, S., and A. Hall. 2002. Rho GTPases in cell biology. Nature. 420:629-635.

Evans, L.L., and P.C. Bridgman. 1995. Particles move along actin filament bundles in nerve growth cones. Proc Natl Acad Sci U S A. 92:10954-10958.

Eymard-Pierre, E., G. Lesca, S. Dollet, F.M. Santorelli, M. di Capua, E. Bertini, and O. Boespflug-Tanguy. 2002. Infantile-onset ascending hereditary spastic paralysis is associated with mutations in the alsin gene. Am J Hum Genet. 71:518-527.

Fabrizi, G.M., T. Cavallaro, C. Angiari, L. Bertolasi, I. Cabrini, M. Ferrarini, and N. Rizzuto. 2004. Giant axon and neurofilament accumulation in Charcot-Marie-Tooth disease type 2E. Neurology. 62:1429-1431.

207 Ferrante, R.J., L.A. Shinobu, J.B. Schulz, R.T. Matthews, C.E. Thomas, N.W. Kowall, M.E. Gurney, and M.F. Beal. 1997. Increased 3-nitrotyrosine and oxidative damage in mice with a human copper/zinc superoxide dismutase mutation. Ann.Neurol. 42:326- 334.

Ferreirinha, F., A. Quattrini, M. Pirozzi, V. Valsecchi, G. Dina, V. Broccoli, A. Auricchio, F. Piemonte, G. Tozzi, L. Gaeta, G. Casari, A. Ballabio, and E.I. Rugarli. 2004. Axonal degeneration in paraplegin-deficient mice is associated with abnormal mitochondria and impairment of axonal transport. J Clin Invest. 113:231-242.

Figlewicz, D.A., A. Krizus, M.G. Martinoli, V. Meininger, M. Dib, G.A. Rouleau, and J.-P. Julien. 1994. Variants of the heavy neurofilament subunit are associated with the development of amyotrophic lateral sclerosis. Hum. Mol. Genet. 3:1757-1761.

Fleming, I.N., A. Gray, and C.P. Downes. 2000. Regulation of the Rac1-specific exchange factor Tiam1 involves both phosphoinositide 3-kinase-dependent and - independent components. Biochem J. 351:173-182.

Flood, D.G., Reaume, A.G., Gruner, J.A., Hoffman, E.K., Hirsch, J.D., Lin, Y.G., Dorfman, K.S. and R.W. Scott. 1999. Hindlimb motor neurons require Cu/Zn superoxide dismutase for maintenance of neuromuscular junctions. Am J Pathol. 155:663-672

Flowers, J.M., J.F. Powell, P.N. Leigh, P. Andersen, and C.E. Shaw. 2001. Intron 7 retention and exon 9 skipping EAAT2 mRNA variants are not associated with amyotrophic lateral sclerosis. Ann Neurol. 49:643-649.

Fournier, A.E., B.T. Takizawa, and S.M. Strittmatter. 2003. Rho kinase inhibition enhances axonal regeneration in the injured CNS. J Neurosci. 23:1416-1423.

Frangiskakis, J.M., A.K. Ewart, C.A. Morris, C.B. Mervis, J. Bertrand, B.F. Robinson, B.P. Klein, G.J. Ensing, L.A. Everett, E.D. Green, C. Proschel, N.J. Gutowski, M. Noble, D.L. Atkinson, S.J. Odelberg, and M.T. Keating. 1996. LIM-kinase1

208 hemizygosity implicated in impaired visuospatial constructive cognition. Cell. 86:59-69.

Fransen, E., R. D'Hooge, G. Van Camp, M. Verhoye, J. Sijbers, E. Reyniers, P. Soriano, H. Kamiguchi, R. Willemsen, S.K. Koekkoek, C.I. De Zeeuw, P.P. De Deyn, A. Van der Linden, V. Lemmon, R.F. Kooy, and P.J. Willems. 1998. L1 knockout mice show dilated ventricles, vermis hypoplasia and impaired exploration patterns. Hum Mol Genet. 7:999-1009.

Frost, J.A., A. Khokhlatchev, S. Stippec, M.A. White, and M.H. Cobb. 1998. Differential effects of PAK1-activating mutations reveal activity-dependent and - independent effects on cytoskeletal regulation. J Biol Chem. 273:28191-28198.

Frost, J.A., J.L. Swantek, S. Stippec, M.J. Yin, R. Gaynor, and M.H. Cobb. 2000. Stimulation of NFkappa B activity by multiple signaling pathways requires PAK1. J Biol Chem. 275:19693-19699.

Frost, J.A., S. Xu, M.R. Hutchison, S. Marcus, and M.H. Cobb. 1996. Actions of Rho family small G proteins and p21-activated protein kinases on mitogen-activated protein kinase family members. Mol Cell Biol. 16:3707-3713.

Fujita, Y., K. Okamoto, A. Sakurai, N.K. Gonatas, and A. Hirano. 2000. Fragmentation of the Golgi apparatus of the anterior horn cells in patients with familial amyotrophic lateral sclerosis with SOD1 mutations and posterior column involvement. J. Neurol. Sci. 174:137-140.

Fujita, Y., K. Okamoto, A. Sakurai, H. Kusaka, H. Aizawa, B. Mihara, and N.K. Gonatas. 2002. The Golgi apparatus is fragmented in spinal cord motor neurons of amyotrophic lateral sclerosis with basophilic inclusions. Acta Neuropathol (Berl). 103:243-247.

Gampel, A., P.J. Parker, and H. Mellor. 1999. Regulation of epidermal growth factor receptor traffic by the small GTPase rhoB. Curr Biol. 9:955-958.

209 Gascon, G.G., P. Chavis, A. Yaghmour, B. Stigsby, A. Shums, P. Ozand, and T. Siddique. 1995. Familial childhood primary lateral sclerosis with associated gaze paresis. Neuropediatrics. 26:313-319.

Gauthier-Rouviere, C., E. Vignal, M. Meriane, P. Roux, P. Montcourier, and P. Fort. 1998. RhoG GTPase controls a pathway that independently activates Rac1 and Cdc42Hs. Mol Biol Cell. 9:1379-1394.

Gelbard, H.A. 2004. Synapses and Sisyphus: life without paraplegin. J Clin Invest. 113:185-187.

Gélinas, S., C. Chapados, M. Beauregard, I. Gosselin, and M.G. Martinoli. 2000. Effect of oxidative stress on stability and structure of neurofilament proteins. Biochem.Cell Biol. 78:667-674.

Georgiou, D.M., J. Zidar, M. Korosec, L.T. Middleton, T. Kyriakides, and K. Christodoulou. 2002. A novel NF-L mutation Pro22Ser is associated with CMT2 in a large Slovenian family. Neurogenetics. 4:93-96.

Ghatak, N.R., W.W. Campbell, R.H. Lippman, and M.G. Hadfield. 1986. Anterior horn changes of motor neuron disease associated with demyelinating radiculopathy. J Neuropathol Exp Neurol. 45:385-395.

Giasson, B.I., and W.E. Mushynski. 1998. Intermediate filament disassembly in clutured dorsal root ganglion neurons is associated with amino-terminal head domain phosphorylation of specific subunits. J. Neurochem. 70:1869-1875.

Gineitis, D., and R. Treisman. 2001. Differential usage of signal transduction pathways defines two types of serum response factor target gene. J Biol Chem. 276:24531-24539.

Gonatas, N.K., J.O. Gonatas, and A. Stieber. 1998. The involvement of the Golgi apparatus in the pathogenesis of amyotrophic lateral sclerosis, Alzheimer's disease, and ricin intoxication. Histochem Cell Biol. 109:591-600.

210 Gong, Y.H., A.S. Parsadanian, A. Andreeva, W.D. Snider, and J.L. Elliot. 2000. Restricted expression of G86R Cu/Zn superoxide dismutase in astrocytes results in astrocytosis but does not cause motoneuron degeneration. J. Neuroscience. 20:660-665.

Gorvel, J.P., P. Chavrier, M. Zerial, and J. Gruenberg. 1991. rab5 controls early endosome fusion in vitro. Cell. 64:915-925.

Graves, M.C., M. Fiala, L.A. Dinglasan, N.Q. Liu, J. Sayre, F. Chiappelli, C. van Kooten, and H.V. Vinters. 2004. Inflammation in amyotrophic lateral sclerosis spinal cord and brain is mediated by activated macrophages, mast cells and T cells. Amyotroph Lateral Scler Other Motor Neuron Disord. 5:213-219.

Greenway, M.J., M.D. Alexander, S. Ennis, B.J. Traynor, B. Corr, E. Frost, A. Green, and O. Hardiman. 2004. A novel candidate region for ALS on chromosome 14q11.2. Neurology. 63:1936-1938.

Grodberg, J., and J.J. Dunn. 1988. ompT encodes the Escherichia coli outer membrane protease that cleaves T7 RNA polymerase during purification. J Bacteriol. 170:1245- 1253.

Gros-Louis, F., R. Lariviere, G. Gowing, S. Laurent, W. Camu, J.P. Bouchard, V. Meininger, G.A. Rouleau, and J.P. Julien. 2004. A frameshift deletion in peripherin gene associated with amyotrophic lateral sclerosis. J Biol Chem. 279:45951-45956.

Gros-Louis, F., S. Laurent, A.A. Lopes, J. Khoris, V. Meininger, W. Camu, and G.A. Rouleau. 2003a. Absence of mutations in the hypoxia response element of VEGF in ALS. Muscle Nerve. 28:774-775.

Gros-Louis, F., I.A. Meijer, C.K. Hand, M.P. Dube, D.L. MacGregor, M.H. Seni, R.S. Devon, M.R. Hayden, F. Andermann, E. Andermann, and G.A. Rouleau. 2003b. An ALS2 gene mutation causes hereditary spastic paraplegia in a Pakistani kindred. Ann. Neurol. 53:144-145.

211 Grosse, G., J. Grosse, R. Tapp, J. Kuchinke, M. Gorsleben, I. Fetter, B. Hohne-Zell, M. Gratzl, and M. Bergmann. 1999. SNAP-25 requirement for dendritic growth of hippocampal neurons. J Neurosci Res. 56:539-546.

Grunnet, M.L., C. Leicher, A. Zimmerman, E. Zalneraitis, and M. Barwick. 1989. Primary lateral sclerosis in a child. Neurology. 39:1530-1532.

Guan, R.J., B.S. Khatra, and J.A. Cohlberg. 1991. Phosphorylation of bovine neurofilament proteins by protein kinase FA (Glycogen synthase kinase 3). J. Biol. Chem. 266:8262-8267.

Guan, Z., S.Y. Buckman, A.P. Pentland, D.J. Templeton, and A.R. Morrison. 1998. Induction of cyclooxygenase-2 by the activated MEKK1 --> SEK1/MKK4 --> p38 mitogen-activated protein kinase pathway. J Biol Chem. 273:12901-12908.

Guegan, C., Vila, M., Rosoklija, G., Hays, A.P., and S. Przedborski. 2001. Recruitment of the mitochondrial-dependent apoptotic pathway in amyotrophic lateral sclerosis. J. Neurosci. 21:6569-6576.

Guidato, S., L.-H. Tsai, J. Woodgett, and C.C.J. Miller. 1996. Differential cellular phosphorylation of neurofilament heavy side-arms by glycogen synthase kinase-3 and cyclin-dependent kinase-5. J. Neurochem. 66:1698-1706.

Guipponi, M., H.S. Scott, H. Chen, A. Schebesta, C. Rossier, and S.E. Antonarakis. 1998. Two isoforms of a human intersectin (ITSN) protein are produced by brain- specific alternative splicing in a stop codon. Genomics. 53:369-376.

Gurney, M.E. 1997. The use of transgenic mouse models of amyotrophic lateral sclerosis in preclinical drug studies. J.Neurol.Sci. 152 Suppl. 1:S67-S73.

Gurney, M.E., H. Pu, A.Y. Chiu, M.C. Dal Canto, C.Y. Polchow, D.D. Alexander, J. Caliendo, A. Hentati, Y.W. Kwon, H.X. Deng, and et al. 1994. Motor neuron degeneration in mice that express a human Cu,Zn superoxide dismutase mutation.

212 Science. 264:1772-1775.

Hadano, S., C.K. Hand, H. Osuga, Y. Yanagisawa, A. Otomo, R.S. Devon, N. Miyamoto, J. Showguchi-Miyata, Y. Okada, R. Singaraja, D.A. Figlewicz, T. Kwiatkowski, B.A. Hosler, T. Sagie, J. Skaug, J. Nasir, R.H. Brown, Jr., S.W. Scherer, G.A. Rouleau, M.R. Hayden, and J.E. Ikeda. 2001. A gene encoding a putative GTPase regulator is mutated in familial amyotrophic lateral sclerosis 2. Nature Genet. 29:166- 173.

Hadano, S., A. Otomo, K. Suzuki-Utsunomiya, R. Kunita, Y. Yanagisawa, J. Showguchi-Miyata, H. Mizumura, and J.E. Ikeda. 2004. ALS2CL, the novel protein highly homologous to the carboxy-terminal half of ALS2, binds to Rab5 and modulates endosome dynamics. FEBS Lett. 575:64-70.

Hafezparast, M., A. Ahmad-Annuar, H. Hummerich, P. Shah, M. Ford, C. Baker, S. Bowen, J.E. Martin, and E.M. Fisher. 2003a. Paradigms for the identification of new genes in motor neuron degeneration. Amyotroph Lateral Scler Other Motor Neuron Disord. 4:249-257.

Hafezparast, M., R. Klocke, C. Ruhrberg, A. Marquardt, A. Ahmad-Annuar, S. Bowen, G. Lalli, A.S. Witherden, H. Hummerich, S. Nicholson, P.J. Morgan, R. Oozageer, J.V. Priestley, S. Averill, V.R. King, S. Ball, J. Peters, T. Toda, A. Yamamoto, Y. Hiraoka, M. Augustin, D. Korthaus, S. Wattler, P. Wabnitz, C. Dickneite, S. Lampel, F. Boehme, G. Peraus, A. Popp, M. Rudelius, J. Schlegel, H. Fuchs, M.H. de Angelis, G. Schiavo, D.T. Shima, A.P. Russ, G. Stumm, J.E. Martin, and E.M. Fisher. 2003b. Mutations in dynein link motor neuron degeneration to defects in retrograde transport. Science. 300:808-812.

Hall, A. 1998. Rho GTPases and the actin cytoskeleton. Science. 279:509-514.

Hall, E.D., J.A. Oostveen, and M.E. Gurney. 1998. Relationship of microglial and astrocytic activation to disease onset and progression in a transgenic model of familial ALS. Glia. 23:249-256.

213 Hama, H., G.G. Tall, and B.F. Horazdovsky. 1999. Vps9p is a guanine nucleotide exchange factor involved in vesicle-mediated vacuolar protein transport. J Biol Chem. 274:15284-15291.

Hammer, R.P., Jr., U. Tomiyasu, and A.B. Scheibel. 1979. Degeneration of the human Betz cell due to amyotrophic lateral sclerosis. Exp Neurol. 63:336-346.

Han, J., K. Luby-Phelps, B. Das, X. Shu, Y. Xia, R.D. Mosteller, U.M. Krishna, J.R. Falck, M.A. White, and D. Broek. 1998. Role of substrates and products of PI 3-kinase in regulating activation of Rac-related guanosine triphosphatases by Vav. Science. 279:558-560.

Hanahan, D. 1983. Studies on transformation of Escherichia coli with plasmids. J Mol Biol. 166:557-580.

Hand, C.K., J. Khoris, F. Salachas, F. Gros-Louis, A.A.S. Lopes, V. Mayeux-Portas, R.H. Brown, Jr., V. Meininger, W. Camu, and G.A. Rouleau. 2002. A novel locus for familial amyotrophic lateral sclerosis, on chromosome 18q. Am.J.Hum.Genet. 70:251- 256.

Hansen, J.J., A. Durr, I. Cournu-Rebeix, C. Georgopoulos, D. Ang, M.N. Nielsen, C.S. Davoine, A. Brice, B. Fontaine, N. Gregersen, and P. Bross. 2002. Hereditary spastic paraplegia SPG13 is associated with a mutation in the gene encoding the mitochondrial chaperonin Hsp60. Am J Hum Genet. 70:1328-1332.

Hart, M.J., A. Eva, T. Evans, S.A. Aaronson, and R.A. Cerione. 1991. Catalysis of guanine nucleotide exchange on the CDC42Hs protein by the dbl oncogene product. Nature. 354:311-314.

Hart, M.J., A. Eva, D. Zangrilli, S.A. Aaronson, T. Evans, R.A. Cerione, and Y. Zheng. 1994. Cellular transformation and guanine nucleotide exchange activity are catalyzed by a common domain on the dbl oncogene product. J Biol Chem. 269:62-65.

214 Hart, M.J., X. Jiang, T. Kozasa, W. Roscoe, W.D. Singer, A.G. Gilman, P.C. Sternweis, and G. Bollag. 1998. Direct stimulation of the guanine nucleotide exchange activity of p115 RhoGEF by Galpha13. Science. 280:2112-2114.

Hayashi, K., T. Ohshima, and K. Mikoshiba. 2002. Pak1 is involved in dendrite initiation as a downstream effector of Rac1 in cortical neurons. Mol Cell Neurosci. 20:579-594.

Hayashi, S., A. Sakurai, M. Amari, and K. Okamoto. 2001. Pathological study of the diffuse myelin pallor in the anterolateral columns of the spinal cord in amyotrophic lateral sclerosis. J Neurol Sci. 188:3-7.

Hayward, C., S. Colville, R.J. Swingler, and D.J. Brock. 1999. Molecular genetic analysis of the APEX nuclease gene in amyotrophic lateral sclerosis. Neurology. 52:1899-1901.

Hazan, J., N. Fonknechten, D. Mavel, C. Paternotte, D. Samson, F. Artiguenave, C.S. Davoine, C. Cruaud, A. Durr, P. Wincker, P. Brottier, L. Cattolico, V. Barbe, J.M. Burgunder, J.F. Prud'homme, A. Brice, B. Fontaine, B. Heilig, and J. Weissenbach. 1999. Spastin, a new AAA protein, is altered in the most frequent form of autosomal dominant spastic paraplegia. Nat Genet. 23:296-303.

He, C.Z., and A.P. Hays. 2004. Expression of peripherin in ubiquinated inclusions of amyotrophic lateral sclerosis. J. Neurol. Sci. 217:47-54.

He, Y., F. Francis, K.A. Myers, W. Yu, M.M. Black, and P.W. Baas. 2005. Role of cytoplasmic dynein in the axonal transport of microtubules and neurofilaments. J Cell Biol. 168:697-703.

Heath, P.R., and P.J. Shaw. 2002. Update on the glutamatergic neurotransmitter system and the role of excitotoxicity in amyotrophic lateral sclerosis. Muscle Nerve. 26:438- 458.

215 Hentati, A., K. Ouahchi, M.A. Pericak-Vance, D. Nijhawan, A. Ahmad, Y. Yang, J. Rimmler, W. Hung, B. Schlotter, A. Ahmed, M. Ben Hamida, F. Hentati, and T. Siddique. 1998. Linkage of a commoner form of recessive amyotrophic lateral sclerosis to chromosome 15q15-q22 markers. Neurogenetics. 2:55-60.

Higgins, C.M.J., C.W. Jung, H.L. Ding, and Z.S. Xu. 2002. Mutant Cu, Zn superoxide dismutase that causes motoneuron degeneration is present in mitochondria in the CNS - art. no. RC215. J.Neurosci. 22: 1-6.

Hill, C.S., J. Wynne, and R. Treisman. 1995. The Rho family GTPases RhoA, Rac1, and CDC42Hs regulate transcriptional activation by SRF. Cell. 81:1159-1170.

Hirano, A., H. Donnenfeld, S. Sasaki, and I. Nakano. 1984. Fine structural observations of neurofilamentous changes in amyotrophic lateral sclerosis. J. Neuropathol. Exp. Neurol. 43:461-470.

Hirling, H., P. Steiner, C. Chaperon, R. Marsault, R. Regazzi, and S. Catsicas. 2000. Syntaxin 13 is a developmentally regulated SNARE involved in neurite outgrowth and endosomal trafficking. Eur J Neurosci. 12:1913-1923.

Hodgson, E.K., and I. Fridovich. 1975. The interaction of bovine erythrocyte superoxide dismutase with hydrogen peroxide: chemiluminescence and peroxidation. Biochemistry. 14:5299-5303.

Hoffman, E.K., H.M. Wilcox, R.W. Scott, and R. Siman. 1996. Proteasome inhibition enhances the stability of mouse Cu/Zn superoxide dismutase with mutations linked to familial amyotrophic lateral sclerosis. J Neurol Sci. 139:15-20.

Horiuchi, H., R. Lippe, H.M. McBride, M. Rubino, P. Woodman, H. Stenmark, V. Rybin, M. Wilm, K. Ashman, M. Mann, and M. Zerial. 1997. A novel Rab5 GDP/GTP exchange factor complexed to Rabaptin-5 links nucleotide exchange to effector recruitment and function. Cell. 90:1149-1159.

216 Hosler, B.A., T. Siddique, P.C. Sapp, W. Sailor, M.C. Huang, A. Hossain, J.R. Daube, M. Nance, C.H. Fan, J. Kaplan, W.Y. Hung, D. McKenna-Yasek, J.L. Haines, M.A. Pericak-Vance, H.R. Horvitz, and R.H. Brown, Jr. 2000. Linkage of familial amyotrophic lateral sclerosis with frontotemporal dementia to chromosome 9q21-q22. JAMA. 284:1664-1669.

Hough, M.A., J.G. Grossmann, S.V. Antonyuk, R.W. Strange, P.A. Doucette, J.A. Rodriguez, L.J. Whitson, P.J. Hart, L.J. Hayward, J.S. Valentine, and S.S. Hasnain. 2004. Dimer destabilization in superoxide dismutase may result in disease-causing properties: Structures of motor neuron disease mutants. Proc. Natl. Acad. Sci. USA. 101:5976-5981.

Howland, D.S., J. Liu, Y.J. She, B. Goad, N.J. Maragakis, B. Kim, J. Erickson, J. Kulik, L. DeVito, G. Psaltis, L.J. DeGennaro, D.W. Cleveland, and J.D. Rothstein. 2002. Focal loss of the glutamate transporter EAAT2 in a transgenic rat model of SOD1 mutant- mediated amyotrophic lateral sclerosis (ALS). Proc.Natl.Acad.Sci.USA. 99:1604-1609.

Hsu, S.M., and L. Raine. 1981. Protein A, avidin, and biotin in immunohistochemistry. J Histochem Cytochem. 29:1349-1353.

Hu, J.H., K. Chernoff, S. Pelech, and C. Krieger. 2003a. Protein kinase and protein phosphatase expression in the central nervous system of G93A mSOD over-expressing mice. J Neurochem. 85:422-431.

Hu, J.H., H. Zhang, R. Wagey, C. Krieger, and S.L. Pelech. 2003b. Protein kinase and protein phosphatase expression in amyotrophic lateral sclerosis spinal cord. J. Neurochem. 85:432-442.

Huber, A.B., A.L. Kolodkin, D.D. Ginty, and J.F. Cloutier. 2003. Signaling at the growth cone: ligand-receptor complexes and the control of axon growth and guidance. Annu Rev Neurosci. 26:509-563.

Huber, L.A., P. Dupree, and C.G. Dotti. 1995. A deficiency of the small GTPase rab8

217 inhibits membrane traffic in developing neurons. Mol. Cell Biol. 15:918-924.

Hughes, C.A., P.C. Byrne, S. Webb, P. McMonagle, V. Patterson, M. Hutchinson, and N.A. Parfrey. 2001. SPG15, a new locus for autosomal recessive complicated HSP on chromosome 14q. Neurology. 56:1230-1233.

Hughes, J.T. 1982. Pathology of amyotrophic lateral sclerosis. Adv Neurol. 36:61-74.

Hughes, M., C. Zhang, J.M. Avis, C.J. Hutchison, and P.R. Clarke. 1998. The role of the ran GTPase in nuclear assembly and DNA replication: characterisation of the effects of Ran mutants. J Cell Sci. 111 ( Pt 20):3017-3026.

Hurd, D.D., and W.M. Saxton. 1996. Kinesin mutations cause motor neuron disease phenotypes by disrupting fast axonal transport in Drosophila. Genetics. 144:1075-1085.

Hussain, N.K., S. Jenna, M. Glogauer, C.C. Quinn, S. Wasiak, M. Guipponi, S.E. Antonarakis, B.K. Kay, T.P. Stossel, N. Lamarche-Vane, and P.S. McPherson. 2001. Endocytic protein intersectin-l regulates actin assembly via Cdc42 and N-WASP. Nat Cell Biol. 3:927-932.

Igarashi, M., S. Kozaki, S. Terakawa, S. Kawano, C. Ide, and Y. Komiya. 1996. Growth cone collapse and inhibition of neurite growth by Botulinum neurotoxin C1: a t-SNARE is involved in axonal growth. J Cell Biol. 134:205-215.

Imamura, H., K. Takaishi, K. Nakano, A. Kodama, H. Oishi, H. Shiozaki, M. Monden, T. Sasaki, and Y. Takai. 1998. Rho and Rab small G proteins coordinately reorganize stress fibers and focal adhesions in MDCK cells. Mol Biol Cell. 9:2561-2575.

Ince, P., N. Stout, P. Shaw, J. Slade, W. Hunziker, C.W. Heizmann, and K.G. Baimbridge. 1993. Parvalbumin and calbindin D-28k in the human motor system and in motor neuron disease. Neuropathol Appl Neurobiol. 19:291-299.

Irving, N.G., and C.C.J. Miller. 1997. Tau phosphorylation in cells transfected with

218 wild-type or an Alzheimer's disease mutant Presenilin-1. Neurosci.Lett. 222:71-74.

Ishizaki, T., M. Maekawa, K. Fujisawa, K. Okawa, A. Iwamatsu, A. Fujita, N. Watanabe, Y. Saito, A. Kakizuka, N. Morii, and S. Narumiya. 1996. The small GTP- binding protein Rho binds to and activates a 160 kDa Ser/Thr protein kinase homologous to myotonic dystrophy kinase. Embo J. 15:1885-1893.

Jaarsma, D., E.D. Haasdijk, J.A. Grashorn, R. Hawkins, W. van Duijn, H.W. Verspaget, J. London, and J.C. Holstege. 2000. Human Cu/Zn superoxide dismutase (SOD1) overexpression in mice causes mitochondrial vacuolization, axonal degeneration, and premature motoneuron death and accelerates motoneuron disease in mice expressing a familial amyotrophic lateral sclerosis mutant SOD1. Neurobiol Dis. 7:623-643.

Jaarsma, D., F. Rognoni, W. Van Duijn, H.W. Verspaget, E.D. Haasdijk, and J.C. Holstege. 2001. CuZn superoxide dismutase (SOD1) accumulates in vacuolated mitochondria in transgenic mice expressing amyotrophic lateral sclerosis-linked SOD1 mutations. Acta Neuropathol.(Berl.). 102:293-305.

Jablonka, S., S. Wiese, and M. Sendtner. 2004. Axonal defects in mouse models of motoneuron disease. J Neurobiol. 58:272-286.

Jalink, K., T. Eichholtz, F.R. Postma, E.J. van Corven, and W.H. Moolenaar. 1993. Lysophosphatidic acid induces neuronal shape changes via a novel, receptor-mediated signaling pathway: similarity to thrombin action. Cell Growth Differ. 4:247-255.

Jalink, K., E.J. van Corven, T. Hengeveld, N. Morii, S. Narumiya, and W.H. Moolenaar. 1994. Inhibition of lysophosphatidate- and thrombin-induced neurite retraction and neuronal cell rounding by ADP ribosylation of the small GTP-binding protein Rho. J Cell Biol. 126:801-810.

Jin, Z., and S.M. Strittmatter. 1997. Rac1 mediates collapsin-1-induced growth cone collapse. J Neurosci. 17:6256-6263.

219 Johnston, J.A., M.J. Dalton, M.E. Gurney, and R.R. Kopito. 2000. Formation of high molecular weight complexes of mutant Cu,Zn-superoxide dismutase in a mouse model for familial amyotrophic lateral sclerosis. Proc. Natl. Acad. Sci. USA. 97:12571-12576.

Jones, S.M., K.E. Howell, J.R. Henley, H. Cao, and M.A. McNiven. 1998. Role of dynamin in the formation of transport vesicles from the trans-Golgi network. Science. 279:573-577.

Joneson, T., and D. Bar-Sagi. 1998. A Rac1 effector site controlling mitogenesis through superoxide production. J Biol Chem. 273:17991-17994.

Joneson, T., M. McDonough, D. Bar-Sagi, and L. Van Aelst. 1996. RAC regulation of actin polymerization and proliferation by a pathway distinct from Jun kinase. Science. 274:1374-1376.

Jordanova, A., P. De Jonghe, C.F. Boerkoel, H. Takashima, E. De Vriendt, C. Ceuterick, J.J. Martin, I.J. Butler, P. Mancias, S. Papasozomenos, D. Terespolsky, L. Potocki, C.W. Brown, M. Shy, D.A. Rita, I. Tournev, I. Kremensky, J.R. Lupski, and V. Timmerman. 2003. Mutations in the neurofilament light chain gene (NEFL) cause early onset severe Charcot-Marie-Tooth disease. Brain. 126:590-597.

Jouet, M., A. Rosenthal, G. Armstrong, J. MacFarlane, R. Stevenson, J. Paterson, A. Metzenberg, V. Ionasescu, K. Temple, and S. Kenwrick. 1994. X-linked spastic paraplegia (SPG1), MASA syndrome and X-linked hydrocephalus result from mutations in the L1 gene. Nat Genet. 7:402-407.

Kajiho, H., K. Saito, K. Tsujita, K. Kontani, Y. Araki, H. Kurosu, and T. Katada. 2003. RIN3: a novel Rab5 GEF interacting with amphiphysin II involved in the early endocytic pathway. J Cell Sci. 116:4159-4168.

Kamo, H., H. Haebara, I. Akiguchi, M. Kameyama, H. Kimura, and P.L. McGeer. 1987. A distinctive distribution of reactive astroglia in the precentral cortex in amyotrophic lateral sclerosis. Acta Neuropathol (Berl). 74:33-38.

220 Kanekura, K., Y. Hashimoto, Y. Kita, J. Sasabe, S. Aiso, I. Nishimoto, and M. Matsuoka. 2005. A Rac1/phosphatidylinositol 3-kinase/Akt3 anti-apoptotic pathway, triggered by AlsinLF, the product of the ALS2 gene, antagonizes Cu/Zn-superoxide dismutase (SOD1) mutant-induced motoneuronal cell death. J Biol Chem. 280:4532- 4543.

Kanekura, K., Y. Hashimoto, T. Niikura, S. Aiso, M. Matsuoka, and I. Nishimoto. 2004. Alsin, the product of ALS2 gene, suppresses SOD1 mutant neurotoxicity through RhoGEF domain by interacting with SOD1 mutants. J. Biol. Chem. 279:19247-19256.

Karki, S., and E.L. Holzbaur. 1999. Cytoplasmic dynein and dynactin in cell division and intracellular transport. Curr Opin Cell Biol. 11:45-53.

Kato, S., H. Hayashi, K. Nakashima, E. Nanba, M. Kato, A. Hirano, I. Nakano, K. Asayama, and E. Ohama. 1997. Pathological characterization of astrocytic hyaline inclusions in familial amyotrophic lateral sclerosis. Am J Pathol. 151:611-620.

Katoh, H., and M. Negishi. 2003. RhoG activates Rac1 by direct interaction with the Dock180-binding protein Elmo. Nature. 424:461-464.

Kaupmann, K., D. Simon-Chazottes, J.L. Guenet, and H. Jockusch. 1992. Wobbler, a mutation affecting motoneuron survival and gonadal functions in the mouse, maps to proximal chromosome 11. Genomics. 13:39-43.

Kawahara, Y., K. Ito, H. Sun, H. Aizawa, I. Kanazawa, and S. Kwak. 2004. Glutamate receptors: RNA editing and death of motor neurons. Nature. 427:801.

Kawamata, T., H. Akiyama, T. Yamada, and P.L. McGeer. 1992. Immunologic reactions in amyotrophic lateral sclerosis brain and spinal cord tissue. Am J Pathol. 140:691-707.

Kawasaki, H., T. Morooka, S. Shimohama, J. Kimura, T. Hirano, Y. Gotoh, and E. Nishida. 1997. Activation and involvement of p38 mitogen-activated protein kinase in

221 glutamate-induced apoptosis in rat cerebellar granule cells. J. Biol. Chem. 272:18518- 18521.

Kawashima, T., A. Furuta, K. Doh-ura, H. Kikuchi, and T. Iwaki. 2000. Ubiquitin- immunoreactive skein-like inclusions in the neostriatum are not restricted to amyotrophic lateral sclerosis, but are rather aging-related structures. Acta Neuropathol (Berl). 100:43-49.

Kheradmand, F., E. Werner, P. Tremble, M. Symons, and Z. Werb. 1998. Role of Rac1 and oxygen radicals in collagenase-1 expression induced by cell shape change. Science. 280:898-902.

Kieran, D., M. Hafezparast, S. Bohnert, J.R. Dick, J. Martin, G. Schiavo, E.M. Fisher, and L. Greensmith. 2005. A mutation in dynein rescues axonal transport defects and extends the life span of ALS mice. J Cell Biol. 169:561-567.

Kimura, F., R.G. Smith, O. Delbono, O. Nyormoi, T. Schneider, W. Nastainczyk, F. Hofmann, E. Stefani, and S.H. Appel. 1994. Amyotrophic lateral sclerosis patient antibodies label Ca2+ channel alpha 1 subunit. Ann Neurol. 35:164-171.

Kirby, J., F.M. Menzies, M.R. Cookson, K. Bushby, and P.J. Shaw. 2002. Differential gene expression in a cell culture model of SOD1-related familial motor neurone disease. Hum.Mol.Genet. 11:2061-2075.

Kirkinezos, I.G., S.R. Bacman, D. Hernandez, J. Oca-Cossio, L.J. Arias, M.A. Perez- Pinzon, W.G. Bradley, and C.T. Moraes. 2005. Cytochrome c association with the inner mitochondrial membrane is impaired in the CNS of G93A-SOD1 mice. J Neurosci. 25:164-172.

Kita, Y., K.D. Kimura, M. Kobayashi, S. Ihara, K. Kaibuchi, S. Kuroda, M. Ui, H. Iba, H. Konishi, U. Kikkawa, S. Nagata, and Y. Fukui. 1998. Microinjection of activated phosphatidylinositol-3 kinase induces process outgrowth in rat PC12 cells through the Rac-JNK signal transduction pathway. J.Cell Sci. 111:907-915.

222 Kiyokawa, E., Y. Hashimoto, S. Kobayashi, H. Sugimura, T. Kurata, and M. Matsuda. 1998. Activation of Rac1 by a Crk SH3-binding protein, DOCK180. Genes Dev. 12:3331-3336.

Knaus, U.G., Y. Wang, A.M. Reilly, D. Warnock, and J.H. Jackson. 1998. Structural requirements for PAK activation by Rac GTPases. J Biol Chem. 273:21512-21518.

Koehnle, T.J., and A. Brown. 1999. Slow axonal transport of neurofilament protein in cultured neurons. J.Cell Biol. 144:447-458.

Kong, J., and Z. Xu. 1998. Massive mitochondrial degeneration in motor neurons triggers the onset of amyotrophic lateral sclerosis in mice expressing a mutant SOD1. J Neurosci. 18:3241-3250.

Kostic, V., V. Jackson-Lewis, F. De Bilbao, M. Dubois-Dauphin, and S. Przedborski. 1997. Bcl-2: Prolonging life in a transgenic mouse model of familial amyotrophic lateral sclerosis. Science. 277:559-562.

Krekoski, C.A., I.M. Parhad, T.S. Fung, and A.W. Clark. 1996. Aging is associated with divergent effects on Nf-L and GFAP transcription in rat brain. Neurobiol Aging. 17:833- 841.

Krieger, C., R.A. Lanius, S.L. Pelech, and C.A. Shaw. 1996. Amyotrophic lateral sclerosis: the involvement of intracellular Ca2+ and protein kinase C. Trends Pharmacol Sci. 17:114-120.

Kriz, J., M. Nguyen, and J. Julien. 2002. Minocycline slows disease progression in a mouse model of amyotrophic lateral sclerosis. Neurobiol. Dis. 10:268-278.

Kroemer, G., and J.C. Reed. 2000. Mitochondrial control of cell death. Nat Med. 6:513- 519.

Kubiseski, T.J., J. Culotti, and T. Pawson. 2003. Functional analysis of the

223 Caenorhabditis elegans UNC-73B PH domain demonstrates a role in activation of the Rac GTPase in vitro and axon guidance in vivo. Mol Cell Biol. 23:6823-6835.

Kubo, T., T. Yamashita, A. Yamaguchi, H. Sumimoto, K. Hosokawa, and M. Tohyama. 2002. A novel FERM domain including guanine nucleotide exchange factor is involved in Rac signaling and regulates neurite remodeling. J Neurosci. 22:8504-8513.

Kuhn, T.B., M.D. Brown, and J.R. Bamburg. 1998. Rac1-dependent actin filament organization in growth cones is necessary for beta1-integrin-mediated advance but not for growth on poly-D-lysine. J Neurobiol. 37:524-540.

Kuhn, T.B., P.J. Meberg, M.D. Brown, B.W. Bernstein, L.S. Minamide, J.R. Jensen, K. Okada, E.A. Soda, and J.R. Bamburg. 2000. Regulating actin dynamics in neuronal growth cones by ADF/cofilin and rho family GTPases. J Neurobiol. 44:126-144.

Kuipers-Upmeijer, J., A.E. de Jager, J.M. Hew, J.W. Snoek, and T.W. van Weerden. 2001. Primary lateral sclerosis: clinical, neurophysiological, and magnetic resonance findings. J Neurol Neurosurg Psychiatry. 71:615-620.

Kummer, J.L., P.K. Rao, and K.A. Heidenreich. 1997. Apoptosis induced by withdrawal of trophic factors is mediated by p38 mitogen-activated protein kinase. J Biol Chem. 272:20490-20494.

Kunda, P., G. Paglini, S. Quiroga, K. Kosik, and A. Caceres. 2001. Evidence for the involvement of Tiam1 in axon formation. J Neurosci. 21:2361-2372.

Kuner, R., A.J. Groom, I. Bresink, H.C. Kornau, V. Stefovska, G. Muller, B. Hartmann, K. Tschauner, S. Waibel, A.C. Ludolph, C. Ikonomidou, P.H. Seeburg, and L. Turski. 2005. Late-onset motoneuron disease caused by a functionally modified AMPA receptor subunit. Proc Natl Acad Sci U S A. 102:5826-5831.

Kunita, R., A. Otomo, H. Mizumura, K. Suzuki, J. Showguchi-Miyata, Y. Yanagisawa, S. Hadano, and J.E. Ikeda. 2004. Homo-oligomerization of ALS2 through its unique

224 carboxy-terminal regions is essential for the ALS2-associated Rab5 guanine nucleotide exchange activity and its regulatory function on endosome trafficking. J Biol Chem 279:38626-38635.

Kusaka, H. 1999. [Neuropathology of the motor neuron disease--Bunina body]. Rinsho Shinkeigaku. 39:65-66.

Kust, B.M., N. Brouwer, I.J. Mantingh, H.W. Boddeke, and J.C. Copray. 2003. Reduced p75NTR expression delays disease onset only in female mice of a transgenic model of familial amyotrophic lateral sclerosis. Amyotroph Lateral Scler Other Motor Neuron Disord. 4:100-105.

Kwon, Y.T., A. Gupta, Y. Zhou, M. Nikolic, and L.-H. Tsai. 2000. Regulation of N- cadherin-mediated adhesion by the p35-Cdk5 kinase. Curr. Biol. 10:363-372.

La Spada, A.R., Wilson, E.M., Lubahn, D.B., Harding, A.E., and K.H. Fischbeck. 1991. Androgen receptor gene mutations in X-linked spinal and bulbar muscular atrophy. Nature 352:77-79

Lacomblez, L., G. Bensimon, P.N. Leigh, P. Guillet, L. Powe, S. Durrleman, J.C. Delumeau, and V. Meininger. 1996. A confirmatory dose-ranging study of riluzole in ALS. ALS/Riluzole Study Group-II. Neurology. 47:S242-S250.

Lamarche, N., N. Tapon, L. Stowers, P.D. Burbelo, P. Aspenstrom, T. Bridges, J. Chant, and A. Hall. 1996. Rac and Cdc42 induce actin polymerization and G1 cell cycle progression independently of p65PAK and the JNK/SAPK MAP kinase cascade. Cell. 87:519-529.

Lamaze, C., T.H. Chuang, L.J. Terlecky, G.M. Bokoch, and S.L. Schmid. 1996. Regulation of receptor-mediated endocytosis by Rho and Rac. Nature. 382:177-179.

Lambrechts D., Storkebaum E., Morimoto M., Del-Favero J., Desmet F., Marklund S.L., Wyns S., Thijs V., Andersson J., van Marion I., Al-Chalabi A., Bornes S., Musson

225 R., Hansen V., Beckman L., Adolfsson R., Pall H.S., Prats H., Vermeire S., Rutgeerts P., Katayama S., Awata T., Leigh N., Lang-Lazdunski L., Dewerchin M., Shaw C., Moons L., Vlietinck R., Morrison K.E., Robberecht W., Van Broeckhoven C., Collen D., Andersen P.M., and P. Carmeliet. 2003. VEGF is a modifier of amyotrophic lateral sclerosis in mice and humans and protects motoneurons against ischemic death. Nat Genet. 33:383-394.

LaMonte, B.H., K.E. Wallace, B.A. Holloway, S.S. Shelly, J. Ascano, M. Tokito, T. Van Winkle, D.S. Howland, and E.L. Holzbaur. 2002. Disruption of dynein/dynactin inhibits axonal transport in motor neurons causing late-onset progressive degeneration. Neuron. 34:715-727.

Lamoureux, P., Z.F. Altun-Gultekin, C. Lin, J.A. Wagner, and S.R. Heidemann. 1997. Rac is required for growth cone function but not neurite assembly. J Cell Sci. 110 ( Pt 5):635-641.

Lampson, L.A., P.D. Kushner, and R.A. Sobel. 1990. Major histocompatibility complex antigen expression in the affected tissues in amyotrophic lateral sclerosis. Ann Neurol. 28:365-372.

Lanier, L.M., and F.B. Gertler. 2000. From Abl to actin: Abl tyrosine kinase and associated proteins in growth cone motility. Curr Opin Neurobiol. 10:80-87.

Lanius, R.A., H.B. Paddon, M. Mezei, R. Wagey, C. Krieger, S.L. Pelech, and C.A. Shaw. 1995. A role for amplified protein kinase C activity in the pathogenesis of amyotrophic lateral sclerosis. J Neurochem. 65:927-930.

Lanzetti, L., A. Palamidessi, L. Areces, G. Scita, and P.P. Di Fiore. 2004. Rab5 is a signalling GTPase involved in actin remodelling by receptor tyrosine kinases. Nature. 429:309-314.

Le Forestier, N., T. Maisonobe, A. Piquard, S. Rivaud, L. Crevier-Buchman, F. Salachas, P.F. Pradat, L. Lacomblez, and V. Meininger. 2001. Does primary lateral

226 sclerosis exist? A study of 20 patients and a review of the literature. Brain. 124:1989- 1999.

Lee, K.W., H.J. Kim, J.J. Sung, K.S. Park, and M. Kim. 2002. Defective neurite outgrowth in aphidicolin/cAMP-induced motor neurons expressing mutant Cu/Zn superoxide dismutase. Int J Dev Neurosci. 20:521-526.

Lee, M.S., Y.T. Kwon, M. Li, J. Peng, R.M. Friedlander, and L.H. Tsai. 2000. Neurotoxicity induces cleavage of p35 to p25 by calpain. Nature. 405:360-364.

Lee, Y.S., Y.S. Kang, J.S. Lee, S. Nicolova, and J.A. Kim. 2004. Involvement of NADPH oxidase-mediated generation of reactive oxygen species in the apototic cell death by capsaicin in HepG2 human hepatoma cells. Free Radic Res. 38:405-412.

Leeuwen, F.N., H.E. Kain, R.A. Kammen, F. Michiels, O.W. Kranenburg, and J.G. Collard. 1997. The guanine nucleotide exchange factor Tiam1 affects neuronal morphology; opposing roles for the small GTPases Rac and Rho. J Cell Biol. 139:797- 807.

Lefebvre, S., L. Burglen, S. Reboullet, O. Clermont, P. Burlet, L. Viollet, B. Benichou, C. Cruaud, P. Millasseau, M. Zeviani, and et al. 1995. Identification and characterization of a spinal muscular atrophy-determining gene. Cell. 80:155-165.

Leigh, P.N., and O. Garofolo. 1995. The molecular pathology of motor neurone disease. In Motor neurone disease. M. Swash and P.N. Leigh, editors. Springer Verlag, London. 139-161.

Leigh, P.N., H. Whitwell, O. Garofalo, J. Buller, M. Swash, J.E. Martin, J.M. Gallo, R.O. Weller, and B.H. Anderton. 1991. Ubiquitin-immunoreactive intraneuronal inclusions in amyotrophic lateral sclerosis. Morphology, distribution, and specificity. Brain. 114 ( Pt 2):775-788.

Lemmon, M.A., and K.M. Ferguson. 2000. Signal-dependent membrane targeting by

227 pleckstrin homology (PH) domains. Biochem J. 350 Pt 1:1-18.

Letourneau, P.C. 1983. Differences in the organization of actin in the growth cones compared with the neurites of cultured neurons from chick embryos. J Cell Biol. 97:963-973.

Leung, C.L., C.Z. He, P. Kaufmann, S.S. Chin, A. Naini, R.K. Liem, H. Mitsumoto, and A.P. Hays. 2004. A pathogenic peripherin gene mutation in a patient with amyotrophic lateral sclerosis. Brain Pathol. 14:290-296.

Li, G.H., C.D. Qin, and M.H. Li. 1994. On the mechanisms of growth cone locomotion: modeling and computer simulation. J Theor Biol. 169:355-362.

Li, M., V.O. Ona, C. Guegan, M. Chen, V. Jackson-Lewis, L.J. Andrews, A.J. Olszewski, P.E. Stieg, J.P. Lee, S. Przedborski, and R.M. Friedlander. 2000. Functional role of caspase-1 and caspase-3 in an ALS transgenic mouse model. Science. 288:335- 339.

Li, Z., Van Aelst, L. and H. T. Cline. 2000b. Rho GTPases regulate distinct aspects of dendritic arbor growth in xenopus central neurons. Nat Neurosci. 3:217-225.

Lin, C.-L.G., L.A. Bristol, L. Jin, M. Dykes-Hoberg, T. Crawford, L. Clawson, and J.D. Rothstein. 1998. Aberrant RNA processing in a neurodegenerative disease: the cause of absent EAAT2, a glutamate transporter, in Amyotrophic Lateral Sclerosis. Neuron. 20:589-602.

Lino, M.M., C. Schneider, and P. Caroni. 2002. Accumulation of SOD1 mutants in postnatal motoneurons does not cause motoneuron pathology or motoneuron disease. J.Neurosci. 22:4825-4832.

Liu, J., C. Lillo, P.A. Jonsson, C.V. Velde, C.M. Ward, T.M. Miller, J.R. Subramaniam, J.D. Rothstein, S. Marklund, P.M. Andersen, T. Brannstrom, O. Gredal, P.C. Wong, D.S. Williams, and D.W. Cleveland. 2004. Toxicity of familial ALS-linked SOD1

228 mutants from selective recruitment to spinal mitochondria. Neuron. 43:5-17.

Liu, R., B. Li, S.W. Flanagan, L.W. Oberley, D. Gozal, and M. Qiu. 2002. Increased mitochondrial antioxidative activity or decreased oxygen free radical propagation prevent mutant SOD1-mediated motor neuron cell death and increase amyotrophic lateral sclerosis-like transgenic mouse survival. J Neurochem. 80:488-500.

Liu, R.G., J.S. Althaus, B.R. Ellerbrock, D.A. Becker, and M.E. Gurney. 1998a. Enhanced oxygen radical production in a transgenic mouse model of familial amyotrophic lateral sclerosis. Ann.Neurol. 44:763-770.

Liu, X., H. Wang, M. Eberstadt, A. Schnuchel, E.T. Olejniczak, R.P. Meadows, J.M. Schkeryantz, D.A. Janowick, J.E. Harlan, E.A. Harris, D.E. Staunton, and S.W. Fesik. 1998b. NMR structure and mutagenesis of the N-terminal Dbl homology domain of the nucleotide exchange factor Trio. Cell. 95:269-277.

Lo Nigro, C., R. Cusano, M. Scaranari, R. Cinti, P. Forabosco, V.B. Morra, G. De Michele, L. Santoro, S. Davies, J. Hurst, M. Devoto, R. Ravazzolo, and M. Seri. 2000. A refined physical and transcriptional map of the SPG9 locus on 10q23.3-q24.2. Eur J Hum Genet. 8:777-782.

Lundquist, E.A., P.W. Reddien, E. Hartwieg, H.R. Horvitz, and C.I. Bargmann. 2001. Three C. elegans Rac proteins and several alternative Rac regulators control axon guidance, cell migration and apoptotic cell phagocytosis. Development. 128:4475-4488.

Luo, L., T.K. Hensch, L. Ackerman, S. Barbel, L.Y. Jan, and Y.N. Jan. 1996. Differential effects of the Rac GTPase on Purkinje cell axons and dendritic trunks and spines. Nature. 379:837-840.

Luo, L., Y.J. Liao, L.Y. Jan, and Y.N. Jan. 1994. Distinct morphogenetic functions of similar small GTPases: Drosophila Drac1 is involved in axonal outgrowth and myoblast fusion. Genes Dev. 8:1787-1802.

229 Machesky, L.M., and R.H. Insall. 1999. Signaling to actin dynamics. J Cell Biol. 146:267-272.

Mack, C.P., A.V. Somlyo, M. Hautmann, A.P. Somlyo, and G.K. Owens. 2001. Smooth muscle differentiation marker gene expression is regulated by RhoA-mediated actin polymerization. J Biol Chem. 276:341-347.

Maekawa, M., T. Ishizaki, S. Boku, N. Watanabe, A. Fujita, A. Iwamatsu, T. Obinata, K. Ohashi, K. Mizuno, and S. Narumiya. 1999. Signaling from Rho to the actin cytoskeleton through protein kinases ROCK and LIM-kinase. Science. 285:895-898.

Maekawa, S., S. Al-Sarraj, M. Kibble, S. Landau, J. Parnavelas, D. Cotter, I. Everall, and P.N. Leigh. 2004. Cortical selective vulnerability in motor neuron disease: a morphometric study. Brain. 127:1237-1251.

Maihofner, C., S. Probst-Cousin, M. Bergmann, W. Neuhuber, B. Neundorfer, and D. Heuss. 2003. Expression and localization of cyclooxygenase-1 and -2 in human sporadic amyotrophic lateral sclerosis. Eur J Neurosci. 18:1527-1534.

Malecz, N., P.C. McCabe, C. Spaargaren, R. Qiu, Y. Chuang, and M. Symons. 2000. Synaptojanin 2, a novel Rac1 effector that regulates clathrin-mediated endocytosis. Curr Biol. 10:1383-1386.

Mallavarapu, A., and T. Mitchison. 1999. Regulated actin cytoskeleton assembly at filopodium tips controls their extension and retraction. J Cell Biol. 146:1097-1106.

Manser, E., H.Y. Huang, T.H. Loo, X.Q. Chen, J.M. Dong, T. Leung, and L. Lim. 1997. Expression of constitutively active alpha-PAK reveals effects of the kinase on actin and focal complexes. Mol Cell Biol. 17:1129-1143.

Martin, L.J. 1999. Neuronal death in amyotrophic lateral sclerosis is apoptosis: Possible contribution of a programmed cell death mechanism. J.Neuropathol.Exp.Neurol. 58:459-471.

230 Martin, N., J. Jaubert, P. Gounon, E. Salido, G. Haase, M. Szatanik, and J.L. Guenet. 2002. A missense mutation in Tbce causes progressive motor neuronopathy in mice. Nat. Genet. 32:443-447.

Martinez Murillo, F., H. Kobayashi, E. Pegoraro, G. Galluzzi, G. Creel, C. Mariani, E. Farina, E. Ricci, G. Alfonso, R.M. Pauli, and E.P. Hoffman. 1999. Genetic localization of a new locus for recessive familial spastic paraparesis to 15q13-15. Neurology. 53:50- 56.

Martinez-Arca, S., S. Coco, G. Mainguy, U. Schenk, P. Alberts, P. Bouille, M. Mezzina, A. Prochiantz, M. Matteoli, D. Louvard, and T. Galli. 2001. A common exocytotic mechanism mediates axonal and dendritic outgrowth. J Neurosci. 21:3830- 3838.

Matsumoto, S., S. Goto, H. Kusaka, T. Imai, N. Murakami, Y. Hashizume, H. Okazaki, and A. Hirano. 1993. Ubiquitin-positive inclusion in anterior horn cells in subgroups of motor neuron diseases: A comparative study of adult-onset amyotrophic lateral sclerosis, juvenile amyotrophic lateral sclerosis and Werdnig-Hoffmann disease. J.Neurol.Sci. 115:208-213.

Matsuo, N., M. Hoshino, M. Yoshizawa, and Y. Nabeshima. 2002. Characterization of STEF, a guanine nucleotide exchange factor for Rac1, required for neurite growth. J Biol Chem. 277:2860-2868.

May, R.C., and L.M. Machesky. 2001. Phagocytosis and the actin cytoskeleton. J Cell Sci. 114:1061-1077.

May, V., M.R. Schiller, B.A. Eipper, and R.E. Mains. 2002. Kalirin Dbl-homology guanine nucleotide exchange factor 1 domain initiates new axon outgrowths via RhoG- mediated mechanisms. J Neurosci. 22:6980-6990.

McDermott, C.J., A.J. Grierson, J.D. Wood, M. Bingley, S.B. Wharton, K.M. Bushby, and P.J. Shaw. 2003. Hereditary spastic paraparesis: disrupted intracellular transport

231 associated with spastin mutation. Ann Neurol. 54:748-759.

McLauchlan, H., J. Newell, N. Morrice, A. Osborne, M. West, and E. Smythe. 1998. A novel role for Rab5-GDI in ligand sequestration into clathrin-coated pits. Curr Biol. 8:34-45.

McPherson, P.S. 2002. The endocytic machinery at an interface with the actin cytoskeleton: a dynamic, hip intersection. Trends Cell Biol. 12:312-315.

McWhorter, M.L., U.R. Monani, A.H. Burghes, and C.E. Beattie. 2003. Knockdown of the survival motor neuron (Smn) protein in zebrafish causes defects in motor axon outgrowth and pathfinding. J Cell Biol. 162:919-931.

Meller, N., M. Irani-Tehrani, W.B. Kiosses, M.A. Del Pozo, and M.A. Schwartz. 2002. Zizimin1, a novel Cdc42 activator, reveals a new GEF domain for Rho proteins. Nat Cell Biol. 4:639-647.

Mellor, H., P. Flynn, C.D. Nobes, A. Hall, and P.J. Parker. 1998. PRK1 is targeted to endosomes by the small GTPase, RhoB. J. Biol. Chem. 273:4811-4814.

Menzies, F.M., A.J. Grierson, M.R. Cookson, P.R. Heath, J. Tomkins, D.A. Figlewicz, P.G. Ince, and P.J. Shaw. 2002. Selective loss of neurofilament expression in Cu/Zn superoxide dismutase (SOD1) linked amyotrophic lateral sclerosis. J. Neurochem. 82:1118-1128.

Mersiyanova, I.V., A.V. Perepelov, A.V. Polyakov, V.F. Sitnikov, E.L. Dadali, R.B. Oparin, A.N. Petrin, and O.V. Evgrafov. 2000. A new variant of Charcot-Marie-Tooth disease type 2 is probably the result of a mutation in the neurofilament-light gene. Am.J.Hum.Genet. 67:37-46.

Michiels, F., G.G. Habets, J.C. Stam, R.A. van der Kammen, and J.G. Collard. 1995. A role for Rac in Tiam1-induced membrane ruffling and invasion. Nature. 375:338-340.

232 Michiels, F., J.C. Stam, P.L. Hordijk, R.A. van der Kammen, L. Ruuls-Van Stalle, C.A. Feltkamp, and J.G. Collard. 1997. Regulated membrane localization of Tiam1, mediated by the NH2-terminal pleckstrin homology domain, is required for Rac-dependent membrane ruffling and C-Jun NH2-terminal kinase activation. J Cell Biol. 137:387-398.

Mielke, K., and T. Herdegen. 2001. JNK and p38 stresskinases -degenerative effectors of signal-transduction-cascades in the nervous system. Prog. Neurobiol. 61:45-60.

Migheli, A., C. Atzori, R. Piva, M. Tortarolo, M. Girelli, D. Schiffer, and C. Bendotti. 1999. Lack of apoptosis in mice with ALS. Nat Med. 5:966-967.

Migheli, A., T. Pezzulo, A. Attanasio, and D. Schiffer. 1993. Peripherin immunoreactive structures in amyotrophic lateral sclerosis. Lab Invest. 68:185-191.

Migheli, A., R. Piva, C. Atzori, D. Troost, and D. Schiffer. 1997. c-Jun, JNK/SAPK kinases and transcription factor NF-kappa B are selectively activated in astrocytes, but not motor neurons, in amyotrophic lateral sclerosis. J Neuropathol Exp Neurol. 56:1314-1322.

Miki, H., H. Yamaguchi, S. Suetsugu, and T. Takenawa. 2000. IRSp53 is an essential intermediate between Rac and WAVE in the regulation of membrane ruffling. Nature. 408:732-735.

Miller, C.C.J., S. Ackerley, J. Brownlees, A.J. Grierson, N.J.O. Jacobsen, and P. Thornhill. 2002. Axonal transport of neurofilaments in normal and disease states. Cell Mol. Life Sci. 59:323-330.

Miller, R.G., D. Moore, L.A. Young, C. Armon, R.J. Barohn, M.B. Bromberg, W.W. Bryan, D.F. Gelinas, M.C. Mendoza, H.E. Neville, G.J. Parry, J.H. Petajan, J.M. Ravits, S.P. Ringel, and M.A. Ross. 1996. Placebo-controlled trial of gabapentin in patients with amyotrophic lateral sclerosis. WALS Study Group. Western Amyotrophic Lateral Sclerosis Study Group. Neurology. 47:1383-1388.

233 Miller, R.G., D.H. Moore, 2nd, D.F. Gelinas, V. Dronsky, M. Mendoza, R.J. Barohn, W. Bryan, J. Ravits, E. Yuen, H. Neville, S. Ringel, M. Bromberg, J. Petajan, A.A. Amato, C. Jackson, W. Johnson, R. Mandler, P. Bosch, B. Smith, M. Graves, M. Ross, E.J. Sorenson, P. Kelkar, G. Parry, and R. Olney. 2001. Phase III randomized trial of gabapentin in patients with amyotrophic lateral sclerosis. Neurology. 56:843-848.

Minden, A., A. Lin, F.X. Claret, A. Abo, and M. Karin. 1995. Selective activation of the JNK signaling cascade and c-Jun transcriptional activity by the small GTPases Rac and Cdc42Hs. Cell. 81:1147-1157.

Mitchison, T., and M. Kirschner. 1988. Cytoskeletal dynamics and nerve growth. Neuron. 1:761-772.

Mitsumoto, H., A.L. Ferut, K. Kurahashi, and I.G. McQuarrie. 1990. Impairment of retrograde axonal transport in wobbler mouse motor neuron disease. Muscle Nerve. 13:121-126.

Mitsumoto, H., and P. Gambetti. 1986. Impaired slow axonal transport in wobbler mouse motor neuron disease. Ann Neurol. 19:36-43.

Moon, S.Y., and Y. Zheng. 2003. Rho GTPase-activating proteins in cell regulation. Trends Cell Biol. 13:13-22.

Moulard, B., A. Sefiani, A. Laamri, A. Malafosse, and W. Camu. 1996. Apolipoprotein E genotyping in sporadic amyotrophic lateral sclerosis: evidence for a major influence on the clinical presentation and prognosis. J Neurol Sci. 139 Suppl:34-37.

Mourelatos, Z., N.K. Gonatas, A. Stieber, M.E. Gurney, and M.C. Dal Canto. 1996. The Golgi apparatus of spinal cord motor neurons in transgenic mice expressing mutant Cu,Zn superoxide dismutase becomes fragmented in early, preclinical stages of the disease. Proc. Natl. Acad.Sci. USA. 93:5472-5477.

Movilla, N., and X.R. Bustelo. 1999. Biological and regulatory properties of Vav-3, a

234 new member of the Vav family of oncoproteins. Mol Cell Biol. 19:7870-7885.

Muglia, M., C. Criscuolo, A. Magariello, G. De Michele, V. Scarano, P. D'Adamo, G. Ambrosio, A.L. Gabriele, A. Patitucci, R. Mazzei, F.L. Conforti, T. Sprovieri, L. Morgante, A. Epifanio, P. La Spina, P. Valentino, P. Gasparini, A. Filla, and A. Quattrone. 2004. Narrowing of the critical region in autosomal recessive spastic paraplegia linked to the SPG5 locus. Neurogenetics. 5:49-54.

Mulder, D.W., L.T. Kurland, K.P. Offord, and C.M. Beard. 1986. Familial adult motor neuron disease: amyotrophic lateral sclerosis. Neurology. 36:511-517.

Munch, C., R. Sedlmeier, T. Meyer, V. Homberg, A.D. Sperfeld, A. Kurt, J. Prudlo, G. Peraus, C.O. Hanemann, G. Stumm, and A.C. Ludolph. 2004. Point mutations of the p150 subunit of dynactin (DCTN1) gene in ALS. Neurology. 63:724-726.

Munoz, D.G., C. Greene, D.P. Perl, and D.J. Selkoe. 1988. Accumulation of phosphorylated neurofilaments in anterior horn motoneurons of amyotrophic lateral sclerosis patients. J Neuropathol Exp Neurol. 47:9-18.

Murayama, S., K. Inoue, H. Kawakami, T.W. Bouldin, and K. Suzuki. 1991. A unique pattern of astrocytosis in the primary motor area in amyotrophic lateral sclerosis. Acta Neuropathol (Berl). 82:456-461.

Murayama, S., Y. Ookawa, H. Mori, I. Nakano, Y. Ihara, S. Kuzuhara, and M. Tomonaga. 1989. Immunocytochemical and ultrastructural study of Lewy body-like hyaline inclusions in familial amyotrophic lateral sclerosis. Acta Neuropathol (Berl). 78:143-152.

Murga, C., M. Zohar, H. Teramoto, and J.S. Gutkind. 2002. Rac1 and RhoG promote cell survival by the activation of PI3K and Akt, independently of their ability to stimulate JNK and NF-kappaB. Oncogene. 21:207-216.

Murphy, A.M., and D.J. Montell. 1996. Cell type-specific roles for Cdc42, Rac, and

235 RhoL in Drosophila oogenesis. J Cell Biol. 133:617-630.

Murphy, C., R. Saffrich, M. Grummt, H. Gournier, V. Rybin, M. Rubino, P. Auvinen, A. Lutcke, R.G. Parton, and M. Zerial. 1996. Endosome dynamics regulated by a Rho protein. Nature. 384:427-432.

Nagao, M., S. Kato, M. Oda, and S. Hirai. 1998. Decrease of protein kinase C in the spinal motor neurons of amyotrophic lateral sclerosis. Acta Neuropathol (Berl). 96:52- 56.

Nagata, K., M. Driessens, N. Lamarche, J.L. Gorski, and A. Hall. 1998. Activation of G1 progression, JNK mitogen-activated protein kinase, and actin filament assembly by the exchange factor FGD1. J Biol Chem. 273:15453-15457.

Nakayama, A.Y., M.B. Harms, and L. Luo. 2000. Small GTPases Rac and Rho in the maintenance of dendritic spines and branches in hippocampal pyramidal neurons. J Neurosci. 20:5329-5338.

Nelson, M., and M. McClelland. 1992. Use of DNA methyltransferase/endonuclease enzyme combinations for megabase mapping of . Methods Enzymol. 216:279-303.

Nguyen, M.D., R.C. Lariviere, and J.P. Julien. 2001. Deregulation of Cdk5 in a mouse model of ALS: Toxicity alleviated by perikaryal neurofilament inclusions. Neuron. 30:135-147.

Nguyen, M.D., R.C. Larivière, and J.P. Julien. 2000. Reduction of axonal caliber does not alleviate motor neuron disease caused by mutant superoxide dismutase 1. Proc.Natl.Acad.Sci.USA. 97:12306-12311.

Nicoziani, P., F. Vilhardt, A. Llorente, L. Hilout, P.J. Courtoy, K. Sandvig, and B. van Deurs. 2000. Role for dynamin in late endosome dynamics and trafficking of the cation- independent mannose 6-phosphate receptor. Mol Biol Cell. 11:481-495.

236 Niebroj-Dobosz, I., D. Dziewulska, and H. Kwiecinski. 2004. Oxidative damage to proteins in the spinal cord in amyotrophic lateral sclerosis (ALS). Folia Neuropathol. 42:151-156.

Niebroj-Dobosz, I., Z. Jamrozik, P. Janik, I. Hausmanowa-Petrusewicz, and H. Kwiecinski. 1999. Anti-neural antibodies in serum and cerebrospinal fluid of amyotrophic lateral sclerosis (ALS) patients. Acta Neurol Scand. 100:238-243.

Nielsen, E., F. Severin, J.M. Backer, A.A. Hyman, and M. Zerial. 1999. Rab5 regulates motility of early endosomes on microtubules. Nat Cell Biol. 1:376-382.

Nikolic, M. 2002. The role of Rho GTPases and associated kinases in regulating neurite outgrowth. Int J Biochem Cell Biol. 34:731-745.

Nikolic, M., H. Dudek, Y.T. Kwon, Y.F.M. Ramos, and L.H. Tsai. 1996. The cdk5/p35 kinase is essential for neurite outgrowth during neuronal differentiation. Genes Dev. 10:816-825.

Nimnual, A.S., L.J. Taylor, and D. Bar-Sagi. 2003. Redox-dependent downregulation of Rho by Rac. Nat Cell Biol. 5:236-241.

Nimnual, A.S., B.A. Yatsula, and D. Bar-Sagi. 1998. Coupling of Ras and Rac guanosine triphosphatases through the Ras exchanger Sos. Science. 279:560-563.

Nishimura, A.L., M. Mitne-Neto, H.C. Silva, A. Richieri-Costa, S. Middleton, D. Cascio, F. Kok, J.R. Oliveira, T. Gillingwater, J. Webb, P. Skehel, and M. Zatz. 2004. A mutation in the vesicle-trafficking protein VAPB causes late-onset spinal muscular atrophy and amyotrophic lateral sclerosis. Am J Hum Genet. 75:822-831.

Niwa, J., S. Ishigaki, N. Hishikawa, M. Yamamoto, M. Doyu, S. Murata, K. Tanaka, N. Taniguchi, and G. Sobue. 2002. Dorfin ubiquitylates mutant SOD1 and prevents mutant SOD1-mediated neurotoxicity. J Biol Chem. 277:36793-36798.

237 Nobes, C.D., and A. Hall. 1995. Rho, rac, and cdc42 GTPases regulate the assembly of multimolecular focal complexes associated with actin stress fibers, lamellipodia, and filopodia. Cell. 81:53-62.

Obermeier, A., S. Ahmed, E. Manser, S.C. Yen, C. Hall, and L. Lim. 1998. PAK promotes morphological changes by acting upstream of Rac. Embo J. 17:4328-4339.

Okado-Matsumoto, A., and I. Fridovich. 2002. Amyotrophic lateral sclerosis: A proposed mechanism. Proc.Natl.Acad.Sci.USA. 99:9010-9014.

Okamoto, K., S. Hirai, M. Amari, M. Watanabe, and A. Sakurai. 1993. Bunina bodies in amyotrophic lateral sclerosis immunostained with rabbit anti-cystatin C serum. Neurosci Lett. 162:125-128.

Okamoto, K., S. Hirai, M. Shoji, Y. Senoh, and T. Yamazaki. 1990. Axonal swellings in the corticospinal tracts in amyotrophic lateral sclerosis. Acta Neuropathol (Berl). 80:222-226.

Olson, M.F., N.G. Pasteris, J.L. Gorski, and A. Hall. 1996. Faciogenital dysplasia protein (FGD1) and Vav, two related proteins required for normal embryonic development, are upstream regulators of Rho GTPases. Curr Biol. 6:1628-1633.

Ono, K., and J. Han. 2000. The p38 signal transduction pathway: activation and function. Cell Signal. 12:1-13.

Oosthuyse, B., L. Moons, E. Storkebaum, H. Beck, D. Nuyens, K. Brusselmans, J. Van Dorpe, P. Hellings, M. Gorselink, S. Heymans, G. Theilmeier, M. Dewerchin, V. Laudenbach, P. Vermylen, H. Raat, T. Acker, V. Vleminckx, L. Van Den Bosch, N. Cashman, H. Fujisawa, M.R. Drost, R. Sciot, F. Bruyninckx, D.J. Hicklin, C. Ince, P. Gressens, F. Lupu, K.H. Plate, W. Robberecht, J.M. Herbert, D. Collen, and P. Carmeliet. 2001. Deletion of the hypoxia-response element in the vascular endothelial growth factor promoter causes motor neuron degeneration. Nat. Genet. 28:131-138.

238 Osen-Sand, A., J.K. Staple, E. Naldi, G. Schiavo, O. Rossetto, S. Petitpierre, A. Malgaroli, C. Montecucco, and S. Catsicas. 1996. Common and distinct fusion proteins in axonal growth and transmitter release. J Comp Neurol. 367:222-234.

Otomo, A., S. Hadano, T. Okada, H. Mizumura, R. Kunita, H. Nishijima, J. Showguchi- Miyata, Y. Yanagisawa, E. Kohiki, E. Suga, M. Yasuda, H. Osuga, T. Nishimoto, S. Narumiya, and J.E. Ikeda. 2003. ALS2, a novel guanine nucleotide exchange factor for the small GTPase Rab5, is implicated in endosomal dynamics. Hum. Mol. Genet. 12:1671-1687.

Pardo, C.A., Z. Xu, D.R. Borchelt, D.L. Price, S.S. Sisodia, and D.W. Cleveland. 1995. Superoxide dismutase is an abundant component in cell bodies, dendrites, and axons of motor neurons and in a subset of other neurons. Proc.Natl.Acad.Sci.USA. 92:954-958.

Parysek, L.M., and R.D. Goldman. 1988. Distribution of a novel 57kDa intermediate filament IF protein in the nervous system. J.Neurosci. 8:555-563.

Pasinelli, P., M.E. Belford, N. Lennon, B.J. Bacskai, B.T. Hyman, D. Trotti, and R.H. Brown, Jr. 2004. Amyotrophic lateral sclerosis-associated SOD1 mutant proteins bind and aggregate with Bcl-2 in spinal cord mitochondria. Neuron. 43:19-30.

Pasinelli, P., M.K. Houseweart, R.H. Brown, Jr., and D.W. Cleveland. 2000. Caspase-1 and-3 are Sequentially activated in motor neuron death in Cu,Zn superoxide dismutase- mediated familial amyotrophic lateral sclerosis. Proc. Natl. Acad. Sci. USA. 97:13901- 13906.

Patel, H., H. Cross, C. Proukakis, R. Hershberger, P. Bork, F.D. Ciccarelli, M.A. Patton, V.A. McKusick, and A.H. Crosby. 2002. SPG20 is mutated in Troyer syndrome, an hereditary spastic paraplegia. Nat Genet. 31:347-348.

Patel, H., P.E. Hart, T.T. Warner, R.S. Houlston, M.A. Patton, S. Jeffery, and A.H. Crosby. 2001. The Silver syndrome variant of hereditary spastic paraplegia maps to chromosome 11q12-q14, with evidence for genetic heterogeneity within this subtype.

239 Am J Hum Genet. 69:209-215.

Pedersen, W.A., W. Fu, J.N. Keller, W.R. Markesbery, S. Appel, R.G. Smith, E. Kasarskis, and M.P. Mattson. 1998. Protein modification by the lipid peroxidation product 4-hydroxynonenal in the spinal cords of amyotrophic lateral sclerosis patients. Ann Neurol. 44:819-824.

Pehar, M., P. Cassina, M.R. Vargas, R. Castellanos, L. Viera, J.S. Beckman, A.G. Estevez, and L. Barbeito. 2004. Astrocytic production of nerve growth factor in motor neuron apoptosis: implications for amyotrophic lateral sclerosis. J Neurochem. 89:464- 473.

Peluffo, H., J.J. Shacka, K. Ricart, C.G. Bisig, L. Martinez-Palma, O. Pritsch, A. Kamaid, J.P. Eiserich, J.P. Crow, L. Barbeito, and A.G. Estevez. 2004. Induction of motor neuron apoptosis by free 3-nitro-l-tyrosine. J. Neurochem. 89:602-612.

Penzes, P., R.C. Johnson, M.R. Alam, V. Kambampati, R.E. Mains, and B.A. Eipper. 2000. An isoform of kalirin, a brain-specific GDP/GTP exchange factor, is enriched in the postsynaptic density fraction. J Biol Chem. 275:6395-6403.

Penzes, P., R.C. Johnson, V. Kambampati, R.E. Mains, and B.A. Eipper. 2001. Distinct roles for the two Rho GDP/GTP exchange factor domains of kalirin in regulation of neurite growth and neuronal morphology. J Neurosci. 21:8426-8434.

Perkinton, M.S., C.L. Standen, K.F. Lau, S. Kesavapany, H.L. Byers, M. Ward, D.M. McLoughlin, and C.C. Miller. 2004. The c-Abl tyrosine kinase phosphorylates the Fe65 adaptor protein to stimulate Fe65/amyloid precursor protein nuclear signalling. J. Biol. Chem. 279:22084-22091.

Plaitakis, A., and J.T. Caroscio. 1987. Abnormal glutamate metabolism in amyotrophic lateral sclerosis. Ann Neurol. 22:575-579.

Pompl, P.N., L. Ho, M. Bianchi, T. McManus, W. Qin, and G.M. Pasinetti. 2003. A

240 therapeutic role for cyclooxygenase-2 inhibitors in a transgenic mouse model of amyotrophic lateral sclerosis. Faseb J. 17:725-727.

Pramatarova, A., J. Laganière, J. Roussel, K. Brisebois, and G.A. Rouleau. 2001. Neuron-specific expression of mutant superoxide dismutase 1 in transgenic mice does not lead to motor impairment. J.Neurosci. 21:3369-3374.

Price, L.S., J. Leng, M.A. Schwartz, and G.M. Bokoch. 1998. Activation of Rac and Cdc42 by integrins mediates cell spreading. Mol Biol Cell. 9:1863-1871.

Prieto-Sanchez, R.M., and X.R. Bustelo. 2003. Structural basis for the signaling specificity of RhoG and Rac1 GTPases. J Biol Chem. 278:37916-37925.

Pringle, C.E., A.J. Hudson, D.G. Munoz, J.A. Kiernan, W.F. Brown, and G.C. Ebers. 1992. Primary lateral sclerosis. Clinical features, neuropathology and diagnostic criteria. Brain. 115 ( Pt 2):495-520.

Puls, I., C. Jonnakuty, B.H. LaMonte, E.L. Holzbaur, M. Tokito, E. Mann, M.K. Floeter, K. Bidus, D. Drayna, S.J. Oh, R.H. Brown, C.L. Ludlow, and K.H. Fischbeck. 2003. Mutant dynactin in motor neuron disease. Nat. Genet. 33:455-456.

Puls, I., S.J. Oh, C.J. Sumner, K.E. Wallace, M.K. Floeter, E.A. Mann, W.R. Kennedy, G. Wendelschafer-Crabb, A. Vortmeyer, R. Powers, K. Finnegan, E.L. Holzbaur, K.H. Fischbeck, and C.L. Ludlow. 2005. Distal spinal and bulbar muscular atrophy caused by dynactin mutation. Ann Neurol. 57:687-694.

Qualmann, B., M.M. Kessels, and R.B. Kelly. 2000. Molecular links between endocytosis and the actin cytoskeleton. J Cell Biol. 150:F111-F116.

Raftopoulou, M., and A. Hall. 2004. Cell migration: Rho GTPases lead the way. Dev Biol. 265:23-32.

Rainier, S., J.H. Chai, D. Tokarz, R.D. Nicholls, and J.K. Fink. 2003. NIPA1 gene

241 mutations cause autosomal dominant hereditary spastic paraplegia (SPG6). Am J Hum Genet. 73:967-971.

Rakhit, R., P. Cunningham, A. Furtos-Matei, S. Dahan, X.F. Qi, J. Crow, N.R. Cashman, L.H. Kondejewski, and A. Chakrabartty. 2002. Oxidation-induced misfolding and aggregation of superoxide dismutase and Its implications for amyotrophic lateral sclerosis. J. Biol. Chem. 277:47551-47556.

Raoul, C., A. Estevez, H. Nishimune, D. Cleveland, O. deLapeyriere, C. Henderson, G. Haase, and B. Pettmann. 2002. Motoneuron Death Triggered by a Specific Pathway Downstream of Fas. Potentiation by ALS-Linked SOD1 Mutations. Neuron. 35:1067- 1083.

Rathke-Hartlieb, S., V.C. Schmidt, H. Jockusch, T. Schmitt-John, and J.W. Bartsch. 1999. Spatiotemporal progression of neurodegeneration and glia activation in the wobbler neuropathy of the mouse. Neuroreport. 10:3411-3416.

Ray, S.S., and P.T. Lansbury, Jr. 2004. A possible therapeutic target for Lou Gehrig's disease. Proc Natl Acad Sci U S A. 101:5701-5702.

Reaume, A.B., J.L. Elliott, E.K. Hoffman, N.W. Kowall, R.J. Ferrante, D.F. Siwek, H.M. Wilcox, D.G. Flood, M.F. Beal, R.H.J. Brown, R.W. Scott, and W.D. Snider. 1996. Motor neurons in Cu/Zn superoxide dismutase-deficient mice develop normally but exhibit enhanced cell death after axotomy. Nature Genet. 13:43-47.

Rebecchi, M.J., and S. Scarlata. 1998. Pleckstrin homology domains: a common fold with diverse functions. Annu Rev Biophys Biomol Struct. 27:503-528.

Reid, E., A.M. Dearlove, O. Osborn, M.T. Rogers, and D.C. Rubinsztein. 2000. A locus for autosomal dominant "pure" hereditary spastic paraplegia maps to chromosome 19q13. Am J Hum Genet. 66:728-732.

Reid, E., A.M. Dearlove, M. Rhodes, and D.C. Rubinsztein. 1999a. A new locus for

242 autosomal dominant "pure" hereditary spastic paraplegia mapping to chromosome 12q13, and evidence for further genetic heterogeneity. Am J Hum Genet. 65:757-763.

Reid, E., A.M. Dearlove, M.L. Whiteford, M. Rhodes, and D.C. Rubinsztein. 1999b. Autosomal dominant spastic paraplegia: refined SPG8 locus and additional genetic heterogeneity. Neurology. 53:1844-1849.

Reid, E., M. Kloos, A. Ashley-Koch, L. Hughes, S. Bevan, I.K. Svenson, F.L. Graham, P.C. Gaskell, A. Dearlove, M.A. Pericak-Vance, D.C. Rubinsztein, and D.A. Marchuk. 2002. A kinesin heavy chain (KIF5A) mutation in hereditary spastic paraplegia (SPG10). Am J Hum Genet. 71:1189-1194.

Reiner, A., L. Medina, G. Figueredo-Cardenas, and S. Anfinson. 1995. Brainstem motoneuron pools that are selectively resistant in amyotrophic lateral sclerosis are preferentially enriched in parvalbumin: evidence from monkey brainstem for a calcium- mediated mechanism in sporadic ALS. Exp Neurol. 131:239-250.

Ren, X.D., and M.A. Schwartz. 2000. Determination of GTP loading on Rho. Methods Enzymol. 325:264-272.

Ridley, A.J., and A. Hall. 1992. The small GTP-binding protein rho regulates the assembly of focal adhesions and actin stress fibers in response to growth factors. Cell. 70:389-399.

Ridley, A.J., H.F. Paterson, C.L. Johnston, D. Diekmann, and A. Hall. 1992. The small GTP-binding protein rac regulates growth factor-induced membrane ruffling. Cell. 70:401-410.

Ringel, S.P., J.R. Murphy, M.K. Alderson, W. Bryan, J.D. England, R.G. Miller, J.H. Petajan, S.A. Smith, R.I. Roelofs, F. Ziter, and et al. 1993. The natural history of amyotrophic lateral sclerosis. Neurology. 43:1316-1322.

Robberecht, W. 2000. Oxidative stress in amyotrophic lateral sclerosis. J Neurol. 247

243 Suppl 1:I1-I6.

Robinson, M.J., and M.H. Cobb. 1997. Mitogen-activated protein kinase pathways. Curr Opin Cell Biol. 9:180-186.

Rocco, P., M. Vainzof, S.C. Froehner, M.F. Peters, S.K. Marie, M.R. Passos-Bueno, and M. Zatz. 2000. Brazilian family with pure autosomal dominant spastic paraplegia maps to 8q: analysis of muscle beta 1 syntrophin. Am J Med Genet. 92:122-127.

Rodrigues, N.R., N. Owen, K. Talbot, J. Ignatius, V. Dubowitz, and K.E. Davies. 1995. Deletions in the survival motor neuron gene on 5q13 in autosomal recessive spinal muscular atrophy. Hum Mol Genet. 4:631-634.

Rogers, D.C., J. Peters, J.E. Martin, S. Ball, S.J. Nicholson, A.S. Witherden, M. Hafezparast, J. Latcham, T.L. Robinson, C.A. Quilter, and E.M. Fisher. 2001. SHIRPA, a protocol for behavioral assessment: validation for longitudinal study of neurological dysfunction in mice. Neurosci Lett. 306:89-92.

Rosen, D.R., T. Siddique, D. Patterson, D.A. Figlewicz, P. Sapp, A. Hentati, D. Donaldson, J. Goto, J.P. O'Regan, H.X. Deng, and R. Brown. 1993. Mutations in the Cu/Zn superoxide dismutase gene are associated with familial amyotrophic lateral sclerosis. Nature. 362:59-62.

Rossman, K.L., D.K. Worthylake, J.T. Snyder, D.P. Siderovski, S.L. Campbell, and J. Sondek. 2002. A crystallographic view of interactions between Dbs and Cdc42: PH domain-assisted guanine nucleotide exchange. Embo J. 21:1315-1326.

Rothstein, J.D., M. Dykes-Hoberg, C.A. Pardo, L.A. Bristol, L. Jin, R.W. Kuncl, Y. Kanai, M.A. Hediger, Y. Wang, J. Schielke, and D. Welty. 1996. Antisense knockout of glutamate transporters reveals a predominant role for astroglial glutamate transport in excitotoxicity and clearance of extracellular glutamate. Neuron. 16:675-686.

Rothstein, J.D., L.J. Martin, and R.W. Kuncl. 1992. Decreased glutamate transport in

244 the brain and spinal cord in amyotrophic lateral sclerosis. New England J Med. 326:1464-1468.

Rothstein, J.D., G. Tsai, R.W. Kuncl, L. Clawson, D.R. Cornblath, D.B. Drachman, A. Pestronk, B.L. Stauch, and J.T. Coyle. 1990. Abnormal excitatory amino acid metabolism in amyotrophic lateral sclerosis. Ann.Neurol. 28:18-25.

Rothstein, J.G., M. Van Kammen, A.I. Levey, L.J. Martin, and R.W. Kuncl. 1995. Selective loss of glial glutamate transporter GLT-1 in amyotrophic laterlal sclerosis. Ann.Neurol. 38:73-84.

Rottner, K., A. Hall, and J.V. Small. 1999. Interplay between Rac and Rho in the control of substrate contact dynamics. Curr Biol. 9:640-648.

Rouleau, G.A., A.W. Clark, K. Rooke, A. Pramatarova, A. Krizus, O. Suchowersky, J.P. Julien, and D. Figlewicz. 1996. SOD1 mutation is associated with accumulation of neurofilaments in amyotrophic lateral sclerosis. Ann. Neurol. 39:128-131.

Ruchhoeft, M.L., S. Ohnuma, L. McNeill, C.E. Holt, and W.A. Harris. 1999. The neuronal architecture of Xenopus retinal ganglion cells is sculpted by rho-family GTPases in vivo. J Neurosci. 19:8454-8463.

Ruddy, D.M., M.J. Parton, A. Al-Chalabi, C.M. Lewis, C. Vance, B.N. Smith, P.N. Leigh, J.F. Powell, T. Siddique, E.P. Meyjes, F. Baas, V. de Jong, and C.E. Shaw. 2003. Two families with familial amyotrophic lateral sclerosis are linked to a novel locus on chromosome 16q. Am J Hum Genet. 73:390-396.

Russo, C., Y. Gao, P. Mancini, C. Vanni, M. Porotto, M. Falasca, M.R. Torrisi, Y. Zheng, and A. Eva. 2001. Modulation of oncogenic DBL activity by phosphoinositol phosphate binding to pleckstrin homology domain. J Biol Chem. 276:19524-19531.

Sagot, Y., T. Rosse, R. Vejsada, D. Perrelet, and A.C. Kato. 1998. Differential effects of neurotrophic factors on motoneuron retrograde labeling in a murine model of

245 motoneuron disease. J Neurosci. 18:1132-1141.

Sahin, M., P.L. Greer, M.Z. Lin, H. Poucher, J. Eberhart, S. Schmidt, T.M. Wright, S.M. Shamah, S. O'Connell, C.W. Cowan, L. Hu, J.L. Goldberg, A. Debant, G. Corfas, C.E. Krull, and M.E. Greenberg. 2005. Eph-dependent tyrosine phosphorylation of ephexin1 modulates growth cone collapse. Neuron. 46:191-204.

Saito, K., J. Murai, H. Kajiho, K. Kontani, H. Kurosu, and T. Katada. 2002. A novel binding protein composed of homophilic tetramer exhibits unique properties for the small GTPase Rab5. J Biol Chem. 277:3412-3418.

Sambrook, J., E.F. Fritsch, and T. Maniatis. 1989. Molecular cloning. A laboratory Manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor.

Sanders, L.C., F. Matsumura, G.M. Bokoch, and P. de Lanerolle. 1999. Inhibition of myosin light chain kinase by p21-activated kinase. Science. 283:2083-2085.

Sanelli, T.R., M.M. Sopper, and M.J. Strong. 2004. Sequestration of nNOS in neurofilamentous aggregate bearing neurons in vitro leads to enhanced NMDA- mediated calcium influx. Brain Res. 1004:8-17.

Santos, M.F., S.A. McCormack, Z. Guo, J. Okolicany, Y. Zheng, L.R. Johnson, and G. Tigyi. 1997. Rho proteins play a critical role in cell migration during the early phase of mucosal restitution. J Clin Invest. 100:216-225.

Sapp, P.C., B.A. Hosler, D. McKenna-Yasek, W. Chin, A. Gann, H. Genise, J. Gorenstein, M. Huang, W. Sailer, M. Scheffler, M. Valesky, J.L. Haines, M. Pericak- Vance, T. Siddique, H.R. Horvitz, and R.H. Brown, Jr. 2003. Identification of Two Novel Loci for Dominantly Inherited Familial Amyotrophic Lateral Sclerosis. Am. J. Hum. Genet. 73:397-403.

Sasaki, S., and M. Iwata. 1996. Ultrastructural study of synapses in the anterior horn neurons of patients with amyotrophic lateral sclerosis. Neurosci Lett. 204:53-56.

246 Sasaki, S., and S. Maruyama. 1994. Immunocytochemical and ultrastructural studies of the motor cortex in amyotrophic lateral sclerosis. Acta Neuropathol (Berl). 87:578-585.

Schaefer, A.W., N. Kabir, and P. Forscher. 2002. Filopodia and actin arcs guide the assembly and transport of two populations of microtubules with unique dynamic parameters in neuronal growth cones. J Cell Biol. 158:139-152.

Schafer, D.A. 2002. Coupling actin dynamics and membrane dynamics during endocytosis. Curr Opin Cell Biol. 14:76-81.

Schiffer, D., S. Cordera, P. Cavalla, and A. Migheli. 1996. Reactive astrogliosis of the spinal cord in amyotrophic lateral sclerosis. J Neurol Sci. 139 Suppl:27-33.

Schmalbruch, H., H.J. Jensen, M. Bjaerg, Z. Kamieniecka, and L. Kurland. 1991. A new mouse mutant with progressive motor neuronopathy. J Neuropathol Exp Neurol. 50:192-204.

Schmidt, A., and A. Hall. 2002. Guanine nucleotide exchange factors for Rho GTPases: turning on the switch. Genes Dev. 16:1587-1609.

Schmidt, S., S. Diriong, J. Mery, E. Fabbrizio, and A. Debant. 2002. Identification of the first Rho-GEF inhibitor, TRIPalpha, which targets the RhoA-specific GEF domain of Trio. FEBS Lett. 523:35-42.

Schurmann, A., A.F. Mooney, L.C. Sanders, M.A. Sells, H.G. Wang, J.C. Reed, and G.M. Bokoch. 2000. p21-activated kinase 1 phosphorylates the death agonist bad and protects cells from apoptosis. Mol Cell Biol. 20:453-461.

Schwamborn, J.C., and A.W. Puschel. 2004. The sequential activity of the GTPases Rap1B and Cdc42 determines neuronal polarity. Nat Neurosci. 7:923-929.

Scita, G., P. Tenca, E. Frittoli, A. Tocchetti, M. Innocenti, G. Giardina, and P.P. Di Fiore. 2000. Signaling from Ras to Rac and beyond: not just a matter of GEFs. Embo J.

247 19:2393-2398.

Sebok, A., N. Nusser, B. Debreceni, Z. Guo, M.F. Santos, J. Szeberenyi, and G. Tigyi. 1999. Different roles for RhoA during neurite initiation, elongation, and regeneration in PC12 cells. J Neurochem. 73:949-960.

Segev, N. 2001. Ypt and Rab GTPases: insight into functions through novel interactions. Curr Opin Cell Biol. 13:500-511.

Sells, M.A. 1999. Pictures in cell biology. Pak1 kinase activity affects the character of cell morphology and movement. Trends Cell Biol. 9:355.

Sells, M.A., J.T. Boyd, and J. Chernoff. 1999. p21-activated kinase 1 (Pak1) regulates cell motility in mammalian fibroblasts. J Cell Biol. 145:837-849.

Sells, M.A., U.G. Knaus, S. Bagrodia, D.M. Ambrose, G.M. Bokoch, and J. Chernoff. 1997. Human p21-activated kinase (Pak1) regulates actin organization in mammalian cells. Curr Biol. 7:202-210.

Shah, J.V., L.A. Flanagan, P.A. Janmey, and J.-F. Leterrier. 2000. Bidirectional translocation of neurofilaments along microtubules mediated in part by dynein/dynactin. Mol.Biol.Cell. 11:3495-3508.

Sharp, D.J., R. Kuriyama, R. Essner, and P.W. Baas. 1997. Expression of a minus-end- directed motor protein induces Sf9 cells to form axon-like processes with uniform microtubule polarity orientation. J Cell Sci. 110 ( Pt 19):2373-2380.

Shaw, P.J., V. Forrest, P.G. Ince, J.P. Richardson, and H.J. Wastell. 1995a. CSF and plasma amino acid levels in motor neuron disease: elevation of CSF glutamate in a subset of patients. Neurodegeneration. 4:209-216.

Shaw, P.J., P.G. Ince, G. Falkous, and D. Mantle. 1995b. Oxidative damage to protein in sporadic motor neuron disease spinal cord. Ann.Neurol. 38:691-695.

248 Shea, T.B., J.T. Yabe, D. Ortiz, A. Pimenta, P. Loomis, R.D. Goldman, N. Amin, and H.C. Pant. 2004. Cdk5 regulates axonal transport and phosphorylation of neurofilaments in cultured neurons. J. Cell Sci. 117:933-941.

Shefner, J.M., Reaume, A.G., Flood, D.G., Scott, R.W., Kowall, N.W., Ferrante, R.J., Siwek, D.F., Upton-Rice, M., and R.H.Jr. Brown. 1999. Mice lacking cytosolic copper/zinc superoxide dismutase display a distinctive motor axonopathy. Neurology. 53:1239-1246

Shibasaki, Y., H. Tanaka, K. Iwabuchi, S. Kawasaki, H. Kondo, K. Uekawa, M. Ueda, T. Kamiya, Y. Katayama, A. Nakamura, H. Takashima, M. Nakagawa, M. Masuda, H. Utsumi, T. Nakamuro, K. Tada, K. Kurohara, K. Inoue, F. Koike, T. Sakai, S. Tsuji, and H. Kobayashi. 2000. Linkage of autosomal recessive hereditary spastic paraplegia with mental impairment and thin corpus callosum to chromosome 15A13-15. Ann Neurol. 48:108-112.

Shibata, N. 2001. Transgenic mouse model for familial amyotrophic lateral sclerosis with superoxide dismutase-1 mutation. Neuropathology. 21:82-92.

Shibata, N., K. Asayama, A. Hirano, and M. Kobayashi. 1996a. Immunohistochemical study on superoxide dismutases in spinal cords from autopsied patients with amyotrophic lateral sclerosis. Dev.Neurosci. 18:492-498.

Shibata, N., A. Hirano, M. Kobayashi, S. Sasaki, T. Kato, S. Matsumoto, Z. Shiozawa, T. Komori, A. Ikemoto, T. Umahara, and et al. 1994. Cu/Zn superoxide dismutase-like immunoreactivity in Lewy body-like inclusions of sporadic amyotrophic lateral sclerosis. Neurosci Lett. 179:149-152.

Shibata, N., A. Hirano, M. Kobayashi, T. Siddique, H.X. Deng, W.Y. Hung, T. Kato, and K. Asayama. 1996b. Intense superoxide dismutase-1 immunoreactivity in intracytoplasmic hyaline inclusions of familial amyotrophic lateral sclerosis with posterior column involvement. J Neuropathol Exp Neurol. 55:481-490.

249 Shibata, N., R. Nagai, K. Uchida, S. Horiuchi, S. Yamada, A. Hirano, M. Kawaguchi, T. Yamamoto, S. Sasaki, and M. Kobayashi. 2001. Morphological evidence for lipid peroxidation and protein glycoxidation in spinal cords from sporadic amyotrophic lateral sclerosis patients. Brain Res. 917:97-104.

Shin, E.Y., K.S. Shin, C.S. Lee, K.N. Woo, S.H. Quan, N.K. Soung, Y.G. Kim, C.I. Cha, S.R. Kim, D. Park, G.M. Bokoch, and E.G. Kim. 2002. Phosphorylation of p85 beta PIX, a Rac/Cdc42-specific guanine nucleotide exchange factor, via the Ras/ERK/PAK2 pathway is required for basic fibroblast growth factor-induced neurite outgrowth. J Biol Chem. 277:44417-44430.

Shin, E.Y., K.N. Woo, C.S. Lee, S.H. Koo, Y.G. Kim, W.J. Kim, C.D. Bae, S.I. Chang, and E.G. Kim. 2004. Basic fibroblast growth factor stimulates activation of Rac1 through a p85 betaPIX phosphorylation-dependent pathway. J Biol Chem. 279:1994- 2004.

Shinder, G.A., M.C. Lacourse, S. Minotti, and H.D. Durham. 2001. Mutant Cu/Zn- superoxide dismutase proteins have altered solubility and interact with heat shock/stress proteins in models of amyotrophic lateral sclerosis. J. Biol. Chem. 276:12791-12796.

Shinohara, M., Y. Terada, A. Iwamatsu, A. Shinohara, N. Mochizuki, M. Higuchi, Y. Gotoh, S. Ihara, S. Nagata, H. Itoh, Y. Fukui, and R. Jessberger. 2002. SWAP-70 is a guanine-nucleotide-exchange factor that mediates signalling of membrane ruffling. Nature. 416:759-763.

Siddique, T., and A. Hentati. 1995. Familial amyotrophic lateral sclerosis. Clin Neurosci. 3:338-347.

Siddique, T., S.T. Hong, B.R. Brooks, W.Y. Hung, N.A. Siddique, J. Rimmler, J.P. Kaplan, J.L. Haines, R.H. Brown Jr., and M.A. Pericak-Vance. 1998. X-linked dominant locus for late-onset familial amyotrophic lateral sclerosis. Amer. J. of Hum. Genetics. 63:A308.

250 Simpson, M.A., H. Cross, C. Proukakis, A. Pryde, R. Hershberger, A. Chatonnet, M.A. Patton, and A.H. Crosby. 2003. Maspardin is mutated in mast syndrome, a complicated form of hereditary spastic paraplegia associated with dementia. Am J Hum Genet. 73:1147-1156.

Singh, R.J., S.P. Goss, J. Joseph, and B. Kalyanaraman. 1998. Nitration of gamma- tocopherol and oxidation of alpha-tocopherol by copper-zinc superoxide dismutase/H2O2/NO2-: role of nitrogen dioxide free radical. Proc Natl Acad Sci U S A. 95:12912-12917.

Skvortsova, V., M. Shadrina, P. Slominsky, G. Levitsky, E. Kondratieva, A. Zherebtsova, N. Levitskaya, A. Alekhin, A. Serdyuk, and S. Limborska. 2004. Analysis of heavy neurofilament subunit gene polymorphism in Russian patients with sporadic motor neuron disease (MND). Eur. J. Hum. Genet.12:241-244

Smirnova, E., D.L. Shurland, S.N. Ryazantsev, and A.M. van der Bliek. 1998. A human dynamin-related protein controls the distribution of mitochondria. J Cell Biol. 143:351- 358.

Smith, R.G., S. Hamilton, F. Hofmann, T. Schneider, W. Nastainczyk, L. Birnbaumer, E. Stefani, and S.H. Appel. 1992. Serum antibodies to L-type calcium channels in patients with amyotrophic lateral sclerosis. N Engl J Med. 327:1721-1728.

Sobue, G., Y. Hashizume, T. Yasuda, E. Mukai, T. Kumagai, T. Mitsuma, and J.Q. Trojanowski. 1990. Phosphorylated high molecular weight neurofilament protein in lower motor neurons in amyotrophic lateral sclerosis and other neurodegenerative diseases involving ventral horn cells. Acta Neuropathol. 79:402-408.

Sone, M., E. Suzuki, M. Hoshino, D. Hou, H. Kuromi, M. Fukata, S. Kuroda, K. Kaibuchi, Y. Nabeshima, and C. Hama. 2000. Synaptic development is controlled in the periactive zones of Drosophila synapses. Development. 127:4157-4168.

Sotiropoulos, A., D. Gineitis, J. Copeland, and R. Treisman. 1999. Signal-regulated

251 activation of serum response factor is mediated by changes in actin dynamics. Cell. 98:159-169.

Spaargaren, M., and J.L. Bos. 1999. Rab5 induces Rac-independent lamellipodia formation and cell migration. Mol Biol Cell. 10:3239-3250.

Stam, J.C., E.E. Sander, F. Michiels, F.N. van Leeuwen, H.E. Kain, R.A. van der Kammen, and J.G. Collard. 1997. Targeting of Tiam1 to the plasma membrane requires the cooperative function of the N-terminal pleckstrin homology domain and an adjacent protein interaction domain. J Biol Chem. 272:28447-28454.

Stamer, K., R. Vogel, E. Thies, E. Mandelkow, and E.-M. Mandelkow. 2002. Tau blocks traffic of organelles, neurofilaments, and APP vesicles in neurons and enhances oxidative stress. J Cell Biol. 156:1051-1063.

Standen, C.L., M.S. Perkinton, H.L. Byers, S. Kesavapany, K.F. Lau, M. Ward, D. McLoughlin, and C.C. Miller. 2003. The neuronal adaptor protein Fe65 is phosphorylated by mitogen-activated protein kinase (ERK1/2). Mol. Cell Neurosci. 24:851-857.

Steinmuller, R., A. Lantigua-Cruz, R. Garcia-Garcia, M. Kostrzewa, D. Steinberger, and U. Muller. 1997. Evidence of a third locus in X-linked recessive spastic paraplegia. Hum Genet. 100:287-289.

Stenmark, H., and V.M. Olkkonen. 2001. The Rab GTPase family. Genome Biol. 2:3007.1-3007.7.

Stenmark, H., R.G. Parton, O. Steele-Mortimer, A. Lutcke, J. Gruenberg, and M. Zerial. 1994. Inhibition of rab5 GTPase activity stimulates membrane fusion in endocytosis. Embo J. 13:1287-1296.

Steven, R., T.J. Kubiseski, H. Zheng, S. Kulkarni, J. Mancillas, A. Ruiz Morales, C.W. Hogue, T. Pawson, and J. Culotti. 1998. UNC-73 activates the Rac GTPase and is

252 required for cell and growth cone migrations in C. elegans. Cell. 92:785-795.

Stieber, A., J.O. Gonatas, and N.K. Gonatas. 2000. Aggregation of ubiquitin and a mutant ALS-linked SOD1 protein correlate with disease progression and fragmentation of the Golgi apparatus. J Neurol Sci. 173:53-62.

Storkebaum E., Lambrechts D., Dewerchin M., Moreno-Murciano M.P., Appelmans S., Oh H., Van Damme P., Rutten B., Man W.Y., De Mol M., Wyns S., Manka D., Vermeulen K., Van Den Bosch L., Mertens N., Schmitz C., Robberecht W., Conway E.M., Collen D., Moons L., and P. Carmeliet. 2005. Treatment of motoneuron degeneration by intracerebroventricular delivery of VEGF in a rat model of ALS. Nat Neurosci. 8:85-92.

Strong, M.J., A.J. Hudson, and W.G. Alvord. 1991. Familial amyotrophic lateral sclerosis, 1850-1989: a statistical analysis of the world literature. Can J Neurol Sci. 18:45-58.

Strong, M.J., Strong, W.L., Jaffe, H., Traggert, B., Sopper, M.M. and Pant, H.C. 2001. Phosphorylation state of the native high-molecular-weight neurofilament subunit protein from cervical spinal cord in sporadic amyotrophic lateral sclerosis. J.Neurochem. 76:1315-1325.

Sturtz, L.A., Diekert, K., Jensen, L.T., Lill, R. and Culotta, V.C. 2001. A fraction of yeast Cu,Zn-superoxide dismutase and its metallochaperone, CCS, localize to the intermembrane space of mitochondria - A physiological role for SOD1 in guarding against mitochondrial oxidative damage. J.Biol.Chem. 276:38084-38089.

Subramaniam, J. R., Lyons, W. E., Liu, J., Bartnikas, T. B., Rothstein, J., Price, D. L., Cleveland, D. W., Gitlin, J. D., Wong, P. C. 2002. Mutant SOD1 causes motor neuron disease independent of copper chaperone-mediated copper loading Nat. Neurosci. 5: 301-307

Suh, Y.A., Arnold, R.S., Lassegue, B., Shi, J., Xu, X., Sorescu, D., Chung, A.B.,

253 Griendling, K.K. and Lambeth, J.D. 1999. Cell transformation by the superoxide- generating oxidase Mox1. Nature. 401:79-82.

Sun, D., C.L. Leung, and R.K.H. Liem. 1996. Phosphorylation of the high molecular weight neurofilament protein (NF-H) by cdk-5 and p35. J. Biol. Chem. 271:14245- 14251.

Suter, D.M., and P. Forscher. 2000. Substrate-cytoskeletal coupling as a mechanism for the regulation of growth cone motility and guidance. J Neurobiol. 44:97-113.

Takahashi, S., and A.B. Kulkarni. 2004. Mutant superoxide dismutase 1 causes motor neuron degeneration independent of cyclin-dependent kinase 5 activation by p35 or p25. J. Neurochem. 88:1295-1304.

Takai, Y., T. Sasaki, and T. Matozaki. 2001. Small GTP-binding proteins. Physiol Rev. 81:153-208.

Takeshima, H., S. Komazaki, M. Nishi, M. Iino, and K. Kangawa. 2000. Junctophilins: a novel family of junctional membrane complex proteins. Mol Cell. 6:11-22.

Takeuchi, H., Y. Kobayashi, S. Ishigaki, M. Doyu, and G. Sobue. 2002a. Mitochondrial localization of mutant SOD1 triggers caspase-dependent cell death in a cellular model of familial amyotrophic lateral sclerosis. J. Biol. Chem. 277:50966-50972.

Takeuchi, H., Y. Kobayashi, T. Yoshihara, J. Niwa, M. Doyu, K. Ohtsuka, and G. Sobue. 2002b. Hsp70 and Hsp40 improve neurite outgrowth and suppress intracytoplasmic aggregate formation in cultured neuronal cells expressing mutant SOD1. Brain Res. 949:11-22.

Talbot, K. 2002. Motor neurone disease. Postgrad Med J. 78:513-519.

Tall, G.G., M.A. Barbieri, P.D. Stahl, and B.F. Horazdovsky. 2001. Ras-activated endocytosis is mediated by the Rab5 guanine nucleotide exchange activity of RIN1. Dev

254 Cell. 1:73-82.

Tanaka, K., K. Watase, T. Manabe, K. Yamada, M. Watanabe, K. Takahashi, H. Iwama, T. Nishikawa, N. Ichihara, T. Kikuchi, S. Okuyama, N. Kawashima, S. Hori, M. Takimoto, and K. Wada. 1997. Epilepsy and exacerbation of brain injury in mice lacking the glutamate transporter GLT-1. Science. 276:1699-1702.

Tashiro, A., and R. Yuste. 2004. Regulation of dendritic spine motility and stability by Rac1 and Rho kinase: evidence for two forms of spine motility. Mol Cell Neurosci. 26:429-440.

Tateno, M., H. Sadakata, M. Tanaka, S. Itohara, R.M. Shin, M. Miura, M. Masuda, T. Aosaki, M. Urushitani, H. Misawa, and R. Takahashi. 2004. Calcium-permeable AMPA receptors promote misfolding of mutant SOD1 protein and development of amyotrophic lateral sclerosis in a transgenic mouse model. Hum Mol Genet. 13:2183-2196.

Taylor, S.J., and D. Shalloway. 1996. Cell cycle-dependent activation of Ras. Curr Biol. 6:1621-1627.

Teramoto, H., O.A. Coso, H. Miyata, T. Igishi, T. Miki, and J.S. Gutkind. 1996a. Signaling from the small GTP-binding proteins Rac1 and Cdc42 to the c-Jun N-terminal kinase/stress-activated protein kinase pathway. A role for mixed lineage kinase 3/protein-tyrosine kinase 1, a novel member of the mixed lineage kinase family. J Biol Chem. 271:27225-27228.

Teramoto, H., P. Crespo, O.A. Coso, T. Igishi, N. Xu, and J.S. Gutkind. 1996b. The small GTP-binding protein rho activates c-Jun N-terminal kinases/stress-activated protein kinases in human kidney 293T cells. Evidence for a Pak-independent signaling pathway. J Biol Chem. 271:25731-25734.

Thies, E., and R.W. Davenport. 2003. Independent roles of Rho-GTPases in growth cone and axonal behavior. J Neurobiol. 54:358-369.

255 Thompson, P.W., A.M. Randi, and A.J. Ridley. 2002. Intercellular adhesion molecule (ICAM)-1, but not ICAM-2, activates RhoA and stimulates c-fos and rhoA transcription in endothelial cells. J Immunol. 169:1007-1013.

Threadgill, R., K. Bobb, and A. Ghosh. 1997. Regulation of dendritic growth and remodeling by Rho, Rac, and Cdc42. Neuron. 19:625-634.

Tibbles, L.A., and J.R. Woodgett. 1999. The stress-activated protein kinase pathways. Cell. Mol. Life Sci. 55:1230-1254.

Tigyi, G., and R. Miledi. 1992. Lysophosphatidates bound to serum albumin activate membrane currents in Xenopus oocytes and neurite retraction in PC12 pheochromocytoma cells. J Biol Chem. 267:21360-21367.

Tikka, T., B.L. Fiebich, G. Goldsteins, R. Keinanen, and J. Koistinaho. 2001. Minocycline, a tetracycline derivative, is neuroprotective against excitotoxicity by inhibiting activation and proliferation of microglia. J Neurosci. 21:2580-2588.

Tikka, T.M., N.E. Vartiainen, G. Goldsteins, S.S. Oja, P.M. Andersen, S.L. Marklund, and J. Koistinaho. 2002. Minocycline prevents neurotoxicity induced by cerebrospinal fluid from patients with motor neurone disease. Brain. 125:722-731.

Tomkins, J., P. Usher, J.Y. Slade, P.G. Ince, A. Curtis, K. Bushby, and P.J. Shaw. 1998. Novel insertion in the KSP region of the neurofilament heavy gene in amyotrophic lateral sclerosis (ALS). Neuroreport. 9:3967-3970.

Topp, J.D., N.W. Gray, R.D. Gerard, and B.F. Horazdovsky. 2004. Alsin is a Rab5 and Rac1 guanine nucleotide exchange factor. J. Biol. Chem 279:24612-24623.

Tortarolo, M., A.J. Crossthwaite, L. Conforti, J.P. Spencer, R.J. Williams, C. Bendotti, and M. Rattray. 2004. Expression of SOD1 G93A or wild-type SOD1 in primary cultures of astrocytes down-regulates the glutamate transporter GLT-1: lack of involvement of oxidative stress. J. Neurochem. 88:481-493.

256 Tortarolo, M., P. Veglianese, N. Calvaresi, A. Botturi, C. Rossi, A. Giorgini, A. Migheli, and C. Bendotti. 2003. Persistent activation of p38 mitogen-activated protein kinase in a mouse model of familial amyotrophic lateral sclerosis correlates with disease progression. Mol. Cell Neurosci. 23:180-192.

Tosney, K.W., and N.K. Wessells. 1983. Neuronal motility: the ultrastructure of veils and microspikes correlates with their motile activities. J Cell Sci. 61:389-411.

Troost, D., P.A. Sillevis Smitt, J.M. de Jong, and D.F. Swaab. 1992. Neurofilament and glial alterations in the cerebral cortex in amyotrophic lateral sclerosis. Acta Neuropathol (Berl). 84:664-673.

Trotta, N., G. Orso, M.G. Rossetto, A. Daga, and K. Broadie. 2004. The hereditary spastic paraplegia gene, spastin, regulates microtubule stability to modulate synaptic structure and function. Curr Biol. 14:1135-1147.

Trotti, D., M. Aoki, P. Pasinelli, U.V. Berger, N.C. Danbolt, R.H. Brown, Jr., and M.A. Hediger. 2001. Amyotrophic lateral sclerosis-linked glutamate transporter mutant has impaired glutamate clearance capacity. J.Biol.Chem. 276:576-582.

Trotti, D., A. Rolfs, N.C. Danbolt, R.H.J. Brown, and M.A. Hediger. 1999. SOD1 mutants linked to amyotrophic lateral sclerosis selectively inactivate a glial glutamate transporter. Nature Neurosci. 2:427-433.

Troy, C.M., K. Brown, L.A. Greene, and M.L. Shelanski. 1990a. Ontogeny of the neuronal intermediate filament protein, peripherin, in the mouse embryo. Neuroscience. 30:217-237.

Troy, C.M., N.A. Muma, L.A. Greene, D.L. Price, and M.L. Shelanski. 1990b. Regulation of peripherin and neurofilament expression in regenerating rat motor neurons. Brain Res. 529:232-238.

Tu, P.H., P. Raju, K.A. Robinson, M.E. Gurney, J.Q. Trojanowski, and V.M.Y. Lee.

257 1996. Transgenic mice carrying a human mutant superoxide dismutase transgene develop neuronal cytoskeletal pathology resembling human amyotrophic lateral sclerosis lesions. Proc. Natl. Acad. Sci. USA. 93:3155-3160.

Turner, B.J., I.K. Cheah, K.J. Macfarlane, E.C. Lopes, S. Petratos, S.J. Langford, and S.S. Cheema. 2003. Antisense peptide nucleic acid-mediated knockdown of the p75 neurotrophin receptor delays motor neuron disease in mutant SOD1 transgenic mice. J Neurochem. 87:752-763.

Udaka, F., M. Kameyama, and M. Tomonaga. 1986. Degeneration of Betz cells in motor neuron disease. A Golgi study. Acta Neuropathol (Berl). 70:289-295.

Urushitani, M., J. Kurisu, K. Tsukita, and R. Takahashi. 2002. Proteasomal inhibition by misfolded mutant superoxide dismutase 1 induces selective motor neuron death in familial amyotrophic lateral sclerosis. J. Neurochem. 83:1030-1042.

Vadlamudi, R.K., C.J. Barnes, S. Rayala, F. Li, S. Balasenthil, S. Marcus, H.V. Goodson, A.A. Sahin, and R. Kumar. 2005. p21-activated kinase 1 regulates microtubule dynamics by phosphorylating tubulin cofactor B. Mol Cell Biol. 25:3726- 3736.

Valencia, A., P. Chardin, A. Wittinghofer, and C. Sander. 1991. The ras : evolutionary tree and role of conserved amino acids. Biochemistry. 30:4637-4648.

Valente, E.M., F. Brancati, V. Caputo, E. Bertini, C. Patrono, D. Costanti, and B. Dallapiccola. 2002. Novel locus for autosomal dominant pure hereditary spastic paraplegia (SPG19) maps to chromosome 9q33-q34. Ann Neurol. 51:681-685.

Valentine, J.S., and P.J. Hart. 2003. Misfolded CuZnSOD and amyotrophic lateral sclerosis. Proc Natl Acad Sci U S A. 100:3617-3622.

Van Aelst, L., and C. D'Souza-Schorey. 1997. Rho GTPases and signaling networks. Genes Dev. 11:2295-2322.

258 Van Aelst, L., T. Joneson, and D. Bar-Sagi. 1996. Identification of a novel Rac1- interacting protein involved in membrane ruffling. Embo J. 15:3778-3786.

Van Den Bosch, L., P. Tilkin, G. Lemmens, and W. Robberecht. 2002. Minocycline delays disease onset and mortality in a transgenic model of ALS. Neuroreport. 13:1067- 1070.

Van den Haute, C., K. Spittaels, J. Van Dorpe, R. Lasrado, K. Vandezande, I. Laenen, H. Geerts, and F. Van Leuven. 2001. Coexpression of human cdk5 and its activator p35 with human protein tau in neurons in brain of triple transgenic mice. Neurobiol Dis. 8:32-44. van Leeuwen, F.N., S. van Delft, H.E. Kain, R.A. van der Kammen, and J.G. Collard. 1999. Rac regulates phosphorylation of the myosin-II heavy chain, actinomyosin disassembly and cell spreading. Nat Cell Biol. 1:242-248.

Vazza, G., M. Zortea, F. Boaretto, G.F. Micaglio, V. Sartori, and M.L. Mostacciuolo. 2000. A new locus for autosomal recessive spastic paraplegia associated with mental retardation and distal motor neuropathy, SPG14, maps to chromosome 3q27-q28. Am J Hum Genet. 67:504-509.

Vechio, J.D., L.I. Bruijn, Z.S. Xu, R.H. Brown, and D.W. Cleveland. 1996. Sequence variants in human neurofilament proteins -absence of linkage to familial amyotrophic lateral sclerosis. Ann.Neurol. 40:603-610.

Veeranna, N.D. Amin, N.G. Ahn, H. Jaffe, C.A. Winters, P. Grant, and H.C. Pant. 1998. 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.

Veldink, J.H., L.H. van den Berg, J.M. Cobben, R.P. Stulp, J.M. De Jong, O.J. Vogels, F. Baas, J.H. Wokke, and H. Scheffer. 2001. Homozygous deletion of the survival motor neuron 2 gene is a prognostic factor in sporadic ALS. Neurology. 56:749-752.

259 Venter, J.C., M.D. Adams, E.W. Myers, P.W. Li, R.J. Mural, G.G. Sutton, H.O. Smith, M. Yandell, C.A. Evans, R.A. Holt, J.D. Gocayne, P. Amanatides, R.M. Ballew, D.H. Huson, J.R. Wortman, Q. Zhang, C.D. Kodira, X.H. Zheng, L. Chen, M. Skupski, G. Subramanian, P.D. Thomas, J. Zhang, G.L. Gabor Miklos, C. Nelson, S. Broder, A.G. Clark, J. Nadeau, V.A. McKusick, N. Zinder, A.J. Levine, R.J. Roberts, M. Simon, C. Slayman, M. Hunkapiller, R. Bolanos, A. Delcher, I. Dew, D. Fasulo, M. Flanigan, L. Florea, A. Halpern, S. Hannenhalli, S. Kravitz, S. Levy, C. Mobarry, K. Reinert, K. Remington, J. Abu-Threideh, E. Beasley, K. Biddick, V. Bonazzi, R. Brandon, M. Cargill, I. Chandramouliswaran, R. Charlab, K. Chaturvedi, Z. Deng, V. Di Francesco, P. Dunn, K. Eilbeck, C. Evangelista, A.E. Gabrielian, W. Gan, W. Ge, F. Gong, Z. Gu, P. Guan, T.J. Heiman, M.E. Higgins, R.R. Ji, Z. Ke, K.A. Ketchum, Z. Lai, Y. Lei, Z. Li, J. Li, Y. Liang, X. Lin, F. Lu, G.V. Merkulov, N. Milshina, H.M. Moore, A.K. Naik, V.A. Narayan, B. Neelam, D. Nusskern, D.B. Rusch, S. Salzberg, W. Shao, B. Shue, J. Sun, Z. Wang, A. Wang, X. Wang, J. Wang, M. Wei, R. Wides, C. Xiao, C. Yan, et al. 2001. The sequence of the . Science. 291:1304-1351.

Vijayvergiya, C., M.F. Beal, J. Buck, and G. Manfredi. 2005. Mutant superoxide dismutase 1 forms aggregates in the brain mitochondrial matrix of amyotrophic lateral sclerosis mice. J Neurosci. 25:2463-2470.

Villa, P., S.H. Kaufmann, and W.C. Earnshaw. 1997. Caspases and caspase inhibitors. Trends Biochem.Sci. 22:388-393.

Vukosavic, S., M. Dubois-Dauphin, N. Romero, and S. Przedborski. 1999. Bax and Bcl- 2 interaction in a transgenic mouse model of familial amyotrophic lateral sclerosis. J Neurochem. 73:2460-2468.

Wada, M., H. Nakanishi, A. Satoh, H. Hirano, H. Obaishi, Y. Matsuura, and Y. Takai. 1997. Isolation and characterization of a GDP/GTP exchange protein specific for the Rab3 subfamily small G proteins. J Biol Chem. 272:3875-3878.

Wagey, R., S.L. Pelech, V. Duronio, and C. Krieger. 1998. Phosphatidylinositol 3- kinase: increased activity and protein level in amyotrophic lateral sclerosis. J

260 Neurochem. 71:716-722.

Wang, J., H. Slunt, V. Gonzales, D. Fromholt, M. Coonfield, N.G. Copeland, N.A. Jenkins, and D.R. Borchelt. 2003. Copper-binding-site-null SOD1 causes ALS in transgenic mice: aggregates of non-native SOD1 delineate a common feature. Hum Mol Genet. 12:2753-2764.

Wang, L., and A. Brown. 2002. Rapid movement of microtubules in axons. Curr Biol. 12:1496-1501.

Wang, X., Culotta, V.C., and C.B. Klee. 1996. Superoxide dismutase protects calcineurin from inactivation. Nature. 383:434-437.

Warita, H., K. Abe, Y. Setoguchi, and Y. Itoyama. 1998. Expression of adenovirus- mediated E. coli lacZ gene in skeletal muscles and spinal motor neurons of transgenic mice with a mutant superoxide dismutase gene. Neurosci Lett. 246:153-156.

Warita, H., Y. Itoyama, and K. Abe. 1999. Selective impairment of fast anterograde axonal transport in the peripheral nerves of asymptomatic transgenic mice with a G93A mutant SOD1 gene. Brain Res. 819:120-131.

Weiner, M.P., G.L. Costa, W. Schoettlin, J. Cline, E. Mathur, and J.C. Bauer. 1994. Site-directed mutagenesis of double-stranded DNA by the polymerase chain reaction. Gene. 151:119-123.

Welch, H.C., W.J. Coadwell, C.D. Ellson, G.J. Ferguson, S.R. Andrews, H. Erdjument- Bromage, P. Tempst, P.T. Hawkins, and L.R. Stephens. 2002. P-Rex1, a PtdIns(3,4,5)P3- and Gbetagamma-regulated guanine-nucleotide exchange factor for Rac. Cell. 108:809-821.

Wennerberg, K., S.M. Ellerbroek, R.Y. Liu, A.E. Karnoub, K. Burridge, and C.J. Der. 2002. RhoG signals in parallel with Rac1 and Cdc42. J Biol Chem. 277:47810-47817.

261 Westwick, J.K., Q.T. Lambert, G.J. Clark, M. Symons, L. Van Aelst, R.G. Pestell, and C.J. Der. 1997. Rac regulation of transformation, gene expression, and actin organization by multiple, PAK-independent pathways. Mol Cell Biol. 17:1324-1335.

Wharton, S.B., C.J. McDermott, A.J. Grierson, J.D. Wood, C. Gelsthorpe, P.G. Ince, and P.J. Shaw. 2003. The cellular and molecular pathology of the motor system in hereditary spastic paraparesis due to mutation of the spastin gene. J Neuropathol Exp Neurol. 62:1166-1177.

Wherlock, M., and H. Mellor. 2002. The Rho GTPase family: a Racs to Wrchs story. J Cell Sci. 115:239-240.

Whitehead, I., H. Kirk, C. Tognon, G. Trigo-Gonzalez, and R. Kay. 1995. Expression cloning of lfc, a novel oncogene with structural similarities to guanine nucleotide exchange factors and to the regulatory region of protein kinase C. J Biol Chem. 270:18388-18395.

Whitehead, I.P., Q.T. Lambert, J.A. Glaven, K. Abe, K.L. Rossman, G.M. Mahon, J.M. Trzaskos, R. Kay, S.L. Campbell, and C.J. Der. 1999. Dependence of Dbl and Dbs transformation on MEK and NF-kappaB activation. Mol Cell Biol. 19:7759-7770.

Wiedau-Pazos, M., J.J. Goto, S. Rabizadeh, E.B. Gralla, J.A. Roe, M.K. Lee, J.S. Valentine, and D.E. Bredesen. 1996. Altered reactivity of superoxide dismutase in familial amyotrophic lateral sclerosis. Science. 271:515-518.

Wilhelmsen, K.C., M.S. Forman, H.J. Rosen, L.I. Alving, J. Goldman, J. Feiger, J.V. Lee, S.K. Segall, J.H. Kramer, C. Lomen-Hoerth, K.P. Rankin, J. Johnson, H.S. Feiler, M.W. Weiner, V.M. Lee, J.Q. Trojanowski, and B.L. Miller. 2004. 17q-linked frontotemporal dementia-amyotrophic lateral sclerosis without tau mutations with tau and alpha-synuclein inclusions. Arch Neurol. 61:398-406.

Wilkinson, P.A., A.H. Crosby, C. Turner, L.J. Bradley, L. Ginsberg, N.W. Wood, A.H. Schapira, and T.T. Warner. 2004. A clinical, genetic and biochemical study of SPG7

262 mutations in hereditary spastic paraplegia. Brain. 127:973-980.

Wilkinson, P.A., M.A. Simpson, L. Bastaki, H. Patel, J.A. Reed, K. Kalidas, E. Samilchuk, R. Khan, T.T. Warner, and A.H. Crosby. 2005. A new locus for autosomal recessive complicated hereditary spastic paraplegia (SPG26) maps to chromosome 12p11.1-12q14. J Med Genet. 42:80-82.

Williamson, T.L., and D.W. Cleveland. 1999. Slowing of axonal transport is a very early event in the toxicity of ALS-linked SOD1 mutants to motor neurons. Nature Neurosci. 2:50-56.

Windpassinger, C., M. Auer-Grumbach, J. Irobi, H. Patel, E. Petek, G. Horl, R. Malli, J.A. Reed, I. Dierick, N. Verpoorten, T.T. Warner, C. Proukakis, P. Van den Bergh, C. Verellen, L. Van Maldergem, L. Merlini, P. De Jonghe, V. Timmerman, A.H. Crosby, and K. Wagner. 2004. Heterozygous missense mutations in BSCL2 are associated with distal hereditary motor neuropathy and Silver syndrome. Nat Genet. 36:271-276.

Wong, N.K.Y., B.P. He, and M.J. Strong. 2000. 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.

Wong, P.C., and D.R. Borchelt. 1995. Motor neuron disease caused by mutations in superoxide dismutase 1. Curr Opin Neurol. 8:294-301.

Wong, P.C., C. Pardo, D.R. Borchelt, M.K. Lee, N.G. Copeland, N.A. Jenkins, S.S. Sisodia, D.W. Cleveland, and D.L. Price. 1995. An adverse property of a familial ALS- linked SOD1 mutation causes motor neuron disease characterised by vacuolar degeneration of mitochondria. Neuron. 14:1105-1116.

Xia, C.H., E.A. Roberts, L.S. Her, X. Liu, D.S. Williams, D.W. Cleveland, and L.S. Goldstein. 2003. Abnormal neurofilament transport caused by targeted disruption of neuronal kinesin heavy chain KIF5A. J. Cell Biol. 161:55-66.

263 Xiao, J., C.B. Neylon, B. Hunne, and J.B. Furness. 2003. Oligophrenin-1, a Rho GTPase-activating protein (RhoGAP) involved in X-linked mental retardation, is expressed in the enteric nervous system. Anat Rec A Discov Mol Cell Evol Biol. 273:671-676.

Xu, Z., L.C. Cork, J.W. Griffin, and D.W. Cleveland. 1993. Increased expression of neurofilament subunit NF-L produces morphological alterations that resemble the pathology of human motor neuron disease. Cell. 73:23-33.

Yabe, J.T., A. Pimenta, and T.B. Shea. 1999. Kinesin-mediated transport of neurofilament protein oligomers in growing axons. J.Cell Sci. 112:3799-3814.

Yaku, H., T. Sasaki, and Y. Takai. 1994. The Dbl oncogene product as a GDP/GTP exchange protein for the Rho family: its properties in comparison with those of Smg GDS. Biochem Biophys Res Commun. 198:811-817.

Yamanaka, K., C. Vande Velde, E. Eymard-Pierre, E. Bertini, O. Boespflug-Tanguy, and D.W. Cleveland. 2003. Unstable mutants in the peripheral endosomal membrane component ALS2 cause early-onset motor neuron disease. Proc Natl Acad Sci U S A. 100:16041-16046.

Yamashita, T., K.L. Tucker, and Y.A. Barde. 1999. Neurotrophin binding to the p75 receptor modulates Rho activity and axonal outgrowth. Neuron. 24:585-593.

Yang, Y., A. Hentati, H.X. Deng, O. Dabbagh, T. Sasaki, M. Hirano, W.Y. Hung, K. Ouahchi, J.H. Yan, A.C. Azim, N. Cole, G. Gascon, A. Yagmour, M. Ben-Hamida, M. Pericak-Vance, F. Hentati, and T. Siddique. 2001. The gene encoding alsin, a protein with three guanine-nucleotide exchange factor domains, is mutated in a form of recessive amyotrophic lateral sclerosis. Nature Genet. 29:160-165.

Yasojima, K., W.W. Tourtellotte, E.G. McGeer, and P.L. McGeer. 2001. Marked increase in cyclooxygenase-2 in ALS spinal cord: implications for therapy. Neurology. 57:952-956.

264 Yim, M.B., J.H. Kang, H.S. Yim, H.S. Kwak, P.B. Chock, and E.R. Stadtman. 1996. A gain-of-function of an amyotrophic lateral sclerosis-associated Cu,Zn-superoxide dismutase mutant: An enhancement of free radical formation due to a decrease in Km for hydrogen peroxide. Proc Natl Acad Sci U S A. 93:5709-5714.

Yoshihara, T., M. Yamamoto, N. Hattori, K. Misu, K. Mori, H. Koike, and G. Sobue. 2002. 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.

Yoshii, S., M. Tanaka, Y. Otsuki, T. Fujiyama, H. Kataoka, H. Arai, H. Hanai, and H. Sugimura. 2001. Involvement of alpha-PAK-interacting exchange factor in the PAK1-c- Jun NH(2)-terminal kinase 1 activation and apoptosis induced by benzo[a]pyrene. Mol Cell Biol. 21:6796-6807.

Yoshiyama, Y., T. Yamada, K. Asanuma, and T. Asahi. 1994. Apoptosis related antigen, Le(Y) and nick-end labeling are positive in spinal motor neurons in amyotrophic lateral sclerosis. Acta Neuropathol (Berl). 88:207-211.

Zenke, F.T., M. Krendel, C. DerMardirossian, C.C. King, B.P. Bohl, and G.M. Bokoch. 2004. p21-activated kinase 1 phosphorylates and regulates 14-3-3 binding to GEF-H1, a microtubule-localized Rho exchange factor. J Biol Chem. 279:18392-18400.

Zhang, H., D.J. Webb, H. Asmussen, S. Niu, and A.F. Horwitz. 2005. A GIT1/PIX/Rac/PAK signaling module regulates spine morphogenesis and synapse formation through MLC. J Neurosci. 25:3379-3388.

Zhang, P., P.-h. Tu, F. Abtahian, J.Q. Trojanowski, and V.M.-Y. Lee. 1997. Neurofilaments and orthograde transport are reduced in ventral root axons of transgenic mice that express human SOD1 with a G93A mutation. J. Cell Biol. 139:1307-1315.

Zhang, S., J. Han, M.A. Sells, J. Chernoff, U.G. Knaus, R.J. Ulevitch, and G.M. Bokoch. 1995. Rho family GTPases regulate p38 mitogen-activated protein kinase

265 through the downstream mediator Pak1. J Biol Chem. 270:23934-23936.

Zhao, C., J. Takita, Y. Tanaka, M. Setou, T. Nakagawa, S. Takeda, H.W. Yang, S. Terada, T. Nakata, Y. Takei, M. Saito, S. Tsuji, Y. Hayashi, and N. Hirokawa. 2001a. Charcot-marie-tooth disease type 2a caused by mutation in a microtubule motor kif1bbeta. Cell. 105:587-589.

Zhao, X., D. Alvarado, S. Rainier, R. Lemons, P. Hedera, C.H. Weber, T. Tukel, M. Apak, T. Heiman-Patterson, L. Ming, M. Bui, and J.K. Fink. 2001b. Mutations in a newly identified GTPase gene cause autosomal dominant hereditary spastic paraplegia. Nat Genet. 29:326-331.

Zheng, Y. 2001. Dbl family guanine nucleotide exchange factors. Trends Biochem Sci. 26:724-732.

Zhou, Q., M. Xiao, and R.A. Nicoll. 2001. Contribution of cytoskeleton to the internalization of AMPA receptors. Proc Natl Acad Sci U S A. 98:1261-1266.

Zhu, S., I.G. Stavrovskaya, M. Drozda, B.Y.S. Kim, V. Ona, M.W. Li, S. Sarang, A.S. Liu, D.M. Hartley, C.W. Du, S. Gullans, R.J. Ferrante, S. Przedborski, B.S. Kristal, and R.M. Friedlander. 2002. Minocycline inhibits cytochrome c release and delays progression of amyotrophic lateral sclerosis in mice. Nature. 417:74-78.

Zipkin, I.D., R.M. Kindt, and C.J. Kenyon. 1997. Role of a new Rho family member in cell migration and axon guidance in C. elegans. Cell. 90:883-894.

Zuchner, S., M. Vorgerd, E. Sindern, and J.M. Schroder. 2004. The novel neurofilament light (NEFL) mutation Glu397Lys is associated with a clinically and morphologically heterogeneous type of Charcot-Marie-Tooth neuropathy. Neuromuscul Disord. 14:147- 157.

266