factors in huntington’s disease - university of leicester (2).pdf · factors in huntington’s...

228
Interrogating the role of splicing factors in Huntington’s Disease Thesis submitted for the degree of Doctor of Philosophy University of Leicester By Gurdeep Kooner BSc Department of Genetics University of Leicester September 2014

Upload: trinhminh

Post on 28-Apr-2018

217 views

Category:

Documents


2 download

TRANSCRIPT

Interrogating the role of splicing

factors in Huntington’s Disease

Thesis submitted for the degree of Doctor

of Philosophy

University of Leicester

By

Gurdeep Kooner BSc

Department of Genetics

University of Leicester

September 2014

Abstract

Huntington’s disease (HD) is a fatal autosomal dominant neurodegenerative disorder

caused by the expansion of a polyglutamine tract in the huntingtin (HTT) protein.

Given the wide range of cellular interactions involving HTT, pathogenesis is attributed

to both disruption of numerous cellular and metabolic pathways, as well as toxic gain-

of-function effects, most notably a propensity of mutant HTT to misfold and

aggregate. mRNA splicing defects have been observed in several neurological diseases,

though the role of splicing in HD pathogenesis is unclear. A recent genetic modifier

screen in baker’s yeast identified several splicing genes which suppress mutant HTT

toxicity when overexpressed. I initially interrogated these candidate genes in

mammalian cell lines to elucidate the mechanism(s) underlying this protection.

Using HD cell models, I have found that overexpression of these splicing genes reduces

the level of caspase3/7 activation, a mark for apoptosis. Furthermore, while I

observed a mutant HTT-induced impairment of splicing, overexpression of the

suppressors failed to ameliorate this defect. However, using automated aggregation

analyses, these suppressors were found to modulate HTT aggregation dynamics- a

possible mechanism of suppression. I have further interrogated these promising

splicing gene hits in the nematode C. elegans, though the results were inconclusive.

While characterising disease-relevant phenotypes in HD model mice, I uncovered a

robust burrowing defect in these mice, which manifested earlier than impairments in

the more widely-used rota-rod locomotor assay. As such, this test could provide a

means of detecting early behavioural dysfunction in HD mouse models.

Finally I sought to optimise intranasal delivery of lentivirus to the brain, as an

alternative to a stereotactic approach. Our results were incredibly variable yet

promising and with further optimisation, this technique could be a viable alternative

to more invasive methods.

Ultimately my work has provided insight into the modulatory potential of splicing

factors in HD.

Acknowledgements

While the past four years have raced by, writing this thesis has felt like a Sisyphean

task. But as I approach completion, I am reminded of the multitude of people who

have helped me to get to this point.

First and foremost, I would like to thank my two supervisors- Dr Flaviano Giorgini,

whose continued support, guidance and mentorship, coupled with his constant

attempts at humour (one day, you’ll genuinely make me laugh) throughout the four

years, has seen me grow into a bonafide scientist- Professor Giovanna Mallucci, whose

guidance and attention-to-detail has greatly enhanced my research.

I would also like to thank the two postdoctoral researcher who were assigned the

unenviable task of guiding me through the first year of my research, and have endured

endless inane questions throughout the subsequent years. Dr Robert Mason, quite

possibly the smartest man I have ever met, has always been on hand to offer advice

and support, to me and the majority of Department of Genetics. Dr Julie Moreno, the

hardest working person I have ever met, who has remained humble despite publishing

incredible work, and all the while has provided unrivalled guidance throughout my

PhD.

I would also like to thank Dr Mark Halliday and Dr Helois Radford for their support

regarding the C. elegans and in vivo work respectively. By extension, I would like to

thank Colin Molloy, Lucy Onion, Alison Smart, and the other animal technicians who

have helped me with all of the varying aspects of the murine work over the past four

years. I would also like to thank Dr Lucia Pinion for all of her help with the Cellomics

work, who was always on hand to prevent the equipment from “going bananas”. I

would like to acknowledge Dr Kees Strattman and Dr David Read for all of their

support and advice regarding microscopy and Jennifer Edwards for performing all of

the histology work.

Finally, I would like to thank everyone in Lab 106/104 in the Department of Genetics

and Lab 605 in the MRC Toxicology, for making both labs a great place to work.

Abbreviations

AD Alzheimer's disease AD Adenovirus ALS Amyptrophic lateral sclerosis APC Anaphase promoting complex APS Ammonium persulphate ASH Amphid sensory neuron H ASI Amphid sensory neuron I

Aβ-42 β-amyloid peptide BBB Blood brain barrier BC200RNA Brain cytoplasmic RNA 200nt BCL-XL B-cell lymphoma extra-large

BCSF Blood cerebrospinal fluid BDNF Brain-derived neurotrophic factor BS Branch site CBP CREB binding protein ClC-1 Chloride channel 1 CMV Cytomegalovirus CNS Central nervous system CNTF ciliary neurotrophic factor CRNKL1 Crooked neck pre-mRNA splicing factor 1 CSF Cerebrospinal fluid CSTF2 Cleavage stimulatory factor 2 CTD C-terminal domain DM1 Myotonic dystrophy type 1

DMEM Dulbecco's modified eagle media DNA Deoxyribonucleic acid DRH-1 Dicer related helicase 1 DTT Dithiothreitol ECL Enhanced chemiluminescence EFTUD2 Elongation factor Tu GTP binding domain containing 2 EGCG Epigallocatechin-gallate EGFP Enhanced green flourescent protein EGO-1 Enhancer of glip 1 EIPA 5-(N-ethyly-N-isopropyl)-amiloride ERI-1 Enhanced RNAi 1 ESE Exonic splicing enhancer

ESS Exonic splicing silencer EV Empty Vector F1 Barbituric acid-like compound

FTDP-17 Frontotemporal dementia with Parkinsonism linkded to chromosome 17

FTLD Frontotemporal lobar degeneration

GABA Rγ2 γ-aminobutyric acid receptor type A γ2 subunit

GAPDH Glyceraldehyde-3-phosphate dehydrogenase

GFP Green flourescent protein GLT1 Glutamate uptake transportes GLYRα2 α2 subunit of glycine receptor GTP Guanosine triphosphate GYR Glycine-tyrosine-arginine rich HAP1 Huntingtin associated protein 1 HD Huntingtons disease

HEAT Huntingtin, elongation factor 3, regulatory A subunit of phosphatase 2A and TORI

HEK293T Human embryonic kidney cells HIP Huntingtin interacting proteins HNRNPF Heterogeneous ribonucleoprotein F HNRNPK Heteregeneous ribonucleoprotein K

HNRNPQ Heterogeneous ribonucleoprotein Q HNRNPU Heterogeneous ribonucleoprotein U HSF1 Heat shock transcription factor 1

HSF25 Heat shock protein 25 HSP40 Heat shock protein 40 HSP70 Heat shock protein 70 HTT Huntingtin IAA Iso amylalcohol IER3 Immediate early response 3 IFT Intraflagellar transport complex IRES Internal ribosome entry point ISE Intronic splicing enhancer

ISS Intronic splicing silencer KH K homology KMO Kynurenine-3-monoxogenase LB Luria Bertani LMWH Low molecular weight heparin LPS Lipopolysaccharide MAPT Microtubule-associated protein tau MBNL1 Muscleblind-like 1 MCC Mucocillary clearance MCS Multiple cloning site MUB Mushroom body expressed NALP1 Pyrin domain containing 1 NGM Nematode growth media

NLS Nuclear localisation signal NMD Nonsense mediated decay NMDAR N-methyl-D-aspartate receptors NOVA-1 Neuro-oncological ventral antigen 1 NRSE Neuron-restrictive silencer elements ORF Open reading frame PABP Poly(A)-binding protein

PACSIN1 Casein kinase 2 substrate in neurons 1

PBS Phosphate buffered saline PCR Polymerase chain reaction PD Parkinson's disease PGRN Survival factor progranulin PHA Phasmid neuron A PHB Phasmid neuron B PND Paraneoplastic neurological disorders PolyP Polyproline PolyQ Polyglutamine POMA Paraneoplastic opsoclonus myoclonus ataxia PQE-1 PolyQ enhancer 1 PSD-95 Post synaptic density 95 PT Polypyrimidine tract

PTC Premature termination codon RDE-1 RNAi deficient 1 RDE-3 RNAi deficient 3 RDE-4 RNAi deficient 4 REST Repressor element 1 silencing transcription factor RFF-1 RNA-dependent RNA polymerase RFP Red fluorescent protein RISC RNA induced silencing complex RML Rocky Mountains Laboratory RNA Ribonucleic acid RNAPII RNA polymerase II RRM RNA recognition motifs

RRM RNA recognition motifs RTN3 Reticulon 3 RT-PCR Reverse transcriptase polymerase chain reaction SAFA Scaffolding attachment factor A SDS Sodium dodecyl sulphate SEM Standard error of mean SFRS3 Splicing factor arginine serine 3 SH3 Src homology 3 SK Human tropomysosin SMA Spinal muscular atrophy SMN Surivival of motor neuron SNAP-25 Synaptosomal-associated protein, 25kDa SnRNP Small nuclear ribonucleoprotein

SNRPB Small nuclear ribonucleoprotein polypeptide B SOD Superoxide dismutase SP1 Specificity protein 1 SV40 Simian virus 40 TARDBP Tranactivation responsive DNA-binding protein 43 TBE Tris-borate-EDTA TDM tetradecylmaltoside

TEMED Tetramethylethylenediamine

TNF-α Tumour necrosis factor alpha TRKB Tropomyosin receptor kinase B TRN2 Transporter importin 2 TRP Tetratricopeptide TSN-1 Tudor staphylococcal nuclease homolog 1 U2AF U2 auxiliary factor UEA I Ulex europeus agglutinin I VIG-1 Vasa intronic gene 1 VP16 Viral transcriptional activator WGA Wheat germ agglutinin WGA-HRP Horseradish peroxidase conjugated wheat germ agglutinin WT Wild-type WW Tryptophan domain

YAC Yeast artifical chromosome

Table of Contents

Chapter 1

General Introduction

1.1 Introduction

1.11 Huntington’s Disease

1.12 Huntingtin

1.13 Normal HTT function and mutant HTT

1.14 HD and protein homeostasis

1.15 Aggregation dynamics

1.16 mRNA splicing and alternative splicing

1.17 mRNA splicing and neurodegenerative

disease

1.18 Project aims

Chapter 2

In vitro validation and characterisation of splicing factors

2.1 Introduction

2.11 Splicing suppressor genes

2.12 Aims

2.2 Materials and methods

2.21 Materials

2.211 Bacterial strains and cell lines

2.212 Constructs

2.212 Media and agar

2.22 Methods

2.221 Generation of overexpression cell lines

2.222 Validation of splicing gene hits

2.223 Characterisation of splicing gene hits

2.23 Statistics

2.3 Results

2.31 Overexpression of several candidate genetic modifiers

suppresses mutant HTT toxicity in mammalian cells.

1

1

2

5

8

10

13

22

25

27

29

38

39

39

39

40

41

42

42

49

52

55

59

59

2.32 Suppressors are overexpressed and localise in the

nucleus

2.33 Mutant HTT causes a polyQ dependent decline in

splicing efficiency and cell viability

2.34 Overexpression of suppressors fails to ameliorate

polyQ dependent splicing defects

2.35 Overexpression of suppressors alters the aggregation

dynamics of mutant HTT

2.36 Overexpression of suppressors brought about changes

in HTT protein levels

2.37 Quantitative real-time PCR revealed changes in

mutant HTT expression with no variation in HTT copy

number between cell lines.

2.4 Discussion

2.5 Future work

Chapter 3

Exploring the role of splicing genes in Huntington’s disease using C. elegans

3.1 Introduction

3.11 The use of C. elegans as a model organism

3.12 C. elegans and RNAi

3.13 C. elegans and Huntington’s disease

3.14 Aims

3.2 Materials and methods

3.21 Materials

3.211 Bacterial strains and nematode strains

3.212 Constructs

3.212 Media and agar

3.22 Methods

3.3 Results

3.31 HD worms do not exhibit the dye-filling defect

3.32 RNAi knockdown of splicing-related suppressors does not

modulate HD phenotypes

63

68

70

73

76

78

83

86

86

88

92

95

96

96

96

96

97

98

101

101

103

60

3.4 Discussion

3.5 Future work

Chapter 4

Testing new behavioural paradigm s in HD models mice

4.1 Introduction

4.11 The use of mouse models in the study of HD

4.12 HD mouse models

4.13 Knockout mouse models

4.14 Knock-in mouse models

4.15 Transgenic mouse models

4.16 Aims

4.2 Materials and methods

4.21 Materials

4.211 Bacterial strain

4.212 Constructs

4.213 Mouse strain

4.22 Methods

4.221 Genotyping

4.222 Lentiviral construction

4.223 Behavioural tests

4.3 Results

4.31 Knock-in exhibit a rota-rod defect at 14 months

4.32 Activity of HdhQ150 knock-in mice begins to deviate from

wild-type after 12 months

4.33 Homozygous HdhQ150 knock-in mice failed to gain

weight during their life span

4.34 Homozygous HdhQ150 mice display a burrowing defect

at 9 month

4.35 CAG tract in HdhQ150 displayed genomic instability

4.4 Discussion

4.5 Future work

114

114

114

115

116

117

120

121

121

121

123

123

124

124

124

128

129

130

131

133

134

135

136

140

108

112

Chapter 5

Validating intranasal delivery in prion infected mice

5.1 Introduction

5.11 Delivery to the CNS

5.12 Intranasal delivery

5.13 Factors affecting intranasal delivery

5.14 Gene therapy and neurodegenerative diseases

5.15 Aims

5.2 Materials and method

5.21 Material

5.211 Lentiviral constructs

5.212 Mouse models

5.22 Methods

5.3 Results

5.31 GFP expression detected in both the cortex and

hippocampus following intranasal delivery of MW1 or EV

into prion infected hemizygous mice

5.32 The MW1 PRNP knockdown lentiviral constructs failed

to significantly reduce prion levels in C57BL6N wild type mice

5.33 PRNP knockdown increased lifespan and a delayed onset of

spongiosis in prion infected homozygous mice.

5.34 Early intranasal treatment of prion infected hemizygous

revealed a delay in spongiosis though no change in synaptic

protein level changes.

5.35 Lowered lentiviral dose, or the use of chitosan, rifampin

and TMD fail to extend survival of prion infected mice

5.36 Treatment of prion infected hemizygous mice with MW1

at 6, 7 and 8 w.p.i increased survival.

5.4 Discussion

5.5 Future work

141

141

142

146

148

150

153

153

153

153

154

162

162

164

130

131

166

167

172

174

177

182

Chapter 6

Concluding remarks

References

186

188

Tables and Figures

Tables

2.1 Primers for amplification of candidate genetic modifiers.

2.2 Sequencing primers for validation of mouse cDNA

overexpression constructs.

2.3 PCR primers to assess overexpression of candidate modifiers.

2.4 Protein sample dilutions and loading volumes.

2.5 Primers to assess HTT construct expression and copy number.

3.1 Worm orthologues of suppressors and PCR primers.

3.2 Quantitative PCR primers.

4.1 List of genotyping primers and melting temperatures.

4.2 List of primers used cloning of EFTUD2, HNRNPF and HNRNPK

into pLENTI vector.

5.1 Clinical signs of prion disease

5.1 Intranasal enhancers.

5.2 List of RT-PCR primer sequences.

43

47

51

55

57

97

100

124

127

152

155

156

Figures

1.1 Schematic diagram of HTT.

1.2 Putative roles of wildtype and mutant HTT in brain-derived

neurotrophic factor (BDNF) synthesis and transport.

1.2 HTT aggregation dynamics.

1.3 Pre-mRNA splicing mechanism.

1.4 Alternative splicing.

1.5 Homeostatic regulation of the antagonistic splicing factor families: SR

and hnRNP proteins.

1.6 Splicing of survival of motor neuron 2 (SMN2).

2.1 Schematic representation of the murine structures of each of the

putative HTT suppressor proteins.

2.2 Inducible HTT construct found in the PC-12 cell line.

2.3 Overexpression of several splicing genes modulates caspase activation

in neuronal cells expressing mutant HTT.

2.4 The RFP tagged suppressors localised in the nucleus.

2.5 Validation of suppressor expression.

2.6 Splicing efficiency assay used to ascertain the possibility of splicing

defects in HD.

2.7 Expansion of the HTT polyglutamine tract causes a decline in splicing

efficiency and cell viability 48 hours post transfection.

2.8 Mutant HTT constructs do not significantly change splicing efficiency or cell

viability 24 hours post transfection.

2.9 Overexpression of the candidate genes does not alter splicing efficiency in

HTT97Q cells.

2.10 Cellomics cell identification and analysis.

2.11 Mutant HTT aggregation is significantly altered by overexpression of

splicing genes.

2

6

12

16

18

20

23

28

38

60

62

63

65

66

67

69

70

72

2.12 Mutant HTT levels were elevated in cell lines overexpressing CRNKL,

HNRNPK and TARDBP, and reduced in cells overexpressing EFTUD2 and

HNRNPF.

2.13 Mutant HTT aggregation levels were found to be consistent with the data

derived from the automated cell analysis software.

2.14 Overexpression cell lines display variation in HTT expression relative to

untransfected control cells.

2.15 Stably transfected cell lines exhibit little variation in HTT copy number.

3.1 C. elegans life cyle.

3.2 Schematic, confocal and fluorescence diagrams of C. elegans.

3.3 Schematic of post transcriptional silencing in C. elegans.

3.4 Confocal images of worms following incubation within DiD lipophilic dye.

3.5 Knockdown of target genes failed to exacerbate motor defects in HD

nematode strains.

3.6 Knockdown of target genes does not alter sensitivity defects in HD

nematode strains.

3.7 Knockdown constructs brought about a moderate decrease in expression of

CRNKL and HNRNPF, and an increase in expression of HNRNPK and TARDBP.

4.1 Timeline of behavioural and pathological symptoms exhibited by HD mouse

models.

4.2 Homozygous HdhQ150 knock-in mice display a rota-rod defect at 14

months.

4.3 Homozygous HdhQ150 knock-in mice display defects in locomotion at 12

months.

4.4 Homozygous HdhQ150 knock-in mice display defects in rearing at 12

months.

4.5 Homozygous HdhQ150 knock-in mice fail to gain weight.

4.6 Homozygous HdhQ150 knock-in mice develop a burrowing defect at 9

month.

4.7 The CAG tract found in the HTT gene carried by the HdhQ150 mouse

models was found to be incredibly unstable.

5.1 Intranasal Delivery.

77

88

89

91

102

105

106

107

119

77

130

131

132

133

134

135

143

74

75

77

5.2 Transport of intranasally administered compounds from the nasal

passage to the olfactory and trigeminal nerve bundles.

5.3 GFP is expressed in the hippocampi of mice treated with three doses of

either EV or MW1.

5.4 GFP is expressed in the cortex of mice treated with three doses of

either EV or MW1.

5.5 The MW1 PRNP knockdown construct failed to significantly reduce

prion levels brain regions of wild type C57BL6N mice.

5.6 Treatment of homozygous prion inoculated mice with MW1 lentivirus

at 3 and 4 weeks post inoculation extends life span of some mice.

5. 7 Prion infected homozygous mice treated with MW1 at 3 and 4 weeks

post inoculation displayed a similar degree of spongiosis at 7 weeks

compared to EV treated mice, while terminally ill mice exhibited a reversal

of spongiosis.

5.8 The early treatment of prion inoculated hemizygous mice with MW1

construct at 2, 3 and 7 weeks post inoculation appears to extend life span

of some mice.

5.9 Prion infected hemizygous mice treated with MW1 at 2, 3 and 7 weeks

post inoculation displayed a delay in spongiosis relative to EV treated and

RML only, untreated mice.

5.10 Mice treated with MW1 at 2, 3 and 7 w.p.i displayed no significant

increase in PSD-95 relative to NBH or RML only treated mice at 9 w.p.i.,

though they exhibited a greater level of SNAP-25.

5.11 PrPC levels in mice administered the MW1 lentiviral construct at 2, 3

and 7 w.p.i appeared to be similar to those displayed by RML only treated

mice at 9 w.p.i, while PrPSc

levels seemed to be elevated.

5.12 Prion infected hemizygous mice treated with low doses of MW1 at 4,

5 and 6 weeks post inoculation displayed no significant increase in survival

compared to EV and RML only treated animals

5.13 Treatment of homozygous mice with chitosan, rifampin and TMD

failed to improve the survival of mice intranasally administered with MW1.

145

163

164

165

166

167

168

169

170

171

173

174

5.14 Treatment of hemizygous prion inoculated mice with MW1 at 6, 7 and

8 weeks post inoculation extends life span.

5.15 Prion infected hemizygous mice treated with MW1 and EV at 6, 7 and

8 weeks post inoculation displayed comparable levels of spongiosis relative

to RML only, untreated mice.

175

176

1

Chapter 1

General Introduction

1.1 Introduction

Neurodegenerative diseases are defined as disorders of the nervous system, which are

progressive and lead to a decline in both cognitive and motor functions. These disorders

often stem from either a sporadic or genetic origin, the latter of which has been the

focus of extensive research with the use of many genetic disease models, both in vitro

and in vivo. One of the prominent features of many neurodegenerative diseases is the

misfolding of disease-causing proteins into stable non-native states. Such

conformational changes can yield fibrillar deposits, which accumulate in the cell and

interfere with a multitude of cellular processes, ultimately leading to loss or atrophy of

neuronal cells (Soto, 2003, Forman et al., 2004).

Given the increase in average lifespan, the number of individuals aged over 65 has

risen steadily in recent years, predominantly attributed to medical advances.

Unfortunately this increase is coupled with a higher prevalence of neurodegenerative

diseases, and in turn a greater financial drain on the health service. It is believed that

by 2050, the number of people suffering from disorders such as Alzheimer’s disease,

Parkinson’s disease and other neurodegenerative will triple (Ross and Poirier, 2004).

As such, it is imperative that our understanding of genetic factors influencing onset

and severity of these crippling conditions, as well as the mechanisms underlying

pathogenesis, are dissected.

1.11 Huntington’s Disease

Huntington’s disease (HD) is an autosomal-dominant neurodegenerative disorder,

stemming from the expansion of a CAG repeat (encoding glutamine) within exon one

of the HTT gene located on human chromosome 4 (Landles and Bates, 2004). Repeat

lengths beyond 39, lead to full penetrance of disease symptoms, with age of onset of

this disorder inversely correlated to the number of these repeats (Landles and Bates,

2004, Duyao et al., 1993). The expansion of the CAG tract occurs due to the instability

of the mutated gene during meiosis, with larger amplifications being attributed to the

2

greater mutation rates associated with spermatogenesis, which are subsequently

conveyed to offspring with paternal transmission (Gil and Rego, 2008). As such,

subsequent generations display larger number of repeats relative to their parent,

which combined with an increased disease severity, produce a phenomenon known as

anticipation, (Ridley et al., 1988).

Motor and cognitive dysfunction, progressive dementia and impairment of mental

processes such as reasoning and judgement are all synonymous with HD, with many

motor symptoms being attributed to the selective degeneration of the caudate and

putamen (Zuccato et al., Landles and Bates, 2004). Degeneration predominantly

affects the spiny GABAergic neurons of the striatum, with early neuronal dysfunction

observed in neurons linked to the external regions of the globus pallidus. This

neuronal pathway is believed to suppress unwanted movements following cortical

stimulation, disruption of which leads to chorea, a defining characteristic of this

condition. While striatal degeneration is one of the hallmarks of HD, other brain

regions are also affected, albeit to a much lesser degree than the striatum (Eidelberg

and Surmeier, Walker, 2007).

Expansion of the CAG tract in the HTT gene, results in an extension of the

polyglutamine (polyQ) tract within the encoded protein, huntingtin (HTT). Upon

lengthening of the polyQ domain, the protein gains a propensity to misfold, leading to

the formation of protein aggregates. The presence of such mutant HTT aggregates in

the nucleus as well as the cytoplasm is a hallmark of HD, and is believed to contribute

to disease pathology (Zuccato et al. 2010).

1.12 Huntingtin

HTT is a 348kDa protein consisting of 3144 amino acids, and although ubiquitously

expressed, the normal function of the protein remains contentious (see Figure 1.1)

(Cattaneo et al., 2005). The amino terminal domain (amino acids 1-17) of HTT adopts

an amphipathic α-helical structure and is believed to mediate interactions with

cellular membranes encasing organelles such as the mitochondria, Golgi apparatus

and endoplasmic reticulum (Rockabrand et al., 2007, Atwal et al., 2007). This

membrane targeting capacity is especially important during periods of stress, with

3

phosphorylation of serine residues 13 and 16 inducing a conformational change in the

α-helical structure of this region, and the subsequent targeting of HTT to subregions

within the nucleus. Mutant HTT is found to be hypophosphorylated at these residues,

and as such fails to display this translocatory response to stress. (Atwal et al., 2011).

The amino terminal region is followed by a polyglutamine (polyQ) domain, a salient

feature of HTT that governs its propensity to aggregate in mutant forms. Non-

pathogenic polyQ tracts in HTT adopt an alpha helical structure, a form believed to

elicit interactions between HTT and proteins containing a Src homology 3 (SH3) or

tryptophan (WW) domain, the latter of which is a common feature of several splicing

factors (Lin et al., 2004, Gao et al., 2006b, Bocharova et al., 2009, Bugg et al., 2012,

Chellgren et al., 2006).

The polyQ domain of HTT is adjacent to a variable polyproline (polyP) domain that

conforms to a polyproline II helical structure. This region is implicated in the

enhancement of membrane binding, and acts as a cis-acting modulator of the polyQ

tract structure itself (Kim et al., 2009b, Burke et al., 2013). The polyP domain inhibits

fibrillation by inducing the polyQ tract to adopt a structure similar to its own,

Figure 1.1. Schematic diagram of HTT. The 3144 amino acid protein contains a poly-Q

tract (red rectangle) starting at the eighteenth amino acid from the amino terminus,

this is followed by a proline-rich region (blue rectangle). HTT contains several HEAT

domains (orange rectangles), a nuclear export signal (NES) as well as caspase (blue

arrows) and calpain (purple arrows) cleavage sites. Adapted from (Cattaneo et al.,

2005)

4

preventing the formation of aggregation prone β strand/ β-turn conformations (Poirier

et al., 2005, Darnell et al., 2009). Deletion of this region therefore exacerbates the

aggregation and toxicity of mutant HTT (Dehay and Bertolotti, 2006).

Both polyQ-flanking regions are believed to interact with each other, mediated by

protein kinase C and casein kinase 2 substrate in neurons 1 (PACSIN1) and dependent

on the flexibility of the polyQ region. Expansion of the polyQ tract diminishes the

flexibility of this region, preventing the two flanking regions from interacting. Given

the importance of intramolecular interactions, the loss of flexibility within the polyQ

region is likely to impair its capacity to function and in turn exacerbate HTT

aggregation (Caron et al., 2013).

The HTT protein also possesses three highly conserved clusters made up of

approximately 37 HEAT (huntingtin, elongation factor 3, regulatory A subunit of

phosphatase 2A and TORI) repeats, each of which can adopted an alpha-rod structure.

These repeat sequences facilitate interactions with numerous proteins including

huntingtin interacting proteins (HIP) 1, 4 and huntingtin associated protein 1 (HAP1),

which are involved in neuronal membrane trafficking, a process impaired in HD

(Bossy-Wetzel et al., 2004, Cattaneo et al., 2005). These domains are also capable of

associating with themselves, which may contribute to both intramolecular folding

within HTT and the formation of homodimers through intermolecular interactions

(Palidwor et al., 2009).

The main body of the HTT protein contains at least three caspase and two calpain

cleavage sites, and as such is prone to proteolytic cleavage. Activities of these

respective protease families (caspase 2, 3 and 6, and calpain 1, 5, 7 and 10) increase

with the greater expansion of the polyQ domain, and give rise to small amino-terminal

fragments. These fragments exhibit a greater tendency to form aggregates than the

full length protein and may be a contributing factor in the pathogenesis of HD (Gafni

and Ellerby, 2002, Gafni et al., 2004, Wellington et al., 2002).

5

1.13 Normal HTT function and mutant HTT

HTT is ubiquitously expressed, with higher levels found in the central nervous system

(CNS) and the testes. Within the cell, the protein is localised in both the nucleus and

cytoplasm, associating with the Golgi, the endoplasmic reticulum, as well as being

positioned at synapses (Cattaneo et al., 2005).

Due to the low level of sequence homology when compared to other cellular proteins,

the normal endogenous function of HTT is unclear (Landles and Bates, 2004). Despite

this uncertainty, the presence of HTT within the cell appears to be crucial, as

illustrated by the early embryonic lethality seen in mouse models following loss of the

functional protein (Nasir et al., 1995a). Reduced levels of wild-type HTT also result in

abnormal brain development and neuron loss, thus implicating the HTT protein in

neurogenesis, neuron maturation and apoptosis (Metzler et al., 1999, Auerbach et al.,

2001, Cattaneo et al., 2005).Furthermore increased expression of wild-type HTT,

following toxic stimuli such as ischemic pressure and mutant HTT toxicity, conveys

neuroprotection by reducing apoptosis (Leavitt et al., 2001, Zhang et al., 2003).

Wild-type HTT interacts with a wide range of cellular proteins as well as influencing

the activity of proteins indirectly. One such protein is brain-derived neurotrophic

factor (BDNF), a protein synthesized in the cortex and transported to striatal neurons

via anterograde transport, where it stimulates the release of glutamate, and thus

promotes neuron survival by preventing excitotoxicity. HTT interacts with BDNF

indirectly, influencing its transport and expression (see Figure 1.2). Consequently a

reduction in the wildtype HTT protein, through depletion or mutation, results in a

parallel reduction in BDNF and in turn striatal degeneration (Gauthier et al., 2004,

Cattaneo et al., 2005).

6

HTT can be found at synaptic terminals, where it interacts with a number of proteins

involved in vesicular trafficking and in turn synaptic transmission (Cattaneo et al.,

2005). HTT is often found to be associated with post synaptic density 95 (PSD-95), a

scaffolding protein shown to interact and cluster the kainate and N-methyl-D-

Figure 1.2. Putative roles of wildtype and mutant HTT in brain-derived neurotrophic

factor (BDNF) synthesis and transport. (A) Wild-type HTT binds to and sequesters

repressor element 1- silencing transcription factor (REST complex), preventing its

binding to neuron-restrictive silencer elements (NRSE), thus permitting transcription

of BDNF. In addition to influencing transcription, HTT governs the anterograde

transport of BDNF from cortical cells to striatal cells. HTT forms a motor complex with

HTT-associated protein 1 (HAP1), which in turn binds to dynactin and BDNF vesicles.

HAP1 finally binds dyneins, allowing transport along cellular microtubules.

Furthermore phosphorylation of serine-421 is believed to stimulate this anterograde

transport. (B) Mutant HTT is unable to sequester the REST complex, which

consequently translocates to the nucleus and binds to NRSE, inhibiting the

transcription of BDNF. Increased affinity between mutant HTT and HAP1 also weakens

interactions within the motor complex, attenuating BDNF transport. Unlike wild-type

HTT, the mutant protein is often found to be unphosphorylated, which is believed to

promote retrograde transport. Adapted from (Zuccato et al., Cattaneo et al., 2005).

7

aspartate receptors (NMDAR), regulating glutamate induced excitatory

neurotransmission. The interaction between HTT and PSD-95 is abrogated in mutated

forms, leading to increased levels of glutamate stimulated excitotoxicity, a major

contributing factor of neurodegeneration seen in HD (Sun et al., 2001).

Glutamate associated excitotoxicity is especially prominent in the medium spiny

neurons of the striatum, a region of the brain shown to be particularly vulnerable

during HD pathogenesis. This vulnerability is believed to arise from a high prevalence

of NMDARs found at the synaptic membrane of these neurons. Excitotoxicity stems

from an excess of extracellular glutamate, chiefly brought about by a reduction in the

glutamate uptake transporter, GLT1 (Miller et al., 2008). The build-up of this

neurotransmitter leads to the continued stimulation of the NMDARs and a persistent

influx of calcium ions. Unable to cope with the large increase in cellular calcium ions,

this class of neuron endures substantial damage through the activation of various

calcium ion dependent pathways (Raymond et al., 2011). This aspect of HD has

become a major target for therapeutic intervention, predominantly via the

modulation of the NMDAR activity. One such method shown to hold promise is the

manipulation of the tryptophan degradation pathway, largely through the inhibition of

one the degradative enzyme, kynurenine-3-monoxogenase (KMO). Genetic and

pharmacological inhibition of KMO has been demonstrated to increase levels of the

neuroprotective NMDAR agonist, kynurenic acid (KA) (Campesan et al., 2011, Zwilling

et al., 2011). This increase in KA is associated with suppression of not only HTT toxicity,

but other forms of neurodegeneration such as Parkinson’s (PD) and Alzheimer’s

disease (AD)(Labbadia and Morimoto, 2013, Zwilling et al., 2011).

Though some symptoms associated with HD are attributed to the loss of endogenous

HTT function, mutated forms of this protein may acquire novel dominant negative

cellular functions. Such toxic gain of function has been shown to disrupt a number of

cellular pathways, either via aberrant interactions with components of these

pathways, or through sequestration (Gil and Rego, 2008). The expanded poly-Q tract

causes the HTT protein to form aggregates, which subsequently sequesters

transcriptional factors possessing glutamine rich regions, such as CREB binding protein

(CBP). Depletion of this histone acetyltransferase CBP leads to a reduction in

8

transcriptional activation of CRE regulated genes, many of which relate to neuronal

survival (Hughes, 2002, Jiang et al., 2006). Aggregated HTT, along with sequestered

cellular proteins, form structures known as inclusion bodies (Rajan et al., 2001).

The soluble mutant forms of HTT, as opposed to the aggregated forms, also hold the

capacity to sequester transcription factors. One such extensively studied protein is

specificity protein 1 (SP1), which reduces the expression of nerve growth factor

receptor and dopamine D2 receptor, both of which are found to be down-regulated in

HD (Dunah et al., 2002, Li et al., 2002)

The toxic gain-of-function theory in regards to HTT is eloquently illustrated by the

conditional model generated by Yamamoto et al. This study showed that a delayed

knockout of the mutated HTT gene, reversed both behavioural deficits and the

formation of inclusion bodies, thus suggesting a constant requirement of the mutated

protein to maintain a HD phenotype (Yamamoto et al., 2000).

1.14 HD and protein homeostasis

The onset of HD is associated with the impairment of numerous pathways and cellular

processes, many relating to protein homeostasis. This encompasses polypeptide

folding, modification and degradation of all proteins within the cell. Disruption of

these crucial cellular functions is brought about in a variety of ways, including the

sequestration and down-regulation of chaperone proteins. During the advancement of

HD there is a gradual decline in chaperone proteins such as DNAJ, heat shock protein

40 (HSP40) and heat shock protein 10 (HSP70), which exacerbates the misfolding of

mutant HTT (Hay et al., 2004). Overexpression of several of these chaperone proteins

has been shown to suppress mutant HTT toxicity, through a reduction in aggregation

coupled with an increase in degradation (Muchowski and Wacker, 2005, Jana et al.,

2000, Kobayashi and Sobue, 2001). Recently heat shock transcription factor 1 (HSF1), a

master transcriptional regulator of several chaperone proteins, has been recognised

as a potential target in the suppression of mutant HTT aggregation. When exposed to

stress, eukaryotic cells can activate the expression of chaperones proteins such as

HSP25, HSP40 and HSP70 through the actions of HSF1. While the upregulation of

single chaperone protein has been shown to be protective, increasing several

9

chaperones simultaneously appears to convey a synergistic protection. (Neef et al.,

2010, Fujikake et al., 2008). Furthermore a number of compounds capable of

activating expression of this therapeutic transcription factor have been identified,

including a barbituric acid-like compound (F1) This compound is believed to not only

reverse defects in protein homeostasis, but also holds the potential to reverse similar

proteomic deficits seen in other neurodegenerative diseases (Neef et al., 2011,

Labbadia and Morimoto, 2013, Anckar and Sistonen, 2011).

In addition to an impairment in protein folding, mutant HTT causes defects in the

ubiquitin proteasome system (UPS), one of the cellular protein degradation systems.

To prevent aberrant functioning, misfolded or damaged proteins are identified by

chaperones proteins and polyubiquitylated by E3 ubiquitin ligases. Proteins possessing

these ubiquitin marks are recognised and subsequently degraded by the large multi-

subunit complex, the proteasome (Mitra and Finkbeiner, 2008). Although mutant HTT

is found to be polyubiquitylated, it fails to undergo degradation and eventually

accumulates in the cytoplasm, leading to a build-up of the aggregated forms. Although

initially believed to directly inhibit the proteasome function, recent studies suggest

that mutant HTT predominantly affects proteostasis through the sequestration of

chaperone proteins (Venkatraman et al., 2004, Hipp et al., 2012). A drop in chaperone

activity is accompanied by an increase in aberrantly folded HTT and also other

misfolded cellular proteins. This loss of protein homeostasis overwhelms the

proteasome, which becomes incapable of coping with the rising number of

polyubquitylated proteins marked for degradation (Hipp et al., 2012). This is

supported by observations derived from conditional HD mouse models. A loss of

mutant HTT expression is believed to provide the proteasome an opportunity to

degrade the surfeit of misfolded proteins, leading to a decline in mutant HTT

inclusions (Yamamoto et al., 2000).

10

1.15 Aggregation Dynamics

The propensity of mutant HTT to undergo aggregation stems from the lengthening of

the polyQ region, which paradoxically results in a reduction in the inherent flexibility

of this domain (Caron et al., 2013). This increased rigidity has been attributed to the

hydrogen bond interactions of the glutamine amide side chains, which intrinsically

form a polar zipper structure as a means of avoiding the less favourable interaction

with water (Crick et al., 2006). As the polyQ tract expands, these side chains

interactions increase, leading to compaction of the domain, which thus limits the

interactions of the two flanking regions (Perutz et al., 1994, Crick et al., 2006).

Consequently the capacity of the polyP region to maintain the α-helical structure of

the polyQ region diminishes and without such a regulatory effect, the latter region

adopts a more aggregation prone β-sheet conformation, a structural change enhanced

by the amino terminal domain (Tam et al., 2009, Thakur et al., 2009).

While the amino terminal of HTT facilitates the conformational change in the polyQ

tract, this repeat region reciprocally induces a polyQ length-dependent extension of

the alpha helical structure of the amino terminal domain, which reduces its interaction

with other regions of the HTT protein. This extension brings about the initial step of

aggregation, with the amino terminal domains interacting and forming a core

structure, allowing the polyQ regions of adjacent HTT monomers into close proximity

(see Figure 1.3) (Thakur et al., 2009, Dlugosz and Trylska, 2011, Lakhani et al., 2010).

The neighbouring polyQ domains can subsequently polymerise via hydrogen bonding

between both the polar side chains and main chain amides, forming a beta-spine

structure (Nelson et al., 2005). The interactions of nearby polyQ domains is putatively

cemented through domain swapping, a mechanism by which part of a tertiary peptide

can spatially replace that of an identical peptide, effectively interlocking the two

peptides (Newcomer, 2001, Bennett et al., 2006). This gives rise to the formation of

the aggregation nucleus, an energetically unfavourable structure, representing the

rate limiting step, or lag phase, of HTT aggregation. The salience of the amino terminal

in this initial stage has been demonstrated through the disruption or mutation of this

region, the result of which being a substantial reduction in aggregation (Tam et al.,

2009, Kelley et al., 2009). Despite the importance of the amino terminal domain,

11

mutant HTT can also undergo this initial nucleation step via a minor pathway

mediated by the association of beta-sheets of the polyQ domains (Kar et al., 2011,

Jayaraman et al., 2012).

Nucleation leads to the sequential formation of oligomers rich in β-strands, and

subsequently mature amyloid fibrils (Slepko et al., 2006, Kim et al., 2009b). Although

the mechanism underlying the elongation stage of mutant HTT aggregation is widely

debated, some theories suggest the “dock and lock” model, in which monomers bind

to the nucleated intermediate, stimulating an intermolecular conformational change

that “locks” the monomer into the nascent aggregate (Reddy et al., 2009). Other

theories suggest the formation of oligomeric intermediates which come together to

form insoluble rod-like structures or fibrils, with the pool of oligomers diminishing as

the larger fibrils form, at a rate correlating with the length of the polyQ region.

(Legleiter et al., 2010).

The aggregation pathway appears to be devoid of one unifying pathway, instead being

made up of a number of branches leading to the eventual formation of structures of

various sizes and shapes, including amorphous globular aggregates, mature fibrils and

even annular structures. Recent advances in microscopy techniques such as atomic

force microscopy, has provided a means of distinguishing between these various

species on the basis of their morphology, with the potential to examine the different

conditions necessary to bring about such aggregate diversity (Burke et al., 2011,

Polling et al., 2012, Legleiter et al., 2010).

The aggregation pathway gives rise to a number of misfolded aggregate forms, and

although controversial, the soluble oligomers and protofibrils are widely regarded as

the toxic species. These aggregate forms are believed to dictate the survival of

neurons, putatively through generation of reactive oxygen species or initiation of cell

death pathways (Wyttenbach et al., 2002, Arrasate et al., 2004). Inhibition of

oligomerisation, following onset of disease symptoms, has been shown to ameliorate

motor function, survival and also stimulate the clearance of mutant HTT in vivo, thus

emphasising the contribution of oligomerisation in pathogenesis (Sanchez et al.,

2003). The formation of larger insoluble aggregates, or inclusion bodies, may

12

represent a neuroprotective mechanism employed by the cell to reduce the level of

soluble toxic species, while the retained capacity to sequester other cellular proteins

may simply reflect an unavoidable consequence (Arrasate et al., 2004). Indeed, the

presence of these larger aggregates appears to increase the survival of certain cell

lines expressing mutant HTT, compared to cells displaying a diffuse spread of the

oligomeric forms (Arrasate et al., 2004, Bodner et al., 2006).

Figure 1.3. HTT aggregation dynamics. Amplification of the polyQ region within

HTT conveys a propensity to aggregate. Aggregation is believed to take place in an

ordered manner, beginning with the nucleation of monomers, via the region

preceding the poly-Q tract. Nucleation is followed by elongation, ultimately

leading to the formation of mature amyloid fibres. The protofibrils, and to some

extent the oligomers and monomers, are believed to represent the toxic species

implicated in HD, while the mature fibres are potentially neuroprotective. This

pathway contains several of branches, including the formation of large amorphous

aggregates and the more rare annular structure (Adapted from Zuccato et al

(2010) and Polling et al (2012))

13

Given the clear importance of aggregation in the progression of HD, modulating the

dynamics of aggregation, either by encouraging the formation of the mature

aggregates or inhibiting the oligomerisation of the putatively toxic precursor species,

has been shown to hold therapeutic potential. The compound B2 for instance, has

been recognised to diminish pathology in cellular and Drosophilia models of expanded

polyQ disorders and Parkinson’s disease, an effect accompanied by an increase in

inclusion body formation. Although the exact target of the B2 compound is unknown,

authors postulated a possible restoration of proteasomal function (Bodner et al.,

2006, Palazzolo et al., 2010). Other compounds such as green tea (-)- epigallocatechin-

gallate (EGCG), methylene blue and trehalose, have all been shown to inhibit early

aggregation and improve disease relevant phenotypes in numerous HD models. These

potential aggregation suppressors are principally believed stabilise the structure of

mutant HTT, preventing the conformational changes that initiate aggregation

(Ehrnhoefer et al., 2006, Tanaka et al., 2004, Sontag et al., 2012).

1.16 mRNA splicing and alternative splicing

The use of genetic modifier screens has provided a wealth of research relating to the

various pathways and cellular processes altered in HD, and can include both human

genomic studies and also candidate screens performed in other genetic models. These

modifiers are genes, usually with some inherent variation in expression or structure,

which can alter the progression of HD symptoms (Gusella and MacDonald, 2009). Both

forms of screening often yield large amounts of candidate genes, which are

subsequently grouped in terms of their functionality. Groups including protein

homeostasis and transcription are generally found to be enriched, and as such prompt

further study. Other groups such as the post-transcriptional process of mRNA splicing

often appear in genetic screens of HD, but have incited very little research (Nollen et

al., 2004, Lejeune et al., 2012, Kaltenbach et al., 2007).

The human genome is comprised of around 23,000 protein encoding genes, a number

similar to less complex organisms such as Arabidopsis Thaliana. Our higher level of

complexity stems from the greater manipulation of our genes at a post-transcriptional

level (Mills and Janitz, 2012b). The majority of genes in the human genome are found

14

in a fragmented form, consisting of coding regions known as exons interspersed with

non-coding intron regions. Following transcription, the intronic regions are removed

from nascent pre-mRNA and the exonic regions are brought together to form a

continuous open reading frame via mRNA splicing (Singh and Cooper, 2012, Han et al.,

2011).

mRNA splicing is catalysed by a large dynamic complex known as the spliceosome, the

core of which consists of five small nuclear ribonucleoproteins (snRNP), U1, U2, U4, U5

and U6, in addition to numerous splicing factors such as splicing factor 1 (SF1) and U2

auxiliary factor (U2AF). Assembly of the complex is guided by the presence of the

splice consensus sequences flanking the intronic region (3’ and 5’ splice sites), the

branch site adenosine (BS) found within the body of the intron and a polypyrimidine

tract (PT) found downstream of the BS (Singh and Cooper, 2012). Splicing is further

regulated by the presence of highly variable auxiliary elements found within the intron

and exon units, including intronic splicing silencers (ISS), intronic splicing enhancers

(ISE), exonic splicing silencers (ESS) and exonic splicing enhancers (ESE). Many of these

latter elements serve as binding sites for the multitude of splicing factors involved in

splicing, which not only guide spliceosome formation, but facilitate exon recognition

across intronic regions (Pagani and Baralle, 2004, Berget, 1995, Singh and Cooper,

2012).

Upon the stepwise assembly and activation of the spliceosome, the complex removes

the intronic region and ligates the two adjacent exonic regions via two

transesterification reactions (see Figure 1.4). The resulting excised intron, in the form

of a lariat structure, is subsequently degraded while the spliceosome components

disassemble for further splicing events (Singh and Cooper, 2012, Matlin et al., 2005).

The majority of splicing in humans occurs via the assembly of the major spliceosome

described above, known as the U2 spliceosome pathway. However a minor U12

pathway, which is absent in yeast and nematodes though present in some lower

eukaryotes, accounts for a small fraction of pre-mRNA splicing (humans possess

around 621 introns excised from pre-mRNA with the aid of this pathway) (Will and

Luhrmann, 2005, Lin et al., 2010). The U12 pathway makes use of less stringent splice

15

consensus sequences, and generally targets introns lacking a PT and displaying a BS

site closer to the 3’ splice site compared to the more ubiquitous U2 pathway. This U12

spliceosome consists of homologous snRNPs, with the U1, U2, U4 and U6 proteins of

the U2 spliceosome replaced by U11, U12, U4atac and U6atac respectively (atac

denotes what was initially thought to the be the U12 consensus sequences for the 3’

and 5’ splice sites, but subsequently refuted). Despite the differences, the minor

spliceosome shares many of the auxiliary factors, as well as the U5 snRNP, with its

more abundant sister complex. These similarities provide support for the belief that

this pathway may reflect an early evolutionary form of splicing (Singh and Cooper,

2012, Will and Luhrmann, 2005, Lin et al., 2010).

16

17

The exclusion or inclusion of different exons during the splicing of pre-mRNA allows

the formation of a myriad of mature mRNA transcripts, each giving rise to protein

isoforms with varying functions or specificities (see Figure 1.5a). This phenomenon is

referred to as alternative splicing, and occurs in approximately 90% of multi-exon

genes (Singh and Cooper, 2012). The variation in transcripts can be attributed to a

number of different alternative splicing events, including mutually exclusive exons and

the use of competing splice sites. Splicing can also act concertedly with the promoters

and end processing sites linked to certain exons, illustrating a regulatory role (see

Figure 1.5b) (Matlin et al., 2005).

Figure 1.4. Pre-mRNA splicing mechanism. Pre-mRNA splicing is achieved through the

actions of a dynamic complex known as the spliceosome. This complex consists of a large

number of components, which affords the spliceosome a high level of specificity in

addition to its catalytic properties, one of the salient constituent groups being the small

nuclear ribonucleoproteins (snRNP). U1snRNP binds to the 5’ splice site (which has a

general GU consensus sequence found in the intronic region), while U2snRNP binds to the

branch site (an unpaired conserved A nucleotide in the intron). Splicing factors serve to

bridge these two snRNPs, allowing for formation of the inactive pre-spliceosome, and in

turn binding of the U4/U6.U5 tri-snRNP. Unbinding of U4 snRNP leads to catalytic

activation of the spliceosome, a two-step reaction in which the 2’ hydroxyl group of the

branch site adenosine, performs a nucleophilic attack onto the 5’ splice site, leading to the

formation of the lariat structure. This structure facilitates the second step of splicing, in

which the 5’ hydroxyl group of the 5’ splice site carries out a second nucleophillic attack

on the 3’ splice, resulting in excision of the intronic region and ligation of the exons.

(Adapted from Wahl, Will et al., 2009)

18

A.

B.

19

The regulation of alternative splicing is achieved through the cumulative actions of

multiple splicing factors, which bind to auxiliary elements, and subsequently influence

spliceosome assembly. These splicing factors identify and distinguish constitutive

exons from those which are alternatively spliced, and by extension govern the ratios

of different transcripts generated from a single multi-exon gene (Matlin et al., 2005).

Classes of splicing factors include the SR proteins, characterised by a string of serine

and arginine repeats and one or more RNA recognition motifs (RRM). These proteins

predominantly bind ESEs and bring this region into contact with the BP, a crucial step

in the activation of splicing (Shen and Green, 2004). Another class of splicing factors,

the heterogeneous nuclear ribonuclear ribonucleoproteins (hnRNPs), also contain

RRMs as well as K-homology type (KH) RNA binding domains. These factors are known

to bind ESSs and ISSs and in turn repress splicing events via the steric hindrance of

snRNP recruitment (though this mechanism remains contentious) (House and Lynch,

2006, Busch and Hertel, 2012). In many cases knockout of these splicing factors are

known to be lethal due to their specific and essential roles in alternative splicing

(Matlin et al., 2005).

Given the vital role of splicing factors in the correct splicing of immature mRNA, many

of these proteins are themselves regulated to maintain a state of homeostasis. For

example the genes encoding SR and hnRNP proteins contain highly conserved regions

which hold premature termination codons (PTCs) (Ni et al., 2007). These regions are

crucial for the negative autoregulatory feedback inherent to both families of proteins.

Incorrectly spliced transcripts encoding these proteins retain an in-frame PTC,

targeting them for degradation via nonsense mediated decay (NMD). HnRNP proteins

induce splicing repression onto transcripts of their own gene, leading to exon

Figure 1.5. Alternative splicing. A. Pre-mRNA can be spliced into a multitude of

mature mRNA transcripts through the inclusion or exclusion of certain exons. B. Pre-

mRNA can undergo several forms of mRNA splicing, including exon exclusion, the

preferential selection of one exon over another (mutually exclusive exons), and the

use of an alternative splice site within the same exon (completing 5’ and 3’ splice

sites). Changes in mRNA splicing can be modulated by varying promoters and can lead

to the selection of multiple poly(A) sites (Adapted from Matlin et al 2005).

20

exclusion which brings a downstream PTC into frame. SR proteins on the other hand

stimulate the inclusion of an additional exon which contains the PTC. In both cases,

the resulting transcripts are broken down and genes downregulated (see Figure 1.6)

(Kalsotra and Cooper, 2011, Ni et al., 2007).

Figure 1.6. Homeostatic regulation of the antagonistic splicing factor families: SR and

hnRNP proteins. The SR and hnRNP protein represent two families of splicing factors,

that have opposing effects on splicing. While the SR proteins generally promote exon

inclusion, the hnRNP proteins stimulate exclusion. Both families utilise a negative

autoregulatory feedback mechanism to maintain the correct levels of their protein

products, and by extension maintain the ratio of transcripts derived from their

respective target genes. Both splicing factor families regulate their own expression by

binding to their own transcripts and stimulating inclusion or exclusion of a crucial

exon. The resulting transcript contains an in-frame premature termination codon (pre-

x), which subsequently signals its degradation via nonsense mediated decay (NMD).

(Adapted from Kalsotra and Cooper (2011))

21

Although splicing factors play a significant role in establishing constitutive and

alternative splicing events, other nuclear processes can also influence mRNA splicing,

one of which being transcription. The initial recruitment of splicing factors, such as

SF2/ASF, to the site of transcription is attributed to the C-terminal domain (CTD) of

RNA polymerase II (RNAPII), with deletion or truncation of this region resulting in a

substantial reduction in mRNA splicing (Caceres and Kornblihtt, 2002, Hirose and

Manley, 2000, Yuryev et al., 1996). Phosphorylation of RNAPII CTD, a step that

prompts transition from the initiation complex to the elongation complex, is required

for the recruitment of splicing factors and the sequential binding the U1 and U2 snRNP

proteins. As the RNAPII moves along its target DNA, the recruited splicing factors may

then serve to identify splice sites on the nascent RNA (Caceres and Kornblihtt, 2002).

The hypophosphorylated form of the CTD is believed to inhibit this initial assembly,

and in turn reduce splicing (Hirose et al., 1999).

The elongation speed of RNAPII has also been postulated to affect splicing, with a

slower progression allowing sufficient time for the identification of exons. Such a slow

rate is conducive to the inclusion of exons, while a faster RNAPII may simply pass over

these alternatively spliced exons (Caceres and Kornblihtt, 2002). This theory has been

supported by work conducted in human cell lines tranfected with a fibronectin

minigene construct containing the alternatively spliced exon, EDI. These cells lines

were subsequently treated with the SV40 T antigen (T-Ag) to slow down the rate of

RNAPII elongation, which significantly increased inclusion of the EDI exon. Conversely

use of a viral transcriptional activator (VP16) to increase elongation rate, prompted

exclusion of the EDI exon (Cramer et al., 1997, Kadener et al., 2001). In vivo, changes

in the elongation rates of RNAPII are determined by the promoter, which is itself

regulated by transcriptional activators. It is therefore feasible that transcriptional

activators may provide a means of regulating splicing via this method (Caceres and

Kornblihtt, 2002).

22

1.17 mRNA splicing and neurodegenerative disease

Alternative splicing of mRNA within the nervous system is a highly regulated process,

given the complex nature of neurons, and the requirement for both spatial and

temporal changes in protein levels. Mutations in splicing sequence (cis-acting) and/or

genes encoding splicing factors (trans-acting) can have a major impact on the

functioning of neurons. Both classes of mutations can greatly and detrimentally

influence the ratio at which the various transcripts are generated from a single

immature mRNA. Indeed aberrant splicing has been seen in some neurodegenerative

diseases, and it is estimated potentially 50% of all genetic disorders stemming from

point mutations may bring about splicing defects (Licatalosi and Darnell, 2006, Matlin

et al., 2005).

One example of a genetic disorder attributed to a splicing abnormality is spinal

muscular atrophy (SMA), a fatal autosomal recessive disease attributed to the

degeneration of motor neurons. This condition is linked to a deficit in SMN, or survival

of the motor neuron protein which is believed to form a large complex that governs

the biogenesis of small nuclear ribonucleoproteins (snRNP), and in turn RNA

metabolism. Although two genes encode this protein, only the centromeric gene,

SMN1, leads to the formation of an adequate amount of protein. An inherent

mutation in the telomeric gene, SMN2, predominantly gives rise to a transcript lacking

exon 7, encoding a non-functional protein. Disease causing mutations in SMN1 lead to

a reduction in functional protein, a deficiency which is not compensated for by SMN2

(see Figure 1.7) (Dredge et al., 2001, Paushkin et al., 2002). The intrinsic splice

sequences in the SMN2 gene, rendering approximately 80% of transcripts superfluous,

can be overcome by certain splicing factors. For example, over expression of

transactivation responsive DNA-binding protein 43 (TARDBP), has been shown to

stimulate the inclusion of exon 7, in mature SMN2- derived transcripts, increasing the

levels of full length SMN by approximately two-fold. Interestingly TARDBP has recently

generated great interest due to its involvement in other neurodegenerative diseases

including HD, an inference based on colocalization with many disease causing proteins

(Schwab et al., 2008).Chemicals such as 5-(N-ethyl-N-isopropyl)-amiloride (EIPA),

23

which inhibit Na+/H+ exchangers, also increases exon 7 inclusion, through the possible

up-regulation of the splicing factor, SRp20 (SFRS3)(Bose et al., 2008, Yuo et al., 2008).

Other examples of cis-acting mutations associated with splicing include the autosomal

dominant disorder, frontotemporal dementia with Parkinsonism linked to

chromosome 17 (FTDP-17), a condition attributed to mutations in the splice sites of

the microtubule-associated protein tau (MAPT) gene, or components of the splicing

inhibitory structure that regulates it splicing. In both cases, mutations lead to an

increased generation of transcripts containing exon 10. This transcript contains an

additional microtubule binding domain, and exists in a delicate equilibrium with its

sister isoforms lacking this extra exon. Changes in this equilibrium results in neuronal

Figure 1.7. Splicing of survival of motor neuron 2 (SMN2). Mutations in the SMN2

gene (AC to T nucleotide change at position 6 of exon 7) result in exon skipping,

leading to the formation of an unstable protein, which is rapidly degraded. Inclusion of

exon 7 can be stimulated by splicing factors such as TARDBP and SFRS3, leading to the

synthesis of a functional protein (Adapted from Burnett and Summer, 2008).

24

dysfunction and ultimately cell death through a poorly understood mechanism (Lee et

al., 2001, Licatalosi and Darnell, 2006).

Neurological disorders can also arise due to mutations or changes in the expression of

splicing factors. One such example of trans-acting disorders are the paraneoplastic

neurological disorders (PNDs), in which cancer cells present elsewhere in the body

elicits an immune response mounted against antigens found not only on the tumour,

but also regions of the nervous system. Patients develop an autoimmune neurological

disease marked by a substantial neuronal degeneration (Musunuru and Darnell, 2001,

Roberts and Darnell, 2004). Nova is a neuronal splicing regulatory factor implicated in

several forms of PND and act as an antigen during the immune response mounted

against the tumour. This protein regulates the splicing of genes encoding a number of

synaptic proteins involved in synaptic formation and plasticity, which are consequently

downregulated in these disorders (Ule et al., 2006, Musunuru and Darnell, 2001).

Although splicing dysfunction in HD is unclear, recent studies have postulated several

theories by which an expansion in the CAG repeats of the HTT gene can bring about

defects in mRNA splicing. One such theory centres on the role of mutant RNA

transcripts, which are believed to form double stranded structures mediated by the

expanded CAG repeats. Akin to the previously well documented pathogenic role of

expanded CUG repeats in myotonic dystrophy type 1 (DM1), these structures are

thought to bind and sequester certain splicing factors, such as muscleblind-like 1

(MBNL1), and in turn prompt aberrant splicing of genes regulated by these proteins

(Mykowska et al., 2011, Birman, 2008). Furthermore expression of untranslated CAG

repeats in the neurons of Drosophilia has been found to induce motor dysfunction, a

shortened lifespan and brain degeneration, indicating that conditions arising from

nucleotide repeat expansions may by attributed, in part to the toxic properties of the

expanded repeat RNA (Li et al., 2008).

Abnormal splicing of the HTT gene has recently been implicated in HD pathogenesis,

with an expansion of the CAG repeat tract believed to prompt an increased interaction

between HTT RNA transcripts and the splicing factor, SRSF6. (Sathasivam et al., 2013).

This greater affinity stimulates the generation of partially spliced exon 1-intron 1

25

transcripts, containing a conserved translational stop codon within the non-coding

region. As a result, this transcript is translated into the extremely pathogenic exon 1

HTT protein fragment. Such aberrant splicing was shown to be governed by polyQ

length, with knock-in mice expressing repeat lengths of Q50, Q80 and Q100 displaying

much lower levels of the exon 1 transcript relative to animals expressing Q150 and

Q175, a result found to be consistent in juvenile HD patients (Sathasivam et al., 2013).

This study therefore suggests that dysfunctions in mRNA splicing may play a much

more crucial role in HD than previously thought.

1.18 Project aims

This PhD project is based upon a currently unpublished genetic modifier screen carried

out by a member of our research group, Dr Robert Mason. This screen identified ~90

mammalian cDNAs that suppress mutant HTT toxicity when over-expressed in yeast.

After grouping these cDNA hits in terms of their interactions as well as functionality, it

became apparent that many of the genes were key components of physiological

processes such as mRNA processing, apoptosis and energy metabolism. Although the

two latter groups have been studied and documented extensively in relation to HD,

research focused on mRNA processing in HD is fairly sparse (Zuccato et al., 2010).

The overall aims of this project were to validate the splicing cDNA hits in a range of

models, initially using a neuron-like HD cell model derived from a pheochromocytoma

of rat adrenal medulla, PC12 cells. Using this cell line, I generated stable cell lines

overexpressing our splicing gene hits, examined apoptosis with the use of a

luminescence based caspase activation assay, and aggregation dynamics via an

automated imaging based platform. As each of the gene hits have a role in mRNA

splicing, also interrogated these candidates using a dual reporter splicing efficiency

assay.

I also employed a C. elegans HD model to bridge the gap between the PC-12 cell line

and mouse model of HD. This model exhibits neurodegeneration, behavioural deficits,

and formation of mutant HTT aggregates which increase with age. I employed an RNAi

approach to knockdown gene expression of the worm orthologues corresponding to

our candidate genes, and examined for alterations in HD phenotypes, notably motor

26

function and the aggregation, in hopes of providing justification for progression of

promising candidates into mice.

As the ultimate goal with this work is preclinical validation and mechanistic dissection

of HD model mice, I have also undertaken analyses to set the stage for future

work.The most promising of the candidate genes will ultimately be overexpressed in

these murine models to determine whether increased expression ameliorates the HD-

relevant phenotypes. Thus, I have characterised a panel of phenotypes in two HD

mouse models, including a novel behavioural paradigm known as the burrowing assay.

I also performed extensive optimisation of an intranasal delivery technique for

lentiviruses, in order to mediate overexpression of gene candidates.

27

Chapter 2

Exploring the functional role of splicing factors in models of

Huntington’s disease

2.1 Introduction

Gene ontology analysis of hits from the yeast genetic modifier screen carried in the

laboratory by Dr Robert Mason, uncovered nine genes implicated in mRNA splicing

(unpublished; see Chapter 1). Here were aimed to characterise and validate these

splicing genes in our HD cellular models.

The majority of these genes belong to the HNRNP class of splicing factors, with the

remainder largely encoding structural components of the spliceosome. Many of the

splicing factors display similar cellular roles, and as such share some structural

similarities (see Figure 2.1).

28

Figure 2.1. Schematic representation of the murine structures of each of the putative

HTT suppressor proteins. CRNKL1 contains a NLS (nuclear localisation signal) domain, as

well as polyproline region (poly P) and 16 HAT (half a TPR) repeat, the latter of which

facilitate protein to protein interactions within large molecular complexes. EFTUD2 is a

GTPase made up of five domains (G domains). HNRNPF, HNRNPK and HNRNPU contain

either RNA recognition motifs (RRM) or K homology RNA binding domain (KH), all of

which facilitate interactions with RNA transcripts. In addition to its RRM domain,

HNRNPQ also possesses a NLS domain. HNRNPU contains a number of domain relating to

its nuclear function, including a DNA binding SAP (SAF A/B, Acinus, PIAS) domain, an ATP

binding domain, an RNA binding domain rich in arginine and glycine (RGG) and a domain

of unknown function, the SPRY domain. The splicing factor, SFRS3 also contains a single

RRM domain. SNRPB contains a repeat rich regions, which may serve to facilitate protein-

protein interactions. TARDBP contains three RNA binding domains, two RRM and one

RGG domain. All structural information was obtained from the Uniprot database.

29

2.11 Splicing suppressor genes

CRNKL1 (also known as CLF, CRN, Clf1, HCRN, SYF3 and MSTPO21)

The CRNKL1 (crooked neck pre-mRNA splicing factor 1) gene encodes a scaffolding

protein which is essential for spliceosome assembly. In humans, this role is believed to

be mediated by the numerous tetratricopeptide (TRP) motifs present throughout the

body of the protein. These motifs facilitate interactions between the constituent

proteins making up large complexes, such as the anaphase promoting complex (APC)

and indeed the spliceosome (Amada et al., 2003). During spliceosome assembly,

CRNKL is thought to interact with the U4/U5.U6 trimer and the prespliceosome

complex, facilitating the transition to mature spliceosome. Consistent with this role,

deletion or mutation of CRNKL1 results in the accumulation of pre-mRNA in vivo

(Chung et al., 1999, Chung et al., 2002). The importance of this protein during

development has been highlighted by the early embryonic lethality seen in mutant

Drosophila models, an effect attributed to splicing defects that are most prominent in

the central and peripheral nervous systems (Zhang et al., 1991, Amada et al., 2003,

Edenfeld et al., 2006).

EFTUD2 (also known as SNRP116, SNU114, MFDGA, MFDM and U5-116KD)

EFTUD2 (elongation factor Tu GTP binding domain containing 2) encodes the GTPase

Snu114. This protein regulates the unwinding of U4/U6, and therefore catalytic

activation of the spliceosome, and also the disassembly of the active spliceosome

following mRNA splicing (Bartels et al., 2002). In both cases, Snu114 brings about such

changes indirectly, modulating the activity of Brr2p, an ATPase and component of the

U4.U5.U6 trimer (Small et al., 2006). When bound to GTP, Snu114 activates Brr2p,

which in turn stimulates unwinding of U2/U6. Following spliceosome activation, the

GTP is hydrolysed and the GDP-bound Snu114 represses further Brr2p activity,

stabilising the spliceosome’s catalytic core. The sequential excision of the intronic

region signals for a reversion to the GTP bound state of Snu114, thereby lifting the

repression on Brr2p. This stimulates the release of the intron lariat and spliceosome

disassembly (Small et al., 2006, Frazer et al., 2009). Similar to CRNKL1, depletion of

EFTUD2 results in an accumulation of pre-mRNA, thus illustrating an integral role in

30

mRNA splicing (Fabrizio et al., 1997). Mutations in this gene have been identified as

the cause of mandibulofacial dysostosis with microcephaly (MFDM), a condition

synonymous with delays in brain development, a cleft palate and hearing loss (Lines et

al., 2012).

HNRNPF (also known as HNRPF)

HNRNPF encodes HNRNPF, a splicing factor and member of the hnRNP family of

proteins. Like many of the factors belonging to this family, HNRNPF predominantly

stimulates exon inclusion. This splicing factor binds to sequences containing a GGG

motif found at the end of an intron, and working in unison with other HNRNPs, such as

HNRNPA or HNRNPB, which bind to the opposing flanking end of this region. The

concerted actions of these HNRNPs define the intron, and facilitate the “looping out”

of the intron during mRNA splicing (Martinez-Contreras et al., 2006). HNRNPF is

regulated by the posttranslational modification of its GYR (glycine-tyrosine-arginine

rich) domain, a region that governs import into the nucleus via import transporter

importin 2 (Trn2) (Van Dusen et al., 2010). It has been postulated that in addition to

this general role in mRNA splicing, HNRNPF may also serve a more specific role. For

example HNRNPF is a crucial component of the neural-specific complex that regulates

the splicing of the proto-oncogene, c-Src, which encodes tyrosine kinase. In neuronal

cells, c-Src contains an additional SH3 (Src homology 3) domain, and antibody

inhibition of HNRNPF has been shown to lead to the exclusion of the exon encoding

this domain (Roskoski, 2004, Min et al., 1995). However, the functional consequences

of the inclusion of this domain in neuronal cells is unclear.

HNRNPK (also known as CSBP, HNRPK, TUNP, KBBP, NOVA)

HNRNPK encodes HNRNPK, which unlike other member of the HNRNP family has a

multitude of roles in addition to mRNA splicing including chromatin remodelling,

transcription and translation (Bomsztyk et al., 2004). Moreover, protein interaction

studies suggest that only 20% of HNRNPK is present in complexes implicated in RNA

splicing (Mikula et al., 2006). Of this relatively small proportion, HNRNPK regulates the

splicing of a number of genes encoding apoptotic proteins, including pyrin domain

containing 1 (NALP1) and B-cell lymphoma extra-large (BCL-xl) (Venables et al., 2008).

31

For example, HNRNPK was shown to elicit an exon exclusion event in NALP1. In the

case of the anti-apoptotic protein BCL-xl, HNRNPK represses the removal of a short

exon found between exons 2 and 3 (denoted exon 2.1) leading to formation of a pro-

apoptotic protein referred to as BCL-xs, a splicing event stimulated by its sister

protein, HNRNPF (Revil et al., 2009, Garneau et al., 2005). This illustrates the

bidirectional nature of HNRNPK as a splicing effector, and the opposing effects of

members of the HNRNP family (Venables et al., 2008).

Although HNRNPK has been implicated in a number of forms of cancer, including

colorectal cancer and carcinoma, it’s involvement in neurodegeneration has not been

investigated (Roychoudhury and Chaudhuri, 2007, Carpenter et al., 2006). However

several other proteins containing a K homology (KH) RNA-binding domain, a

characteristic feature of HNRNPK, regulate neural specific splicing. For instance the KH

domain contained within neuro-oncological ventral antigen 1 (NOVA-1) protein

regulates the splicing of the α2 subunit of glycine receptor (GlyRα2) and γ-

aminobutyric acid receptor type A γ2 subunit (GABA A Rγ2). Mice lacking NOVA-1

display symptoms similar to those seen in paraneoplastic opsoclonus myoclonus

ataxia (POMA), a condition stemming from a loss of inhibitory regulation of motor

neurons in the brainstem and spinal cord (Musunuru and Darnell, 2001, Dredge et al.,

2001). Further to this, overexpression of KH domain-containing mushroom body

expressed (mub), an essential splicing factor, exacerbates disease phenotypes in a

spinocerebellar ataxia type 1 Drosophila model (Grams and Korge, 1998, Fernandez-

Funez et al., 2000).

While mRNA splicing may be the main function of many of the other member of the

HNRNP family, HNRNPK has primary roles in both transcription and translation. For

example, HNRNPK is involved in the transcriptional regulation of the oncogene, c-myc.

HNRNPK binds to the C-rich (CT) element located upstream of this gene and the

RNAPII machinery, stimulating transcription (Michelotti et al., 1996). Interestingly,

while hnRNPF regulates expression of c-Src at a splicing level, hnRNPK governs the

translation of this enzyme, binding to the 3’ untranslated region of the mRNA and

preventing ribosome formation (Naarmann et al., 2008).

32

HNRNPQ (also known as GRYRBP, SYNCRIP, NSAP1, PP68)

HNRNPQ encodes for HNRNPQ, a splicing factor known predominantly to induce exon

inclusion (Venables et al., 2008). One of the several gene targets linked to HNRNPQ is

SMN2, the inherently mutated gene which is exposed following mutations within its

sister gene, SMN1 in the condition SMA (see Figure 1.7). HNRNPQ binds exon 7 of the

SMN2 mRNA transcripts, and stimulate exon inclusion, leading to the formation of the

functional protein (Chen et al., 2008).

In addition to a role as a splicing factor, hnRNPQ has been implicated in the regulation

of translation and localisation of mRNA transcripts. In the former role, HNRNPQ binds

to the 5’ untranslated region of mature transcripts encoding the tumour suppressor

protein p53, thereby stabilising the RNA for translation and facilitating apoptosis (Kim

et al., 2013). The capacity to bind and stabilise mRNA, as well as non-coding RNA, is of

great importance in neurons, whereby HNRNPQ forms a components of the RNA

granules transported to neuronal dendrites. These structures allow the neuron to

regulate protein synthesis away from the main cell body, and depletion of HNRNPQ

alters the distribution of these granules (Bannai et al., 2004, Chen et al., 2012). For

example, HNRNPQ binds to the mRNA of components making up the Cdc42/N-

WASP/Arp2/3 (cell division control protein 42/ neuronal Wiskott-Aldrich syndrome

protein/ actin-related protein 2/3) complex which governs actin polymerisation, and in

turn migration at the leading edge of neuronal cells. Loss of HNRNPQ alters the

transport of these mRNA transcripts, leading to dispersal throughout the cell body.

Consequently neurons exhibit an altered morphology with an increased degree of

neurite complexity (Chen et al., 2012).

HNRNPQ has also been shown to bind non-coding RNA, such as brain cytoplasmic RNA

200nt (BC200 RNA), a neuron-specific component of mRNA transport complexes

believed to regulate protein synthesis at the dendrites, and thus maintaining synaptic

plasticity (Duning et al., 2008). This short nucleic acid declines with aging, and is

significantly increased and mislocalised in the brains of AD sufferers. These

observations have led some to theorise an ineffective compensatory mechanism

employed by neurons affected by AD, particularly those of the cortex, to stave off

33

synaptic degeneration (Mus et al., 2007). HNRNPQ is postulated to recruit BC200 into

the mRNA complexes, facilitating its transport to the dendrites, and its subsequent

inhibition of translational machinery (Duning et al., 2008). This inhibition is reversed

by poly(A)-binding protein (PABP), which shares the binding site with HNRNPQ,

illustrating potential competition and further translational regulation at the dendrites

(Kondrashov et al., 2005).

HNRNPU (also known as HNRNPU, SAFA, U21.1, RP11-11N7.3)

HNRNPU encodes for the largest member of the HNRNP family, and is confined solely

to the nucleus. It has the capacity to interact with a range of proteins and nucleic acids

as well as the nuclear matrix, leading to its initial characterisation as the nuclear

scaffolding attachment factor A (SAFA). As a nuclear matrix factor, HNRNPU has been

linked to a number of processes including transcription, X chromosome inactivation

and mRNA splicing (Roshon and Ruley, 2005, Hasegawa et al., 2010, Kukalev et al.,

2005). However, while other members of this protein family modulate the splicing of

specific genes, HNRNPU predominantly exerts an effect on splicing at a global level,

regulating the maturation of the U2 snRNP. It is proposed that HNRNPU may govern

the intranuclear transport of this spliceosome component, either by re-directing or

sequestering the protein (Xiao et al., 2012). Depletion of HNRNPU results in an

increase in the pool of U2 snRNP available for incorporation into the spliceosome, as

well as increased levels of both exon inclusion and exon skipping events. These

changes in splicing are thought to be dependent on the relative strength of both the 5’

and 3’ splice sites. During exon definition, the 5’ splice site is believed to influence the

recognition of the upstream 3’ splice site and U2 snRNP binding. Splice sites of equal

strength thereby facilitates exon inclusion, an effect enhanced by increased levels of

U2 snRNP. Conversely a weak 5’ splice site can diminish the recognition of its

corresponding 3’ splice site, leading to exon skipping. This effect is enhanced upon

depletion of HNRNPU and the resulting increase in U2 snRNP. The role of HNRNPU is

therefore crucial in maintaining both the levels of this spliceosome component and

the delicate balance of splicing events that it regulates (Xiao et al., 2012).

34

As a splicing factor, hnRNPU has very few recognised target genes, one of which is

caspase-9. This cysteine protease is a precursor of apoptosis, and is activated by the

apoptosome. Once activated, it triggers a cascade of downstream caspases in the cell

death pathway. Caspase-9 has two splice variants differing at exons 3-6, the inclusion

of these 4 exons leads to the formation of pro-apoptotic caspase-9a isoform, while

exclusion of this cassette generates the anti-apoptotic caspase-9b. HNRNPU binds to

exon 3 of immature mRNA transcripts and stimulate exon inclusion, while binding of

the phosphorylated form of its sister protein, HNRNPL, represses such splicing event

(Vu et al., 2013).

SFRS3 (also known as SRP20, X16)

SFRS3 encodes for the splicing factor SFRS3, a member of the SR family. Like other SR

proteins, SFRS3 contains an RRM domain to facilitate the binding of RNA. This protein

also contains an RS domain, allowing for interaction with other proteins, a crucial

function that defines splice sites for spliceosome binding (Huang and Steitz, 2005).

SFRS3 has been implicated in several cellular processes, including the β-catenin

pathway regulating stem cell regeneration, cellular senescence as well as wound

healing. (Tang et al., 2013, Goncalves et al., 2008, Li-Korotky et al., 2006).

Furthermore, SFRS3 is crucial for the formation of blastocysts during embryonic

development, with mouse embryos mice lacking the gene failing to progress pass the

early morula stage (Jumaa et al., 1999).

While many SR proteins are located in the nucleus, SFRS3 has the capacity to shuttle

between the nucleus and the cytoplasm. SFRS3 may only traverse the nuclear

membrane when attached to intronless mRNA, and mediates this export via its RS

domain (Caceres et al., 1998). Although the role of SFRS3 in the cytoplasm is largely

unknown, given that this splicing factor has been shown to interact with ribosome 80s,

some have speculated a role in coupling splicing to translation (Sanford et al., 2004).

Recently SFRS3 has been linked to AD, regulating the splicing of tropomyosin receptor

kinase B (TrkB), a receptor protein found in the BDNF signalling pathway (Wong et al.,

2012). TrkB exists as three distinct splice variants, the full length protein encoded by

the transcript denoted TrkB-TK+, and two truncated protein forms encoded by

35

transcripts defined as TrkB-TK and TrkB-Shc (Stoilov et al., 2002). The truncated forms

possess either exons 16 or 19 respectively, both of which contain translational stop

signals. While all three proteins retain the capacity to bind BDNF, only the full length

receptor elicits the downstream signalling cascade that promotes neuronal survival

and plasticity. The ratio of the three transcripts is skewed in favour of TrkB-Shc in the

hippocampus of AD sufferers, contributing to the neuronal degeneration. The

increased formation of this variant has been linked to the elevated levels of SFRS3

observed in AD, which stimulates the inclusion of exon 19. Knockdown of this splicing

factor was confirmed to significantly reduce the generation of TrkB-Shc and in turn

rescue disease phenotypes. This demonstrates a role played by splicing in

pathogenesis of AD, and potentially other neurodegenerative diseases (Wong et al.,

2012).

SNRPB (also known as COD, SNRPB1, SMB, SM11)

SNRPB (small nuclear ribonucleoprotein polypeptide B) encodes for SMB, a member of

the Sm family of proteins. This group of proteins are defined by a Sm domain that

facilitates interactions with other members of this family. During formation of the

mature snRNP spliceosomal components, seven Sm proteins, B, D1, D2, D3, E, F and G,

come together to form the snRNP subcore particle, subsequently binding either U1,

U2, U4/U6 or U5 snRNA to form the corresponding snRNPs (Urlaub et al., 2001).

Unlike other higher organisms, which solely express SMB, humans possess two splice

variants of this protein, the second of which being termed SMB’. These two variants

differ only by an additional 9 amino acids found at the carboxy-terminal end of the

SMB’ isoform, which appears to be functionally redundant (Vandam et al., 1989). The

SMB proteins are expressed in all cell types except neuronal cells, where they are

replaced by the orthologous SMN protein, encoded by the paternally imprinted

SNRPN gene. In conditions such as Prader-Willi syndrome, loss of SMN is met by an

upregulation of both SmB and SmB’, thereby compensating for the deficit and

alleviating the phenotype synonymous with this condition (Gray et al., 1999).

36

TARBBP (also known as ALS10, TDP-43)

TARDBP (transactive response region DNA binding protein) encodes for TARDBP, a

member of the HNRNP family of protein found in both the nucleus and the cytoplasm.

This splicing factor binds to more than 6000 RNA transcripts, many of which encodes

for proteins involved in neuronal functioning, including SMN2, myocyte enhancer

factor 2D (MEF2D) and MAP-kinase activating death domain (MADD) (associated with

SMA, PD and AD respectively) (Lee et al., 2012, Tollervey et al., 2011). Knockdown of

TARDBP in the adult mouse brain alters the splicing of more than 1000 mRNA

transcripts, especially those involved in synaptic plasticity and/or involving containing

long introns (Polymenidou et al., 2011). Knockout of the TARDBP gene also causes

early embryonic lethality, illustrating a pivotal role in development (Sephton et al.,

2010).

Following analysis of TARDBP-binding proteins, it became apparent that this protein

possesses a multitude of functions throughout the cell, dependent on cell type and

temporality. These roles including regulation of transcription and translation, in

addition to splicing, and TARDBP may serve to link these processes (Freibaum et al.,

2010, Sephton et al., 2012).

Mutations within the TARDBP gene have been linked to several neurological disorders

that have been grouped under the umbrella term, TDP-43 proteinopathies. In

conditions such as amyotrophic lateral sclerosis (ALS) and frontotemporal lobar

degeneration (FTLD), TARDBP undergoes aberrant post-translational modifications,

including phosphorylation and ubiquitinylation, which subsequently results in its

mislocalisation to the cytoplasm, loss of nuclear function and the formation of

TARDBP aggregates and inclusions (Lee et al., 2012). In common with other

proteinopathies, these inclusions disrupt a myriad of cellular processes and are

compounded further by the autoregulatory properties displayed by this protein. When

TARDBP levels rise above a homeostatic threshold, increased binding to the 3’-UTR of

its own mature transcript reduces mRNA stability and brings about destruction via

nonsense mediated decay (NMD). When mutated forms of the TARDBP aggregate and

37

form inclusion bodies, this autoregulatory mechanism stimulates the increased

synthesis TARDBP, exacerbating its aggregation (Buratti and Baralle, 2011).

Several groups have demonstrated a role of TARDBP in the pathogenesis of other

neurodegenerative diseases, including HD. In neuronal cell models and human brain

tissue TARDBP co-localises with mutant HTT inclusions in the cytoplasm, and this

interaction is facilitated by the glutamine/asparagine (Q/N) rich region located the C-

terminus of TARDBP (Fuentealba et al., 2010). Sequestration away from the nucleus

results in a loss of TARDBP-mediated splicing, which may exacerbate the HD

phenotype (Fuentealba et al., 2010, Schwab et al., 2008). Although overexpression of

TARDBP alone was found to convey some cytotoxicity, due to its own inherent

propensity to aggregate - its overexpression can suppress mutant HTT toxicity by

restoring its nuclear function (Fuentealba et al., 2010). Conversely, TARDBP loss-of-

function mutations in C.elegans and RNAi knockdown in striatal cell lines suppress

mutant HTT toxicity, and delay neuronal degeneration in the former. Although the

mechanism by which a reduction in TARDBP alleviated the HD phenotype in these

models was unclear, suppression was dependent upon the TARDBP-interacting

protein, survival factor progranulin (PGRN) (Tauffenberger et al., 2013).

38

2.12 Aims

Here I sought to validate the capacity of splicing-related gene candidates to

ameliorate mutant HTT-dependent phenotypes in a cell lines derived from a

pheochromocytoma of rat adrenal medulla (PC12 cell models) and human embryonic

kidney (HEK293T) cells. The former cell line expresses an inducible truncated mutant

HTT (exon 1 of the HTT containing 103 CAG repeats) construct tagged with enhanced

green fluorescent protein (EGFP) (see Figure 2.2). Following induction of mutant HTT,

these cells recapitulate many of the early HD events, including transcriptional

abnormalities and the formation of inclusions (Apostol et al., 2006, Apostol et al.,

2003). Using this cell line, we aimed to generate stable cell lines overexpressing our

splicing gene hits, and examine apoptosis, (via a luminescence based caspase

activation assay) as well as study aggregation dynamics (via an automated cell analysis

software system). Given that each of the gene hits has a role in mRNA splicing, we also

aimed to examine this post transcriptional process with the use of a dual reporter

splicing efficiency assay in HEK293T cells.

Figure 2.2 Inducible HTT construct found in the PC-12 cell line. The PC-12 cell line used

to validate the splicing gene hits, express an inducible exon 1 HTT fragment, containing

103Q polyQ region conjugated to a green fluorescent protein tag (EGFP). Induction of

this construct is achieved through the addition of ponasterone. (Adapted from Apostol,

Illes et al., 2006).

39

2.2 Materials and methods

2.21 Materials

2.211 Bacterial strains and cell lines

Escherichia coli strain

The DH5α bacterial strain was used during the course of the study, with the genotype:

F– Φ80lacZΔM15 Δ(lacZYA-argF) U169 recA1 endA1 hsdR17 (rK–, mK+) phoA supE44

λ– thi-1 gyrA96 relA1.

PC12 (adrenal gland pheochromocytoma) cell line

The Htt14A2.5 cell line was generated as described in Apostol at el (2003), expressing

an inducible truncated IT15 construct consisting of amino acids 1 to 17, followed by a

103Q poly-glutamine tract, conjugated to an EGFP encoding gene (all of which cloned

into pIND vector) (Apostol et al., 2003).The cells were maintained in 75 cm2 culture

vessels (Cellstar, Greiner) at 37 oC, 5% CO2, in a high glucose (5%) Dulbecco’s modified

eagle media (DMEM) (Life Technologies) supplemented with 10% heat inactivated

fetal bovine serum (FBS) (Life Technologies), 1% penicillin/streptomycin (PAA

LABORATORIES LTD) and 1% L-glutamate (PAA Laboratories Ltd).

Upon reaching 80% confluency, the media was aspirated off the cells and replaced

with 5 mL 2 mM EDTA (pH 8.0) (Gibco) in 1X Phosphate buffered saline (PBS) (PAA

Laboratories Ltd). To facilitate detachment, cells were incubated at room temperature

for several minutes and agitated. 5 mL of media was then added to flasks and 1 ml of

harvested cells transferred to a fresh culture flask, containing 10 ml of media.

HEK293T (human embryonic kidney) cell line

The tetracycline resistant strain of a cell line, first generated by Graham et al (1977), is

derived from primary HEK cells, putatively immature neurons (Graham et al., 1977,

Shaw et al., 2002). Cells were maintained at 37 oC, 5% CO2 in high glucose DMEM,

containing 10% FBS and 1% penicillin/streptomycin and were generously donated by

Dr Miguel Martins, MRC toxicology, Leicester, UK.

40

2.212 Constructs

pIREShyg3 overexpression construct

Genes hit were amplified from a murine cDNA library (pDEST), and cloned into the

multiple-cloning site of the pIREShyg3 vector, fused in frame with a red fluorescent

protein (RFP)-encoding gene. Given the presence of the internal ribosome entry point

(IRES) between the cDNA-RFP open reading frame (ORF) and the hygromycin

resistance gene ORF, the expression of both ORFs were conjugated.

pRRL-cPPT-CMV-X2-PRE-SIN

The pLENTI constructs expressed 25Q, 47Q, 72Q or 97Q exon 1 HTT fragment

conjugated to GFP and were cloned into the vector plasmid with the use of the SmaI

site. The constructs were generated by Chrstine Cheah and Flaviano Giorgini (Kwan et

al., 2012)

TN23 splicing efficiency construct

The splicing efficiency construct encodes β-galactosidase and luciferase reporter genes

which are separated by an intron flanked by recombinant fragments encoding

adenovirus and skeletal muscle isoforms of human tropomyosin. The introns contain

three in-frame translational stop signals, which prevent luciferase expression when

the construct is inadequately spliced. Splicing efficiency can be determined by

analysing the activity of luciferase in relation to the constitutively expressed β -

galactosidase activity (Nasim and Eperon, 2006). The construct was kindly donated by

Professor Ian Eperon, Biochemistry Department, University of Leicester, UK.

pCDNA 3.1 RFP construct

An RFP construct was used during transfection as a control. The construct was

generated by Dr Mariaelena Repici, and kindly donated to this study

41

2.213 Media and agar

Luria-Bertani agar (LB)

E.coli was grown on selective agar plates, consisting of 1% (w/v) tryptone (Oxoid),0.5%

(w/v) yeast extract (Oxoid), 0.5% sodium chloride (Fisher) and 1% agar (Bioline). The

final pH was adjusted to 7.2 and the agar sterilised via autoclaving. Prior to use, and

while in the molten state, the agar was supplemented with ampicillin (Melford), with a

final concentration of 10 μg/mL.

Luria-Bertani media (LB)

The bacterial strain was also grown in liquid media, composed of the same

constituents as the solid media, though lacking agar.

42

2.22 Methods

2.222 Generation of overexpression cell lines

Polymerase chain reaction (PCR)

Gene hits were amplified from a murine cDNA library (generated by Dr Robert Mason),

with primers containing restriction sites allowing cloning into the pIREShyg3 vector

(see Table 2.1) (TARDBP had previously been cloned by Dr Robert Mason). The

amplification, as originally developed in (Mullis and Faloona, 1987), was performed in

a thermocycler (Applied Biosystems 2720 thermocycler and G-storm dual 48 block

thermocycler) with the use of Phusion High Fidelity polymerase (Thermo Scientific)

according to manufacturer guidelines. A typical 50 μL PCR reaction comprised 100 ng

template DNA template; 2.5 μL of forward and reverse primers (Primers were

generated by Sigma-Aldrich, and diluted down to 10 μM stock); 1 μL of dNTP mix (100

mM (Promega)), 10 μL of 5x HF Phusion reaction buffer, 0.5 μL of Phusion enzyme (1

unit) and 32.5 μL of sterile distilled water.

The reaction consisted of an initial denaturation step of 30 seconds at 98 oC, followed

by 35 cycles consisting of 10 seconds at 98 oC, 30 seconds at the temperature stated in

Table 2.1, and a short period (governed by the length of gene, i.e. 30 seconds for every

500 nucleotides) at 72 oC. After cycling, amplification ended with a final 7 minute hold

at 72 oC.

Gel electrophoresis

PCR amplicons were visualised with the use of gel electrophoresis. Given the small

sizes of the fragments, a 0.8% agarose gel was used for the separation. Gels were

made by dissolving agarose (Seakem, Lonza) in 1x Tris-borate-EDTA (TBE) (10x stock

consisted of 108 g Tris, 55 g Boric acid in 900 mL of water, and 40 mL of 0.5 M

Na2EDTA pH8.0) and supplemented with ethidium bromide while in a molten state.

Upon solidification, gels were placed in gel tanks and immersed in 1x TBE. 5x loading

dye (Bioline) was added to samples, and subsequently loaded onto the gel. 5 μL of the

molecular weight marker, Hyperladder I (Bioline) run alongside samples. An electrical

current was applied to the gel to facilitate separation of fragments.

43

PCR clean up

Primers, salts and other PCR reaction components were removed from reactions using

the E.Z.N.A cycle pure kit (Omega Biotek) according to manufacturer guidelines.

Restriction digest

Following validation with the use of gel electrophoresis and the removal of PCR

reaction components, fragments were digested overnight at 37 oC with the

appropriate restriction enzymes (New England Biolabs Ltd. (NEB)) (see Table 2.1). The

42 μL digest reactions typically contained 30 μL of plasmid (approximately 150 ng/μL);

4 μL of the appropriate 10x NEB reaction buffer; 4 μL of 10x NEB bovine serum

albumin (BSA); 2 μL of each of the endonuclease (20 units).

Gene Hit Forward primer (5’-3’)

Reverse primer (5’-3’)

Tm (oC)

Restriction sites

Amplicon size (bp)

CRNKL GATCTGTACAGTCATGGCAGCCTCCACG

AGCTGCTAGCGCAAAAGATGAGGATTCACTCTCATC

54 BsrgI and NheI

2073

EFTUD2 GATCGCTAGCAACATGGATACTGACTTGTATGATG

AGCTCCCGGGGCCATGGGATAATTGAGCACAACATCC

49 NheI and XmaI

2919

HNRNPF GATCTGTACAGATATGATGCTGGGCCCTGAG

AGCTCCCGGGGCATCATATCCGCCCATGCTGT

56 BsgI and XmaI

1248

HNRNPK GATCTGTACAGATATGGAGACCGAACAGCCAGA

AGCTGCTAGCGCGAAAAACTTTCCAGAATACTGCTTCAC

51 BsrgI and NheI

1392

HNRNPQ GATCGCTAGCAACATGGCTACAGAACATGTTAATGG

AGCTCCCGGGGCCTTCCACTGTTGCCCAAAAG

54 NheI and XmaI

1872

HNRNPU GATCTGTACAGATATGAGTTCTTCGCCTGTTA

AGCTGCTAGCGCATAATATCCTTGGTGATAATGCTGAC

50 BsrgI and NheI

2403

SFRS3 GATCTGTACAGATATGCATCGTGATTCCTGTCC

AGCTGCTAGCGCTTTCCTTTCATTTGACCTAGATCGG

54 BsrgI and NheI

495

44

E.coli Transformation

To generate a sufficient quantity of pIREShyg3 for subsequent digestion, the vector

plasmid constructs were initially transformed into chemically competent E.coli for

subsequent plasmid prepartions.

E.coli were made competant using the method described in (Mandel and Higa,

1970).Following growth at 37 oC with agitation, the density of cells was determined

with the use of a spectrophotometer (Spectronic Unicam) at the 600 nm (OD600). Upon

reaching a density of approximately 0.5, cells were centrifuged at 3000 rpm for 10

minute at 4 oC. Cell pellets were washed in ice cold distilled water, suspended in ice

cold 0.1 M CaCl2 (Sigma-Aldrich), and incubated on ice for 90 minutes. Cells were then

centrifuged for a further 5 minutes at 3000 rpm, and re-suspended in 0.1 M CaCl2 15%

glycerol (Fisher Scientific). The suspended cells were stored at -80 oc for further use.

Prior to transformation, competent cells were thawed on ice. 50 ng of plasmid DNA

was added to cells, which were then incubated on ice for 20 minutes. Cells underwent

heat shock at 42 oc for a further 2 minutes, and were subsequently stored on ice for 2

minutes. LB media was added to bacterial samples, which were then incubated at 37

oc to facilitate phenotypic recovery. Finally cells were pelleted at 13000 rpm for 1

minute, re-suspended in sterile distilled water and plated on selective LB plates.

Plasmid extraction

Single colonies were inoculated into 5 mL of liquid LB and grown 37 oC, with agitation

at 220 rpm, overnight. Plasmids were extracted and purified using the E.Z.N.A Plasmid

Miniprep Kit 1 (Omega Biotek) following the instructions supplied by the

manufacturers, and based on the original protocol described in (Birnboim and Doly,

1979).

SNRPB GATCTGTACAGATATGACGGTGGGCAAGAGC

AGCTGCTAGCGCAAGCAGACCTCGCATGCC

57 BsrgI and NheI

696

Table 2.1. Primers for amplification of candidate genetic modifiers. The putative

suppressors were amplified from a mouse cDNA library, using primers containing

restriction sites to allow insertion into the multiple cloning site of the pIREShyg3 vector.

45

Once extracted, the pIREShyg3 plasmid was digested using the same restriction

enzymes as those used to digest the amplified splicing genes (see Table 2.1),

generating complementary sticky ends to allow ligation of the digested PCR products.

Gel extraction

Once digested, the pIREShyg3 plasmid was run on a 0.8% agarose gel. Bands

corresponding to the linearized plasmid were excised with the aid of a transluminator

(Chromato-Vue). Plasmid DNA was then extracted with the use of the E.Z.N.A gel

extraction kit (Omega Biotek) as stated in the manufacturer’s instructions.

Dephosphorylation

To prevent re-circularisation, the 5’ ends of the linearized plasmid were

dephosphorylated. 1μg of the digested plasmid (gauged relative to the known

amounts of DNA in each band of the molecular marker) was treated with Antarctic

phosphatase (NEB) according to manufacturer’s guidelines. Samples were incubated

at 37 oC for 1 hour, followed by a further incubation of 15 minutes at 65 oC. In addition

to the digested plasmid, a typical 20 μL reaction mixture contained 2 μL of 10x

Antarctic phosphatase buffer; 1 μL of Antarctic phosphatase (5 units), with the

remaining volume being made up of sterile distilled water.

Ligation

To generate the overexpression constructs, digested PCR products were ligated with

the linearized vector, at 3:1 molar ratio, as previously demonstrated in (Dugaiczyk et

al., 1975). Ligation was performed using use of T4 DNA ligase (NEB), consistent with

the manufacturer instructions and samples were incubated at 16 oC overnight. A

typical 20 μL reaction mixture consisted of 1 μL of dephosphorylated vector (50 ng,

0.025 pmol); insert fragment (0.075 pmol); 2 μL of 10X T4 ligase reaction buffer, and 1

μL of T4 DNA ligase (400 units)

Following the overnight incubation, the ligation mixes were transformed into E. coli,

and plated on to LB plates containing ampicillin. Single colonies were picked, and

plasmids isolated via the methods stated above. The success of the ligation was

assessed by digesting the inserts out the plasmid with the aid of same restriction

46

enzymes used to generate complementary fragments. Digested samples were run on a

0.8% agarose gel, with samples derived from successful ligations displaying bands

corresponding to both the vector plasmid backbone and the splicing hit cDNA.

Sequencing

Recombinant plasmids were validated by Sanger sequencing. A typical 10 μL reaction

consisted of 300ng of purified plasmid combined with 2 μL of sequencing primers

(1μM stock, with a final concentration of 0.1 μM) (see table 2.2); 2 μL of 5X BigDye

reaction buffer and 1μl of BigDye® Terminator v3.1 ready reaction mix. Samples were

then placed into a thermal cycler (Applied Biosystems 2720 thermocycler) and were

initially denatured for 1 minute at 96 oC, followed by 25 cycles, made up of a 10

seconds at 96 oC; 5 seconds at 50 oC; and 4 minutes at 60 oC, after which samples were

held at 16oc.

To remove excess dye terminator, 10 μL of 0.4% (w/v) sodium dodecyl sulphate (SDS)

(Fisher) was added to samples, and incubated at 95 oC for 10 minutes, followed by a

further 10 minutes hold at 22 oC. Sequencing products were then purified with the aid

of the Performa Gel Filtration Cartridges (EdgeBio). Spin columns were initially

centrifuged at 2800 rpm for 3 minutes to remove storage buffer from the column

matrix. Samples were placed into the columns, which were placed into a fresh

microcentrifuge tube and centrifuged for 2800 rpm for 3 minutes. DNA sequencing

was then performed with the use of the Applied Biosystems 3730 sequencer by the

Protein Nucleic Acid Chemistry Laboratory (PNACL) services available at the University

of Leicester. Sequences were visualised with ApE- A plasmid Editor v1.12 and

compared to those obtained from the Mouse Genome Informatics (MGI) database.

47

Sequencing

Primer

Sequence (5’-3’) Tm (oC)

T7 Forward TAATACGACTCACTATAGGG 48

MRFP Reverse GCCCTCGATCTCGAACTCGTG 60

CRNKL SEQ F1 CGGTGGAGTTTTTTGGGGAT 57

CRNKL SEQ R1 CCTATGGAAGTTCCCAAAGCTC 57

EFTUD2 SEQ F1 TGAAGCTGCCTCCAACAG 56

EFTUD2 SEQ F2 CACATTGAGGTGAATCGTGT 54

HNRNPQ SEQ F1 TGGCAAAGGTAAAAGTGCTG 55

HNRNPU SEQ F1 CTTCTTCCCTTACTATGGAG 50

HNRNPU SEQ F2 AGACAAATGTGTCTGCTGCT 56

Endo-free DNA miniprep

Following examination of the sequencing data, plasmid DNA was purified with the use

of the EZNA Endo-Free Plasmid Mini Kit II. This kit allowed for the removal of

endotoxins for use in cell culture work, most notably the toxic lipopolysaccharides

(LPS) present on the cell surface of the E. coli strain. The purification was carried out

according to manufacturer guidelines.

DNA purification

Upon the removal of endotoxins, samples were further purified and concentrated by

phenol chloroform extraction. Samples were diluted to 400 μL using sterile distilled

water, transferred to screw cap 1.5 mL tubes, and mixed with an equal volume of

phenol: chloroform: Iso amylalcohol (IAA) (this solution consisted of phenol (Fisher

Scientific): cholorform (Fisher Scientific): IAA (Fisher Scientific) mixed at a 25:24:1

ratio). Samples were vortexed, centrifuged at 13,000 rpm for 10 minutes at 4 oC, and

Table 2.2. Sequencing primers for validation of mouse cDNA overexpression

constructs. Recombinant plasmids were sequenced with the use of several primer sets.

Both the T7 Forward and MRFP Reverse oligonucleotides are plasmid specific primers,

annealing to regions flanking the MCS. Several large constructs also required the use of

additional primers, which annealed to regions within the inserted DNA.

48

the aqueous phase removed to a fresh tube. Plasmid DNA was then precipitated by

the addition of 40 μL of 3 M sodium acetate (pH 5.3) (Fisher Scientific) and 800μl of

cold 100% ethanol (Fisher Scientific). Samples were incubated at -20oC for 1 hour

before being centrifuged at 4oC at 13,000 r.p.m for 15 minutes. The pelleted DNA was

washed with 1 mL of 70% ethanol, centrifuged for 5 minutes, and then washed a

second time in the same volume of 70% ethanol. Following a final centrifugation step,

the ethanol was removed and pellets dried at room temperature before being

suspended in 50 μL of cell culture grade water (PAA Laboratories Ltd).

Prior to transfection, the concentration and purity of the plasmid samples was

determined with the use of a nanophotometer (Implen). The quality of samples was

assessed at 260 nm and 280 nm wavelengths, with the ratio of the two used to

measure the purity, i.e. a 260:280 ratio of greater than 1.8 was deemed to be pure

enough for transfection.

Transfection of PC12 cells

Sub-confluent PC12 cultures were harvested and 2 x 105 PC12 cells (cell number was

determined with the use of a cell counter chamber (Hawksley)) seeded into six-well

plates (Greiner) and incubated for 24 hours. Prior to seeding, the plates were coated

with 10% poly-lysine solution (Thermoscientific) diluted in 1x PBS, incubated at 37 oc

for 20 minutes, and subsequently washed several times in 1X PBS. The following day, 4

µg of plasmid DNA was diluted into 400 μL of serum-free media, along with 6 μL of the

TurbofectTM transfection reagent (Fermentas). The mixture was then vortexed for

several seconds, incubated at room temperature for 20 minutes, and added to cells in

a drop-wise manner. After 6 hours, the media was aspirated and replaced with fresh

media.

Alternatively transfection of PC12 cells was achieved via nucleofection (AMAXA). Sub-

confluent cultures were harvested and 2x106 cells centrifuged at 700 rpm for 5

minutes. The resulting pellet was suspended in 2 μg of plasmid DNA diluted in 100 μL

of electroporation transfection solution (Mirus). Cells were transferred into a cuvette

and nucleofected in an AMAXA nucleofector II using programme U-29. Cell were

49

immediately transferred to six-well plates containing 3 mL of pre-warmed media,

which was changed after a further 24 hours.

After 4 days cells were passaged and maintained under selective antibiotics. Media

was supplemented with 200 µg/ml of G418 (PAA laboratories Ltd) and 200 µg/ml of

hyrogromycin B (Invitrogen) to select for the HTT construct and overexpressed cDNA

respectively, and changed every 72 hrs until resistant colonies were formed.

Freezing stable PC12 cell lines

Following several passage stable transformants were frozen for long term storage.

Cells were grown to 80% confluency, and detached using PBS supplemented with 2

mM EDTA. Cells were centrifuged at 700 rpm for 5 minutes, and the pellet suspended

in 2 mL of media/ 10% dimethyl sulfoxide (DMSO) (Sigma-Aldrich). Cells were

aliquoted and stored at -80 oC for long term use.

2.222 Validation of splicing gene hits

Caspase Assay

In order to determine whether overexpression of the splicing genes conveyed the

same suppression of mutant HTT as previously demonstrated in the yeast screen, a

caspase assay was performed on the stable cell lines. Sub-confluent cultures were

harvested and 1 x 104 cells in 200 μL of media seeded in white 96-well plates

(Greiner). After 24 hours, expression of mutant HTT by the addition of of 1 μL of 1 mM

ponasterone in DMSO (working concentration of 5 μM). After an additional 72 hours,

an equal volume of caspase 3/7 Glo® reagent (Promega) was added to each of the

wells, and plates incubated for approximately 2 hours at room temperature.

Luminescence was then measured with FlUOstar Omega mircoplate reader Plate

Reader (BMG LABTECH) using the MARS data analysis software (top optic, emission

filter lens, gain 3600, and positioning delay 0.2). Background luminescence from

capase 3/7 Glo reagent added to media was subtracted from the test values, which

were subsequently normalised to untransfected control cells.

50

RNA isolation

RNA was isolated from PC12 cell lines using the method originally developed by

(Chomczynski and Sacchi, 1987). PC12 cells were seeded in poly-lysine coated 6-well

plates at a seeding density of 2 x 105 cells, and allowed to grow for 72 hours in 3 mL of

growth media. After the 3 days, the media was then removed and replaced with 1 mL

of TRI reagent (Sigma-Aldrich), for 5 minutes. Cell lysates were then transferred to

screw cap 1.5 mL micro-tubes, 0.2 mL of chloroform added, and samples incubated at

room temperature for 15 minutes. Samples were centrifuged for 15 minutes at 13,000

rpm at a temperature of 4 oC, and the RNA-containing upper phase transferred to a

fresh micro-tube, mixed with 0.5 mL of isopropanol (Fisher Scientific) and incubated at

room temperature for 10 minutes. Samples were centrifuged at 13, 000 rpm for 10

minutes, and the pelleted RNA pellet washed twice in 1 mL of 70% ethanol and

centrifuged at 7,500 rpm for 5 minutes. Samples were air dried and suspended in 50

μL RNAase/DNAase-free distilled water (Life Technologies). The concentration and

purity of samples were determined using a nanophotometer.

Reverse Transcription and Quantitative Real-Time PCR (QPCR)

250 ng of isolated RNA was reverse transcribed to form single stranded cDNA, using

the instructions supplied by manufacturers of the ImProm-II™ Reverse Transcription

System kit (Promega). Initial 5 µL reaction mixtures consisting of 250 ng of RNA, 1μL of

Oligo (dT)15 and sterile distilled water were incubated at 70 oC for 10 minutes. Samples

were then supplemented with 4 μL of 25 mM MgCl2, 2 µL Reverse transcription 10X

buffer, 2 μL of 10 mM dNTP mixture, 0.5 µL recombinant RNasin ribonuclease inhibitor

and 1 µL of AMV reverse transcriptase (15 units). The PCR reaction was performed in a

thermocycler (Applied Biosystems 2720 thermocycler), and consisted of three

temperature steps, an initial 5 minute hold at 25 oC to facilitate annealing, a one hour

extension at 42 oC and finally a 15 minute hold at 72 oC to inactivate the reverse

transcriptase enzyme.

Primer sets for qPCR (see Table 2.3) were tested for their specificity and their capacity

to dimerise. 1 μL of serially diluted cDNA samples (12.5 ng/μL, 1.25 ng/μL, 0.125 ng/μL

and 0.0125 ng/μL) was placed into each well of a 96 well real time PCR plate (three

51

experimental repeats per sample). 19 μL of master mix, consisting of 10 μL of Fast

SYBR® Green Master Mix (Applied Biosystems), 7.4 μL of sterile DNAase/RNAase-free

distilled water, 0.8 μL of 10 μM forward and reverse primers, was then added. Plates

were sealed, centrifuged briefly, and analysed with the use of the Applied Biosystems

StepOne™plus Real-Time PCR System and software. Amplification was achieved via an

initial denaturation step at 95oC for 10 minutes, followed by 40 cycles of 3 seconds at

95 oC and 30 seconds at 60 oC, with fluorescence monitored at the elongation stage of

each cycle.

The specificity of the reaction was determined by melt curve analysis. By analysing the

inverse change in fluorescence against the change in temperature, the reaction

products of the amplification can be deduced. The reaction efficiency was examined

by plotting the logarithmic quantity of starting template against the threshold cycle

(Ct). A correlation coefficient of greater than 0.98 was indicative of an efficient

amplification reaction.

Following validation of primer sets, overexpression of candidate modifiers was

determined using the conditions stated above. Amplification was carried out using 10

µM MRFP forward and reverse primers (see Table 3), and data normalised to a

reference gene, PpiB (peptidylprolyl isomerase B) (10 µM forward and reverse

primers).Subsequent quantification of results was performed using the δδCT method

as described in (Pfaffl, 2001)

Primer Sequence (5’-3’) Tm (oc)

mRFP QPCR Forward CGCCTACAAGACCGACATCA 58

mRFP QPCR Reverse CGCTCGTACTGTTCCACGAT 58

Ppib Forward GGCTCCGTTGTCTTCCTTTT 57

Ppib Reverse ACTCGTCCTACAGGTTCGTCTC 59

Table 2.3. PCR primers to assess overexpression of candidate modifiers.

Overexpression of splicing genes was determined via real-time PCR with the use of

primers designed against conjugated mRFP gene. Results were normalised to the

reference gene, PpiB (peptidylprolyl isomerase B).

52

Confocal Microscopy

Confocal microscopy was employed to evaluate the cellular location of RFP tagged

suppressors in PC12 cells. 2 x 104 PC12 cells were seeded onto round poly-L-lysine

coated coverslips (12 mm diameter), in 24-well plates and incubated for 72-hours.

Cells were then fixed with 4% paraformaldehyde (Fison) for 20 minutes at room

temperature and later washed 3 times in 1X PBS containing 0.1% (v/v) Tween 20

(Sigma-Aldrich) (PBST). Cells were then permeabilised in 1X PBS containing 0.5% (v/v)

Triton X-100 (Fisher Scientific) for 10 minutes at room temperature, and washed a

further 3 times in PBST. The fixed PC12 cells were incubated in 4% (w/v) BSA for 10

minutes at room temperature and then for a further hour in 4% BSA supplemented

with 1:200 RFP booster (Chromotek). Cells were washed three times in PBST, each for

a 10 minute period, before being mounted onto slides with the aid of a DAPI (4’, 6-

diamidino-2-phenylindole) containing mounting media (VECTASHIELD®).

Cells were visualised using an Olympus FV1000 confocal microscopy. The DAPI and red

fluorophores were detected with the aid of the UV (405 nm) and red (635 nm) diode

lasers respectively. Images were acquired using a 60X objective lens, a numerical

aperture (NA) of 1.35 and the Z stack function to generate approximately 30 images

(at 0.38μm increments). The clarity of the images was increased with of deconvolution

software, Huygens Essentials (Huygens Software).

2.223 Characterisation of splicing gene hits

Splicing assay

HEK293T cells were seeded in white 96-well flat bottomed plates (Greiner) at a

seeding density of 2.5 x 104 cells per well in 200 μL of growth media and grown for 24

hours. RFP, HTT pLENTI, suppressor cDNA and TN23 constructs were isolated and

purified and as described above and diluted to 100 ng/μL stocks. Cells were

transfected with either 2 or 3 plasmid constructs, with a total amount of 200ng per

well. Constructs were mixed with 0.4 μL of Turbofect transfection reagent

(Fermentas), in a total of 20 μL of serum free media, which was used to transfect cells

per well, as previously described. All transfections were performed in triplicate.

53

Splicing efficiency determined 48 hours after transfection with this use of the Dual-

Light® Combined Reporter Gene Assay System for Detection of Luciferase and beta-

Galactosidase (Applied Biosystems), following the manufacturer instructions supplied

with the kit. Cells were lysed in 25 μL of lysis solution for 10 minutes at room

temperature. The 25 μL of buffer A, containing components required for the enhanced

luciferase reaction, added to each well. Plates were placed in a FLUOstar Omega

mircoplate Reader (BMG LABTECH), and wells injected (at a speed of 310 μL/s) with

100 μL of buffer B, containing luciferin, supplemented with Galacton-Plus substrate at

a 1:100 dilution. Luminescence was measured a 1 second delay (top optic, emission

filter lens, gain 3600, and positioning delay 0.2). The luciferase signal was then allowed

to deteriorate for 1 hour at room temperature prior to measuring β-galactosidase

activity. Accelerator-II solution, containing luminescence enhancer and alkali (the

higher pH is conducive to the β-galactosidase reaction), was then injected into wells

and a second reading taken. Splicing efficiency was determined by calculating the ratio

of luciferase activity to beta-galactosidase.

WST-1 assay

HEK293T cells were transiently transfected in clear 96-well F-bottom plates (Greiner)

as stated above and assessed in terms of cytotoxicity with the WST-1 assay (Roche). 10

μL of WST-1 reagent was added to cells 48 hours after transfection and microplates

incubated at 37 oC, 5% CO2 for 4 hours. Cells were shaken briefly and the absorbance

at 440 nm (path length correction 100 μL, length 3.31mm, position delay 0.2, number

of flashes per well 22) measured using a FLUOstar Omega microplate Reader (BMG

LABTECH). Background absorbance from media treated with WST-1 reagent was

subtracted and samples normalised to untransfected control.

Aggregation studies- Cellomics

PC12 cells were seeded in a poly-lysine treated 24-well plates at a seeding density of 2

x 104 cells in 0.5 mL of culture media, and induced at varying time points (0, 12, 18, 24,

48, and 72 hours) with 5 μM ponasterone. After 72 hours, media were aspirated and

cells fixed 0.5 mL of 4% paraformaldehyde for 20 minutes at room temperature. After

washing in 1X PBS, 0.5 mL of 1 μg/mL Hoechst 33342 (Invitrogen) diluted in 1x PBS

54

was added and samples incubated at room temperature for 15 minutes. Cells were

washed a further 3 times in 1x PBS and stored in 1X PBS at 4 oC.

Fixed cells were analysed using an ArrayScan* Infinity High Content Reader (Thermo

scientific), with Cellomics® one software (Thermo scientific). The Target activation

bioapplication was used to measure fluorescence, fluorescent intensity and variation

in fluorescence intensity (for ~2000 cells). Cells displaying an average intensity greater

than 20 were scored as expressing HTT, while cells exhibiting a variation in intensity of

greater than 200 were scored as containing HTT inclusion bodies or aggregates. Cells

with variation in intensity of less than 200 displayed diffuse expression of HTT.

Dot blots/ Filter trap assay

Mutant HTT protein levels in PC12 cells were assessed using filter retardation assays

(Wanker et al., 1998). Cells were seeded in poly-lysine treated 6-well plates at a

density of 2.5 x 105 in 500 μL cell culture media and induced with 5 μM ponasterone A

after 24 hours. After a further 48 hours, media was aspirated and replaced with 300 μL

of cell lysis-M-regent (Roche) supplemented with 1x EDTA-free protease inhibitor

cocktail (Roche) in 1X PBS. Cells were agitated and lysed for 15 minutes at room

temperature. Following lysis, 0.3 μL (75 units) of Pierce universal nuclease

(Thermoscientific) was added and samples and incubated on ice for 30 minutes to

remove genomic DNA. Protein samples were subsequently quantitated using a

nanophotometer, diluted in 1X PBS containing 2% SDS and 50 mM dithiothreitol (DTT)

(Melford) and boiled at 98 oc for 3 minutes (see Table 4).

Prior to blotting, membranes and 2 pieces of 3MM blotting paper were equilibrated in

0.1% SDS in 1X PBS, and assembled onto the 96-well dot blot apparatus (Geneflow),

which was then connected to the vacuum manifold. Nitrocellulose (0.2 μm,

Whatman), was used to assess total mutant HTT levels, while cellulose acetate

membrane (0.45 μm, Whatman), was used to detect mutant HTT aggregates.

Denatured proteins samples were loaded onto the dot blotter in duplicate (see Table

2.4), and wells washed 4 times with 100 μL of 0.1% SDS. Membrane were air-dried and

blocked with of 3% (w/v) milk (Sigma) in Tris buffered saline (TBS)/Tween (0.605%

(w/v) Tris (Sigma), 0.876% (w/v) sodium chloride (Fisher), pH7.5 containing 0.05% (v/v)

55

Tween 20 overnight at 4 oC with continued agitation. Nitrocellulose and cellulose

acetate membranes were probed with 1:10,000 rabbit anti GFP (Abcam 6556)

antibody in 3% milk/TBST for one hour at room temperature with constant rocking.

The primary antibody was subsequently removed, membranes were washed 3 times

in TBST, and probed with horseradish peroxidase (HRP) conjugated goat anti-mouse

secondary antibody (PI-2000, Vector Laboratories) for a further hour at room

temperature. Finally, membranes were washed 5 times with excess TBST, and covered

with SuperSignal West Dura Chemiluminescent Substrate (Thermoscientific) for 5

minutes at room temperature. Membranes were exposed to radiographic film (Fuji)

for varying lengths of time (typically 30 seconds for the nitrocellulose and

approximately 10 minutes for the cellulose acetate). Film was developed in developer

agent (RG Universal) until dots became visible, immersed in stop solution (0.3% acetic

acid, Fisher Scientific), and fixed in fixation fluid (RG Universal).

Image J software (http://imagej.nih.gov/ij/) was utilised to measure the average pixel

density of each sample, which were normalised to the amount of protein loaded and

the untransfected control samples.

Membrane type Concentration

(μg/ml)

Volume (μL) Amount of

protein (μg)

Nitrocellulose 133 150 20

Nitrocellulose 133 112.5 15

Nitrocellulose 133 75 10

Nitrocellulose 33.3 150 5

Nitrocellulose 33.3 75 2.5

Nitrocellulose 33.3 30 1

Cellulose acetate 667 150 100

Cellulose acetate 667 75 50

Cellulose acetate 667 37.5 25

Cellulose acetate 66.7 150 10

Cellulose acetate 66.7 75 5

Cellulose acetate 66.7 37.5 2.5

56

HTT expression levels via qPCR

PC12 cells were seeded into 6 well plates as stated above, induced after 24 hours and

allowed to grow for a further 48 hours. RNA isolation, primer set validation and

quantitative real-time PCR was performed as previously described, using primers

designed against EGFP tag conjugated to the inducible mutant HTT expressed by this

cell line (see Table 2.5), coupled with the PPIB primers used previously (see Table 2.3)

(Pfaffl, 2001). Results were normalised to both the expression levels of PPIB, and the

untransfected control cells.

Genomic DNA isolation

Quantitative PCR method was used to assess the copy number of the mutant HTT

construct within each of the stably transfected PC12 cell lines. Upon reaching 80%

confluency, cells were lifted from T25 flasks and centrifuged at 700 rpm for 5 minutes,

and the resulting pellet incubated at 56 oc overnight, in 500 µL of lysis buffer,

consisting of: 100 mM NaCl; 10 mM Tris pH 8.0; 25 mM pH 8.0; 0.5% (w/v)SDS and 0.1

mg/mL of proteinase K (New England Biolabs Inc.) (added prior to use). Following lysis,

55 µL of 3 M sodium acetate and 500 µL of phenol: chloroform: IAA was added, and

samples inverted several times. Samples were centrifuged at 13,000 rpm for 10

minutes, and the resulting aqueous layer transferred to a fresh micro-tube, mixed

with 500µl of chloroform and subsequently centrifuged for a further 10 minutes at

13,000 rpm. The aqueous layer was transferred to a fresh 1.5 mL micro-tube, mixed

with 450 µL of isopropanol, and incubated overnight at -20 oC. Precipitated DNA was

then pelleted by centrifugation at 13,000 rpm for 30 minutes at 4 oC, and washed

twice with 70% ethanol. Samples were air dried and re-suspended in 50 μL

RNAase/DNAase-free distilled water (Life Technologies).

Table 2.4. Protein sample dilutions and loading volumes. Protein samples were

diluted down to four concentrations (column 2); two stock solutions were prepared for

both nitrocellulose and cellulose acetate (column 1). Varying volumes of these samples

were loaded onto each of the membranes (column 3) containing varied amounts of

protein (column 4)

57

To remove any contaminating RNA, samples were incubated at 37 oC with 1 µL of 10

µg/mL of RNase A (Sigma) for 30 minutes. DNA was precipitated by adding 5 µL of 3 M

sodium acetate and 100 µL of 100% ethanol, and tubes incubated on ice for 10

minutes. Samples were then centrifuged at 13,000 rpm for 5 minutes and washed

twice with 70% ethanol, air dried, and suspended in 50 μL sterile 1x TE buffer, made

up of: 10 mM Tris pH 8.0; 1 mM EDTA. The concentration and purity of samples were

determined using a nanophotometer.

Validation of HTT copy number

The number of copies of the HTT constructs in PC12 cells was verified by quantitative

PCR, as previously described. Primers that anneal within the EGFP gene were used,

coupled with primers designed to anneal to intronic and exonic regions of the PPIB

gene (see Table 2.5). Results were normalised to both the copy number of PPIB gene,

and then untransfected control cells.

Primer Sequence (5’-3’) Tm (oc)

EGFP QPCR Forward GAAGTCGTGCTGCTTCATGTG 58

EGFP QPCR Reverse TCGTGACCACCCTGACCTAC 59

Ppib intronic forward CCAGACTCAGCCACAGCATT 60

Ppib exonic Reverse ACTCGTCCTACAGGTTCGTCTC 59

2.23 Statistics

Each of the data sets were analysed with GraphPad Prism 15 software, to determine

statistical significance. The choice of statistical test was governed by both normality,

homogeneity of variances and pairing. Both the D'Agostino & Pearson omnibus and

Shapiro-Wilk normality tests were used to determine the distribution of the data, and

Bartlett’s test was used to gauge the homogeneity of the data. Data was transformed

Table 2.5. Primers to assess HTT construct expression and copy number. Expression

of mutant HTT, following induction with ponasterone, was determined by amplifying a

small region of the conjugated EGFP tag. The same primers were used to determine

copy number of the HTT construct, and were used in conjunction with primers

designed to anneal to the intronic and exonic regions of the PPIB gene.

58

logarithmically when failing to conform to the necessary prerequisites of parametric

analysis.

Following these tests, normally distributed and homogeneous data was analysed via a

two-sample Student t-test (parametric) or a one way ANOVA, while data failing to

display homogeneity and normal distribution after logarithmic transformation was

analysed via the Kruskal-Wallis test and subsequently a pair-wise Mann Whitney test.

In each statistical test a confidence level of 0.05 was used.

59

2.3 Results

2.31 Overexpression of several candidate genetic modifiers suppresses mutant HTT

toxicity in mammalian cells.

The yeast modifier screen performed by our research group identified a number of

genes involved in mRNA splicing, which when overexpressed, suppressed mutant HTT

toxicity. To further validate these hits in a mammalian context, we initially cloned the

nine genes into the pIREShyg3 construct. Genes were inserted into this vector in-

frame and upstream of the RFP gene, thus fluorescently tagging each gene and

facilitating localisation studies. Transcription of the cDNA-RFP ORF was linked to the

hygromycin resistance gene by means of an IRES sequence. We subsequently

generated HTT14A2.5 PC12 cell lines stably overexpressing selected modifiers. These

cells contained a stably integrated HTT103Q exon 1 construct under the control of a

tightly regulated ponasterone A inducible promoter.

After generating these stable cell lines, a caspase 3/7 activation assay was used to

assess the toxicity displayed by these cells in response to the mutant HTT induction.

Given that caspase activation is a precursor of apoptosis, this assay has previously

been used to determine changes in the apoptotic state of cells in response to varying

toxic insults and suppressors(Liu et al., 2004, Ren et al., 2004).

Overexpression of CRNKL, EFTUD2, HNRNPF, HNRNPK and TARDBP reduced levels of

toxicity in relation to untransfected cells. In contrast, cell lines overexpressing

HNRNPQ or SNRNPB demonstrated toxicity levels comparable to this control, which

thus suggests that in mammalian cells overexpression of these gene does not suppress

mutant HTT toxicity. Finally, overexpression of the two remaining splicing genes,

HNRNPU and SFRS, appeared to aggravate toxicity (see Figure 2.3)

Given these results, cell lines overexpressing CRNKL, EFTUD2, HNRNPF, HNRNPK or

TARDBP were examined further to determine the mechanism of suppression.

60

2.32 Suppressors are overexpressed and localise in the nucleus

As a means of gauging cellular location, as well as their possible co-localisation with

mutant HTT, each of the splicing proteins was tagged with mRFP. However given the

inherently weak signal produced by this fluorophore, cells were incubated with an RFP

booster. This booster consisted of a RFP binding protein conjugated to a red

fluorescent ATTO-TEC dye, emitting a stronger and more stable signal.

Supporting their known role in mRNA splicing, all of the proteins were found to reside

in the nucleus (see Figure 2.4). CRNKL and EFTUD2 both localised to distinct areas

Figure 2.3. Overexpression of several splicing genes modulates caspase activation in neuronal cells expressing mutant HTT. Selected splicing genes were stably expressed in PC12 cells containing an inducible HTT exon 1 construct (103Q) (Apostel, Kazantsez at al., 2003). To test for suppression of mutant HTT, a caspase 3/7 activation (a precursor of apoptosis) cytotoxicity assay was performed and results were normalised to the untransfected cells. Overexpression of several genes hits suppressed toxicity, while others exacerbated mutant HTT toxicity. A Kruskal-Wallis test was performed on the data as a whole, resulting in a P<0.0001 (****) (N = 14 for the controls and N = 3-5 for test samples, SEM error bars) . A Mann-Whitney test was carried out on each of the test cell lines relative to the untransfected control cells (*P<0.05 **P<0.01, ***P<0.001).

61

within the nucleus. Unlike the other suppressors, the primary role of these proteins is

the assembly of the spliceosome. As such, the observed structures, reminiscent of

splicing “speckles” described in (Han et al., 2011), may reflect the sub nuclear

locations of mRNA splicing.

HNRNPF and HNRNPK displayed a more diffuse spread throughout the nucleus, a

result consistent with roles in transcription and chromation re-modelling as well as

splicing. In contrast TARDBP localised in discrete sub-nuclear bodies, while retaining

diffuse levels in the rest of the nucleus.

Expression was confirmed via quantitative real-time PCR, with amplification using

primers detecting against the RFP tag. All cell lines displayed RFP expression, levels of

which were normalised to untransfected controls (see Figure 2.5).The cell line

expressing TARDBP displayed the highest level of expression, while cells expressing

CRNKL exhibited the lowest level of expression.

62

Figure 2.4. The RFP tagged suppressors localised in the nucleus. Cell lines were seeded onto cover slips and grown for 48 hours. Cells were then incubated with an RFP-booster to enhance the signal emitted by the conjugated mRFP. Expression of the each of the RFP tagged suppressors lead to red fluorescence within the nucleus, consistent with a role in mRNA splicing. Untransfected cells did not to display a red signal (not shown). Images were taken under a 60X magnification.

63

2.33 Mutant HTT causes a polyQ dependent decline in splicing efficiency and cell

viability

In order to examine the role of mRNA splicing in HD pathogenesis, and by extension

the possible mechanistic route by which our candidate genes conveyed suppression,

we established a splicing efficiency assay in the context of a HD cell model. The

splicing efficiency assay utilised the TN23 construct encoding β-galactosidase and

luciferase, separated by an intron flanked by recombinant fragments encoding

adenovirus (Ad) and skeletal muscle isoforms of human tropomyosin (SK) (see Figure

2.6). The intronic region contained three in-frame translational stop signals (xxx),

which prevented luciferase expression when the construct is inadequately spliced.

Figure 2.5. Validation of suppressor expression. Expression of the suppressors were analysed by measuring levels of the RFP tag by qPCR. The expression levels varied between the different stable cell lines, with the TARDBP expressing cell line displaying the greatest expression level and the CRNKL1 expressing cell line exhibiting the least. A Kruskal-Wallis test was performed on the data as a whole, resulting in a P<0.01 (**). A Mann-Whitney test was carried out on each of the test cell lines relative to the untransfected control cells (P<0.05 *) (SEM error bars, N=3)

64

Splicing efficiency could therefore be determined by analysing the activity of luciferase

in relation to the activity of the constitutively expressed β-galactosidase.

Initially we transiently transfected HEK293T cells with plasmids expressing exon one

HTT fragments of varying polyQ length, as well as the TN23 construct. After 48 hour,

cells were analysed using a dual reporter assay. The results demonstrated a decline in

splicing efficiency in response to increasing polyQ length, an effect coupled with a

mirrored decline in cell viability (see Figure 2.7). The 25Q and 47Q constructs did not

affect mRNA splicing efficiency or cell viability, while the 72Q and 97Q constructs

reduced splicing by ~20%. This splicing defect was associated with a drop in cell

viability of ~50%.

To more precisely determine the onset of the splicing defect, cells were assayed 24

hours post transfection. These assays revealed no significant change in splicing at 24

hours, or polyQ length dependent cell toxicity, suggesting insufficient expression of

these constructs at this time point (see Figure 2.8). Cells appeared to have a

somewhat lower cell viability compared to those transfected with the RFP construct,

though this may simply reflect a recovery state following transfection. Nonetheless,

these data suggest that the toxicity observed and the splicing defects are intimately

linked.

65

Figure 2.6. Splicing efficiency assay used to ascertain the possibility of splicing defects in HD. The splicing efficiency assay utilised a construct encoding β-galactosidase (β-gal) and luciferase which are separated by an intron flanked by recombinant fragments encoding adenovirus (Ad) and skeletal muscle isoforms of human tropomyosin (SK). These fragments contain three in-frame translational stop signals (xxx), which prevent luciferase expression when the construct is inadequately spliced. Splicing efficiency can be determined by analysing the activity of luciferase in relation to the constitutively expressed β -galactosidase activity (Nasim and Eperon, 2006).

66

A.

B.

Figure 2.7. Expansion of the HTT polyglutamine tract causes a decline in splicing efficiency and cell viability 48 hours post transfection. A. Constructs expressing HTT fragments containing different length polyglutamine tracts were transiently transfected into HEK293 cells, alongside a splicing efficiency construct. Splicing efficacy was assayed 48 hours post transfection, and results were subsequently normalised to control cells transfected with an RPF expressing plasmid. The 25Q and 47Q constructs did not alter mRNA splicing, while the 72Q and 97Q fragments reduced splicing efficiency (RFP N=24, 25Q N=12, 47Q N=15, 72Q N=15, 97Q N=14). Kruskal-Wallis, P<0.001 (****) with Dunns multiple comparison, P<0.001 (****). B. The viability of transiently transfected cells was tested using of the WST-1 assay. Cells displayed a trend similar to that seen in the splicing efficiency assay (untransfected N=18, RFP N=18, 25Q N=12, 47Q N=12, 72Q N=12, 97Q N=9, SEM error bars). Kruskal-Wallis, P<0.001 (****) with Dunns multiple comparison, P<0.001 (**)

67

A.

B.

Figure 2.8. Mutant HTT constructs do not significantly change splicing efficiency or cell viability 24 hours post transfection. A. Constructs expressing HTT fragments containing different length polyglutamine tracts were transiently transfected into HEK293 cells, alongside a splicing efficiency construct. Splicing efficiency was assayed 24 hours post transfection, and results were subsequently normalised to control cells transfected with an RPF plasmid. No significant change in splicing efficiency was observed (RFP N=11, 25Q N=12, 47Q N=12, 72Q N=12, 97Q N=12). Kruskal-Wallis with Dunns multiple comparison. B. Transiently transfected cells were tested in terms of cell viability with the use of the WST-1 assay. Cells failed to display the polyQ length dependent decline in cell viability seen at the 48 hour time point. RFP N=12, 25Q N=12, 47Q N=12, 72Q N=12, 97Q N=12, SEM error bars). Kruskal-Wallis with Dunns multiple comparison P<0.05 (*).

68

2.34 Overexpression of suppressors does not ameliorate polyQ-dependent splicing

defects

We next sought to investigate whether our splicing-related suppressors were capable

of correcting the splicing defect observed in HEK293T cells transfected with mutant

HTT97Q constructs at 48 hours.

When analysing the effect of overexpression on splicing efficiency, we found that in

general our candidate genes did not improve this metric, although overexpression of

CRNKL did appear to reduce splicing by ~12% compared to control cells transfected

with RFP (see Figure 2.9a). Furthermore upon transfection of these suppressors

alongside the HTT fragment constructs, we observed no significant change in the

splicing defect brought about by the 97Q construct (see Figure 2.9b). ~20% drop in

splicing efficacy attributed to expression of the 97Q construct, relative to the

construct containing the shorter HTT fragment, was consistent in all cells co-

transfected with the overexpression constructs. The decline in splicing efficiency due

to CRNKL overexpression remained when transfected into cells along with the 25Q

construct. Interestingly, when co-transfected, the 97Q construct and CRNKL reduced

splicing efficiency in an additive manner, though this was not statistically significant.

69

B.

Figure 2.9. Overexpression of the candidate genes does not alter splicing efficiency in HTT97Q cells. A. HEK293T cells were transfected with both the TN23 splicing efficiency construct and the suppressor overexpression constructs. Splicing efficiency was assayed 48 hours post tranfection using a dual reporter system. Overexpression of EFTUD2, HNRNPF, HNRNPK or TARDBP did not alter splicing efficiency, while expression of CRNKL reduced this metric (N=9- 12, SEM error bars). Kruskal-Wallis, P<0.001 (****) with Dunns multiple comparison, P<0.05 (*)B. Transfection of overexpression constructs, along with the TN23 and HTT constructs did not alter the splicing defect caused by the 97Q HTT expression (N=9, SEM error bars). Kruskal-Wallis, P<0.001 (****) with Dunns multiple comparison (no significant difference between cells transfected with the 25Q construct and those transfected with the 97Q construct). Expansion of the polyglutamine tract correlates with a decline in splicing efficiency and cell viability 48 hours post transfection. A. Lentiviral constructs expressing htt fragments of with different polyglutamine lengths were transiently transfected into HEK293 cells, alongside the splicing efficiency construct. The splicing efficiency was assay 48 hours post transfection, and results were subsequently normalised to control cells transfected with an RPF plasmid. The 25Q and 47Q constructs do not alter

A.

70

2.35 Overexpression of suppressors alters the aggregation dynamics of mutant HTT

Given that overexpression of our suppressors appeared to have little effect on the

mRNA splicing defect, we next sought to examine whether these suppressors

functioned by altering mutant HTT aggregation dynamics. To do so, we employed the

automated cell analysis apparatus system, Cellomics, to study mutant HTT aggregation

dynamics in PC12 cells expressing candidate modifiers. Prior to analysis, cells were

stained with the nucleus dye- Hoechst 33342 to facilitate cell identification (see Figure

2.10). The Cellomics platform subsequently extrapolated from the nucleus to define

an area encompassing the cytoplasm, which was then analysed in terms of green

fluorescence.

Figure 2.10. Cellomics cell identification and analysis. (A) PC12 cells displayed diffuse expression of the mutant HTT fragment, as well the condensed formation of inclusions bodies, 24 hours after ponasterone induction. (B and C) The Target activation bioapplication identifies the nucleus of the cell with the aid of the DNA binding dye, Hoechst 33342. (D) The software extrapolates the nucleus to define the cell body. (E) The bioapplication analysis measure GFP intensity and identifies large inclusion bodies (labelling in red).

71

Expression of mutant HTT was induced at various time points and its expression and

aggregation examined using the parameters stated in the methods section. An

average green fluorescent intensity threshold was used to distinguish between cells

expressing mutant HTT and those which failed to respond to the inducing agent,

ponasterone A. At 48 hours, approximately 80% of control cells and those

overexpressing CRNKL, HNRNPF and HNRNPK exceeded this threshold, while cell lines

overexpressing TARDBP and EFTUD2 displayed significantly fewer mutant HTT

expressing cells (see Figure 2.11a). In each case the percentage of HTT expressing cells

appeared to plateau by around 48 hours.

By analysing the variation in GFP intensity, cells displaying inclusion bodies or other

large aggregate species were differentiated from those exhibiting a diffuse,

homogeneous expression of mutant HTT. This metric was used to gauge changes in

the aggregation dynamics in response to overexpression of the suppressors. At 48

hours, cells lines over expressing CRNKL, HNRNPK and TARDBP displayed a significantly

greater proportion of cells containing large aggregates/inclusions, indicative of an

accelerated rate of aggregation. In stark contrast, cell lines overexpressing EFTUD2

and HNRNPF both displayed a diminished level of mutant HTT of aggregation relative

to control cells (see Figure 2.11b).

72

A.

B.

73

2.36 Overexpression of suppressors brought alters HTT protein levels

To further assess the changes in HTT expression and aggregation we next employed a

well-established filter retardation assay to determine whether cell lines displayed

differing levels of total and aggregated HTT protein 48 hours post induction (Wanker

et al., 1998). Protein was extracted from lysed cells, and subsequently denatured in

SDS. Samples were subsequently passed through either a nitrocellulose or cellulose

acetate membrane. While the former membrane was used to detect total protein

levels, only aggregated proteins adhered to the latter membrane.

The nitrocellulose dot blot revealed changes in HTT protein levels in three of the cell

lines. Cells overexpressing CRNKL displayed a slightly elevated level of HTT following

induction, while those expressing HNRNPK and TARDDP contained considerably more

of the mutant protein. Cell line overexpressing EFTUD2 and HNRNPF exhibited HTT

levels comparable to the control cells (See Figure 2.12).

The cellulose acetate filter trap method was used to detect levels of SDS insoluble

mutant HTT aggregates. Consistent with the Cellomics data, cell lines overexpressing

CRNKL, HNRNPK and TARDBP exhibited a marked increase in aggregate levels in

relation in the control cells, 48 hours after induction. In contrast, cell lines expressing

EFTUD2 and HNRNPF demonstrated a significant reduction in aggregation (See Figure

2.13).

Figure 2.11. Mutant HTT aggregation is significantly altered by overexpression of splicing genes. A. Cellomics software was used to analyse aggregation of mutant HTT. The software identified cells with the use of Hoescht nuclear dye, and extrapolated from this area to encompass the cytoplasm. The target activation application was used to examine the green fluorescence found in the cytoplasm, indicative of mutant HTT expression. After 48 hours of ponasterone induction cell lines overexpressing TARDBP or EFTUD2 displayed lower HTT expression relative to control cells B. Overexpression of CRNKL, HNRNPK and TARDBP increased the level of inclusion body formation, while overexpression of EFTUD2 and HNRNPF significantly reduced formation of these aggregate species (N=3 for each cell line and time point). A two way ANOVA was performed on data sets, with a Bonferoni post-test (*P<0.05, ***P<0.001)

74

A.

B.

Figure 2.12. Mutant HTT levels were elevated in cell lines overexpressing CRNKL, HNRNPK and TARDBP, and reduced in cells overexpressing EFTUD2 and HNRNPF. A. Nitrocellulose dot blots (representative blot shown) revealed changes in total HTT protein levels 48 hours post induction. Cell lines overexpressing CRNKL, HNRNPK and TARDBP all displayed an increased level of the HTT protein in relation to control cells. Cell lines overexpressing EFTUD2 and HNRNPF contained HTT levels similar to those of the control cells. B. The nitrocellulose blots were analysed with the use of Image J, and normalised to both the protein amount and to the control cells. (N=4, SEM error bars). The Wilcoxon test was used to determine statistical significance.

75

A.

B.

Figure 2.13. Mutant HTT aggregation levels were found to be consistent with the data derived from the automated cell analysis software. A. The cellulose acetate blot showed (representative blot) pronounced changes in mutant HTT aggregation levels displayed by each of the cell lines 48 hours post induction. Cell lines overexpressing CRNKL, HNRNPK and TARDBP all displayed a substantial increase in aggregates/ inclusions relative to control cells. Cell lines overexpressing EFTUD2 and HNRNPF lower levels of aggregate formation to those of the control cells. B. The cellulose acetate blots were analysed with the use of Image J, and normalised to both the protein amount and to the control cells. (N=3, SEM errors bars). The Kruskal-Wallis test was used to determine statistical significance.

76

2.37 Quantitative real-time PCR revealed changes in mutant HTT expression with no

variation in HTT copy number between cell lines.

In order to determine whether changes in total and aggregated HTT protein levels

were due to transcriptional changes, we analysed RNA transcript levels by RT-PCR.

Quantitative RT-PCR was performed using primers that annealed within the EGFP tag

of the HTT construct, and results normalised to untransfected control cells. The results

were consistent with those obtained from the nitrocellulose dot blot, indicative of

transcriptional rather than translational variations between cell lines (see Figure 2.14).

However, the differences in HTT aggregation appeared to be much more pronounced

than the differences in expression, thus such changes in aggregation dynamics may

not be solely attributed to transcriptional variations between cell lines.

To ensure these changes were independent of HTT copy number, we isolated genomic

DNA from each of the stably transfected cell lines, and analysed samples via RT-PCR.

Our results illustrated no significant difference in HTT construct copy number between

cell lines, though cells overexpressing EFTUD2 displayed a slight reduction of around

5% (see Figure 2.15). Given that changes in copy number should result in large

changes in signal- unless a significant number of integration events have occurred- this

5% reduction most likely represents noise in the assay.

77

Figure 2.14. Overexpression cell lines display variation in HTT expression relative to untransfected control cells. Mutant HTT expression was assessed via quantitative real-time PCR, with results normalised to the reference gene, PIPB. The results illustrated a higher expression of HTT in cell lines overexpressing CRNKL, HNRNPK and TARDBP, and lower expression in those overexpressing EFTUD2 and HNRNPF. These results were analogous to changes in protein levels. (N=3). The Kruskal-Wallis and Man-Whitney test was used to determine statistical significance.

Figure 2.15. Stably transfected cell lines exhibit little variation in HTT copy number. Mutant HTT copy number was assessed via quantitative real-time PCR, with results normalised to the reference gene, PIPB. The results illustrated no dramatic variation in HTT copy number in cells overexpressing CRNKL, HNRNPF, HNRNPK or TARDBP, though cells overexpressing EFTUD2 displayed a significant, though modest reduction (N=3). The Kruskal-Wallis and Man-Whitney test was used to determine statistical significance.

78

2.4 Discussion

Following the generation of stable cell lines overexpressing our gene of interest, we

sought to determine whether these candidates suppressed mutant HTT phenotypes in

mammalian cells, as previously demonstrated in yeast. Our initial cytotoxicity assay

revealed that five of the nine mRNA splicing factors convey suppression in our

mammalian cells line, two did not modulate toxicity relative to untransfected control

cells, and the remaining genes, HNRNPU and SRFS, exacerbated this phenotype.

The mechanisms by which overexpression of both HNRNPU and SRFS increased HTT

induced toxicity in our stable mammalian cells is unclear, and we did not explore this

further during the course of this project. Given that HNRNPU encodes a global splicing

factor, overexpression may have altered the delicate balance of splicing events

regulated by this protein (Xiao et al., 2012). As demonstrated previously,

overexpression of HNRNPU can lead to increased expression of proteins involved in

apoptosis, such as immediate early response (IER3) proteins and tumour necrosis

factor alpha (TNF-α), through the stabilisation of mRNA transcripts (Yugami et al.,

2007). Moreover, HNRNPU overexpression has been shown to enhance splicing of

caspase-9, leading to an increase in the pro-apoptotic splicing variant, caspase-9a. It is

therefore feasible that such effect would induce apoptosis, and in turn influence the

caspase-3/7 assay (Vu et al., 2013). Similarly, overexpression of SFRS has been shown

to cause the aberrant splicing of the gene encoding tropomyosin receptor kinase B

receptor (TrkB), a component of the BDNF signalling pathway. This splicing defect has

been noted in AD, and leads to a reduction in the functional protein, reduced levels of

which have also been observed in HD (Wong et al., 2012, Ginés et al., 2006).

Alternatively, both HNRNPU and SFRS, like the other splicing factors derived from the

initial yeast modifier screen, contain RNA binding domains, which can mediate

aggregation of these proteins (Vanderweyde et al., 2013). Other RNA binding proteins

such as TLS (translocated in liposarcoma), proteins with WW domains, and indeed

TARDBP have been shown to interact with HTT. Therefore it is possible that

overexpression of HNRNPU and SFRS may contribute to HTT aggregation via

sequestration, though further investigation would be required to deduce the correct

79

mechanism/(s) involved (Vanderweyde et al., 2013, Doi et al., 2008, Fuentealba et al.,

2010, Passani et al., 2000).

Of the nine potential suppressor candidates, overexpression of CRNKL, EFTUD2,

HNRNPF, HNRNPK and TARDBP suppressed mutant HTT toxicity. Despite their unifying

role in mRNA splicing, after performing the splicing efficiency assay on cells

transfected with both the suppressors and the HTT constructs, it became apparent

that the mechanisms by which these proteins conveyed suppression was independent

of this post-transcriptional process. Given that our candidate genes were derived

from a yeast genetic modifier screen, an organism in which mRNA splicing is

considered to be a minor process (~3.8% of genes contain introns), it is perhaps

unsurprising these genes act to suppress HTT toxicity via a mechanism other than

splicing (Dujon, 1996, Lopez and Seraphin, 1999). Moreover, as our splicing assay was

based upon the splicing of a single construct, it is feasible that this assay may not be

reflective of global splicing.

We next examined mutant HTT aggregation dynamics by analysing inclusion body

formation with the aid of an automated cell analysis platform. These result suggested

that modulation of aggregation is likely to underline the suppression of toxicity. While

overexpression of CRNKL, HNRNPK, and TARDBP accelerated HTT aggregation,

overexpressing EFTUD2 and HNRNPF delayed aggregation. These results were

confirmed biochemically using filter retardation assays. Further investigation revealed

that such effects were attributed, at least in part, to changes in HTT expression

following induction. Such changes in expression could reflect variation at the

transcriptional level or mRNA stability.

Many of our candidate genes encode for multifunctional proteins, and thus have roles

in addition to mRNA splicing. For example HNRNPK has documented roles in both

mRNA stability and transcription. This nuclear protein has been shown to bind to

promoter regions and alter expression of the regulated gene (Lee et al., 1996,

Ostrowski et al., 2003). The inducible PC12 HD model employed in my study, contains

two stably transfected constructs, one of which encode the ecdysone receptor and the

retinoid X receptor under control of the CMV promoter. The second construct consists

80

of a binding site for the ecdysone-retinoid X receptor heterodimer, which modulates

transcription of the HTT fragment from the HSP minimal promoter when in the

presence of ponasterone A. It is feasible that HNRNPK could bind to either of these

promoter sequences, with overexpression bringing about an increase in transcription

of the genes encoding the receptors or the HTT fragment.

Similarly HNRNPF has been implicated as a negative regulator of polyadenylation.

Working in an antagonistic manner with its sister protein HNRNPH, HNRNPF reduces

gene expression by blocking the binding of cleavage stimulatory factor 2 (CstF2) to

SV40 polyadenylation site (Alkan et al., 2006, Yoshida et al., 1999). CstF2 plays a

crucial role in the polyadenylation of mRNA, and by extension the stability and

translation the mRNA transcripts (MacDonald et al., 1994). Overexpression of HNRNPF

in mammalian cells has previously been shown to reduce the expression of certain

genes, an effect attributed to an altered ratio of HNRNPF to HNRNPH (Veraldi et al.,

2001). Given that both the inherent constructs found in our PC12 cells given rise to

transcripts containing SV40 sequences, it is possible that overexpression of HNRNPF

prevents polyadenylation of these transcripts, leading to a reduction in HTT

expression, a theory consistent with our data.

Surprisingly overexpression of TARDBP was also found to be protective in our

neuronal cell lines. Previous studies conducted in C.elegans and murine models

reported that TARDBP overexpression induced neurological dysfunction and

symptoms synonymous with ALS and frontotemporal lobar degeneration (FTLD) (Tsai

et al., 2010, Ash et al., 2010, Wils et al., 2010). Such contradictory finding may be

attributed to differences in models, or differences in levels of expression. As with the

previous candidate genes, the mechanism surrounding protection appeared to lie at

the transcriptional level. In addition to its role in mRNA splicing, TARDBP has been

linked to transcriptional regulation, modulating assembly of transcription factors,

binding to promoter sequences and promoting mRNA stability (Ou et al., 1995,

Acharya et al., 2006, Strong et al., 2007). It is therefore possible TARBDP may

influence the expression of HTT at any of these transcriptional levels, though further

research would be required to elucidate the precise mechanism.

81

Both EFTUD2 and CRNKL encode for components of the spliceosome, and their

documented functions have been confined to the process of mRNA splicing. It was

therefore somewhat surprising that overexpression of these genes brought about

changes in the expression of the HTT fragment. In both cases depletion or knockout of

the encoded proteins has been shown to have a deleterious effect on splicing, with

lethal or disease-causing consequences (Lines et al., 2012, Frazer et al., 2009, Chung et

al., 2002). In our study, overexpression of CRNKL or EFTUD2, reduced splicing

efficiency by around 13 and 9% respectively (though the latter result failed to achieve

statistical significance). It is therefore possible that overexpression of these genes may

alter splicing of transcription factors, or proteins modulating post-transcriptional

processing such as capping and polyadenylation, ultimately altering expression of our

HTT fragment indirectly. Again, further study is required to determine the mechanisms

by which overexpression of these genes convey suppression against mutant HTT

toxicity in our neuronal cell lines. Interestingly of the our five genes of interest, both

CRNKL and EFTUD2 appear in the HD research crossroads database, and have been

assigned a score of 3, indicating that these genes have a causal relationship with HD,

and as such warrant further study (Kalathur et al., 2012).

Although our results suggest that the candidate genes largely appeared to suppress

mutant HTT via a mechanism independent of mRNA splicing, the splicing efficiency

assay revealed a polyQ length dependent decline in splicing efficiency in HEK293T

cells. Cell transfected with lentiviral constructs expressing 93Q HTT fragments

displayed a 20% drop in splicing efficiency relative to those transfected with the 25Q

construct. As highlighted in Chapter 1, neurodegenerative disorders have recently

been linked to defects in mRNA splicing, though work in the context of HD is scarce.

(Mills and Janitz, 2012b). Given that our splicing efficiency assay is designed to

represent global splicing, a defect in this form of RNA processing is likely to affect the

expression many genes (Licatalosi and Darnell, 2006, Cooper et al., 2009). This was

demonstrated by Jia et al, who found mutations in DNA encoding U2 snRNA, a

component of U2 snRNP, resulted a drop in splicing efficiency of ~28% in vitro (using

the same assay we utilised in our study), as well as global disruption in mRNA splicing

coupled with neurodegeneration (most prominent in the cerebellum) and ataxia in

82

mutant mice (Jia et al., 2012). The same group found many genes encoding mRNA

splicing factors to be greatly enriched in mutant mice, and proposed this to be a

compensatory mechanism to restore the splicing deficit. They further postulated such

attempts may only enhance abnormal splicing patterns seen in disease states (Jia et

al., 2012).

Our results suggest that, similar to the study above, aberrant splicing may play a role

in the neurodegeneration seen in HD. Such claims are supported by recent work

regarding abnormal splicing of genes encoding chloride channel 1 (ClC-1) in the

skeletal muscle of R6/2 mice, and indeed of HTT itself (Sathasivam et al., 2013, Waters

et al., 2013). Furthermore microarray data compiled from both HD patients and

transgenic murine models indicate differential expression of splicing factors in the HD

disease state, while RNAi screens conducted in both C.elegans and Drosophila models

of HD also reveal a significant enrichment in genes relating to RNA splicing (Kalathur

et al., 2012, Nollen et al., 2004, Zhang et al., 2010). Changes in this post-transcriptional

process during disease progression demands greater investigation and may provide

new avenues for therapeutic intervention.

In conclusion, many of our candidate genes modulated aggregation dynamics of

mutant HTT, an affect partly attributed to alterations in the expression of HTT in the

PC12 mammalian cell line. Such expression changes may have been linked to

transcription or post-transcriptional processes. The results derived from our splicing

efficiency assay suggest that aberrant splicing may have a role in the pathogenesis of

HD, a theory consistent with recently published work. The role of mRNA splicing in

disease progression therefore requires substantially more research and may provide

greater insight into this devastating condition.

83

2.5 Future work

Due to time constraints, I was unable to further investigate many of the results

obtained from this portion of the project. For example overexpression of HNRNPU or

SFRS3 exacerbated HTT toxicity in PC12 cells. It is therefore possible that knockdown

of these candidate genes may have a beneficial effect (though as demonstrated in

Chapter 3, such symmetry may not arise).

During the course of the in vitro cellular work, we attempted to examine the co-

localisation of EGFP tagged mutant HTT and the mRFP tagged suppressors. However,

this work was hampered by the extremely strong EGFP signal following HTT

aggregation relative to the weak fluorescent signal emitted by the red fluorophore. In

order to strengthen and shift the red signal further to the far red, I employed an RFP

booster ATTO 594, consisting of an RFP binding protein conjugated to a zwitterionic

dye with an absorbance of 601nm and emission of 627nm. Despite the shift in

fluorescence, the strong signal of the EGFP tag continued to penetrate the red filter,

making differentiation of the two fluorophores difficult. With more time, an

immunocytochemical approach, utilising an RFP-binding primary antibody and far-red

secondary antibody, may sufficiently separate the emissions of the two fluorophores.

Our attempts to elucidate the mechanisms surrounding the protection conveyed

following overexpression of our candidate gene, revealed changes in transcription.

Although transcriptional abnormalities have been noted previously in HD, changes in

the transcription of the HTT gene itself have not (Cha, 2007, Mazarei et al., 2009,

Hodges et al., 2006). Our results could be attributed to global transcriptional changes,

the result of altered mRNA degradation or other post transcriptional processes.

Conversely, given that HTT expression in our cellular model is regulated by a non-

endogenous promoter system, our results may simply reflect the artificial nature of

the experiment. Therefore it is necessary to further examine our candidate genes in a

transcriptional context. This could be achieved via transcriptional profiling of our cells

line in both the induced and repressed states.

Other methods to further explore our results may include the transfection of a

reporter system into our stable PC12 cell lines, using the same promoters as those

84

regulating HTT expression. Variations in the expression of the reporter may reveal

whether our findings are the result of general transcriptional effects of the promoter

or whether this may reflect effects specific to HTT. Furthermore, mRNA stability may

be examined via the isolation of cytoplasmic mRNA and the subsequent analysis using

a qPCR method against HTT. Given that nonsense mediated decay (NMD) occurs in the

cytoplasm, variations in cytoplasmic mRNA levels between cell lines may be indicative

of changes in mRNA degradation (Parker and Sheth, 2007).

One of the more interesting results from this portion of my thesis research came from

analysis of the candidate suppressors using splicing efficiency assays. I found that

transfection of our HTT constructs into HEK293T cells brought about a polyQ length

dependent splicing deficit. Defects in splicing may therefore play a role in the

manifestation and/or progression of HD. Research examining the role of mRNA

splicing in HD pathogenesis is still very much in its infancy, though our work suggests

that it could provide new insight into disease progression. Future work could include

the use of exon arrays and next generation sequencing techniques, such as RNA-seq,

to examine splicing changes of specific genes in HD (Mills and Janitz, 2012a). For

example, the Affymetrix Human Exon Array ST 1.0 consists of 1.4 million probe sets

used to analyse all known and predicted human exons. These probes can be used to

detect the level and relative abundance of different transcript isoforms (Okoniewski et

al., 2007). This technique has been successfully used to examine mRNA splicing

changes arising in certain forms of cancer (Gardina et al., 2006, Thorsen et al., 2008, Xi

et al., 2008).

With advances in next generation sequencing, sequencing of the entire transcriptome

is now possible in the form of RNA-seq. This technique consists of reverse transcribing

an entire population of RNA into cDNA, which is subsequently amplified and

sequenced into reads of approximately 400bps. These short fragments are then

aligned according to a reference and analysed. Similar to the microarray approach, the

RNA-seq technique can be used to determine levels of transcript isoforms, though

with a much higher fidelity (Wang et al., 2009). RNA-seq has been postulated as

means of examining the complex splicing events within brains, and by extension

changes is such events underlying neurological disorders (Sutherland et al., 2011).

85

Finally, in light of recent unpublished work conducted by De Almeida et al, one of

caveats of our splicing efficiency experiment may have been the use of HEK293T cells.

Work conducted by this group have highlighted a divergence in splicing patterns in

HEK293T cells, relative to other commonly used cell lines. As such, it will be necessary

to examine this polyQ length dependent splicing defect in HeLa cells or primary

neurons, as well as in vivo models.

86

Chapter 3

Exploring the role of splicing genes in Huntington’s disease using C. elegans

3.1 Introduction

3.11 The use of C. elegans as a model organism

The potential of Caenorhabitis elegans (C. elegans) as a versatile genetic model was

first highlighted by Sydney Brenner, in his seminal paper entitled “the genetics of

Caenorhabitis elegans”. Brenner described a comprehensive method by which C

.elegans strains were maintained and utilised in both forward genetics and compound

screens, as well as emphasising the advantages of this model over other invertebrate

models such as Drosophilia melanogaster (Brenner, 1974). Since publication, C.

elegans has been extensively studied, revealing a high degree of conservation

between this organism and humans, prompting many to select this model as a means

of bridging the gap between in vitro and in vivo work in rodents (Consortium, 1998).

The ease with which C. elegans can be grown and maintained in a laboratory makes its

extremely attractive as a model organism. Their short generation time, lasting

approximately 3-4 days (see Figure 3.1) and small size, coupled with their capacity to

grow on agar plates surviving on a diet of E. coli, makes this organism ideal for high-

throughput experiments. Furthermore, the transparency inherent to worms

throughout their lifespan allows metabolic processes and multiple cell types to be

viewed in real-time with the use of fluorescent markers (see Figure 3.2). Despite the

advantages of this model, the relative simplicity of their organs and absence of many

molecular pathways can limit comparison to our own physiology (Brignull et al., 2006,

Kaletta and Hengartner, 2006).

87

Figure 3.1. C .elegans life cyle. The C. elegans life span is approximately three to four weeks, with a generation time of around four days. This latter period is made up of an embryonic phase, followed by four larval stages, in which worms shed their cuticle and synthesise a new larval stage specific cuticle. Post embryonic development is prompted by the presence of food, without which, larva have the capacity to inhibit their development, surviving for up to ten days. In the presence of food, larvae begin the developmental process three hours after hatching. The four larval stages (L1-L4) are accompanied by the maturation of the nervous, reproductive and elementary immune (coelomocytes) systems (Wood, 1988). Following the L1 developmental stage, worms can enter the dauer phase, a state of growth arrest in response to overcrowding or scarcity of food, and can remain in this form for up to four months. In the dauer phase animals exhibit a reduced level of locomotion and feeding until favourable conditions are re-established, initiating further larval development (Cassada and Russell, 1975). Upon reaching adulthood, worms are capable of producing oocytes for up to four days, after which they live for an additional ten to fifteen days (Wood, 1988)(Adapted from wormatlas.org).

The rudimentary nervous system of C. elegans consists of a mere 302 neurons, making

it an ideal organism in which to analyse the neural circuitry underlying behaviour (see

Figure 3.2). C. elegans is the only organism to have had its entire nervous system

mapped, a task facilitated by the simplicity of its synaptic connectivity. The majority of

neurons are unbranched forming around 5000 synapses, many of which being dyads

88

or triads (White et al., 1986). Such information is invaluable in deconvoluting our own,

vastly more complex nervous system and in turn understanding diseases of the

nervous system, such as Parkinson’s Disease (PD), Alzheimer’s disease (AD) and

Huntington’s disease (HD)(Kaletta and Hengartner, 2006, Chatterjee and Sinha, 2008,

White et al., 1986).

3.12 C. elegans and RNAi

The potential of C. elegans as a genetic model was further emphasized by the

development of RNA interference (RNAi) technology, the post-transcriptional knock

down of gene expression, which was first described in this nematode. Initial RNAi work

focused on the use of single stranded antisense or sense strand RNA, homologous to a

specific region of a gene, which produced notable, though inconsistent, reduction of

gene expression (Fire et al., 1991, Hunter, 1999). The research spearheaded by

Andrew Fire and Craig C. Mello led to significant advances, revealing the importance

of RNA structure and the greater efficacy of knock down afforded by double stranded

RNA (Fire et al., 1998). It is believed that the variability arising from the use of single

stranded RNA was predominantly attributed to the presence of contaminating

complementary RNA, leading to the formation of a double stranded structure. The use

of purified single stranded RNA was found to result in a significantly lower level of

knock down (Montgomery and Fire, 1998, Hunter, 1999). These studies consolidated

this method as a viable technique for genetic manipulation.

89

Figure 3.2. Schematic, confocal and fluorescence diagrams of C. elegans. C. elegans possesses a simple nervous system comprising of 302 sensory, motor and inter neurons. 282 of these neurons are found along the body of the worm at several ganglia that constitute the somatic nervous system, while the remainder are located within the pharynx forming the pharyngeal nervous system. The nervous system consists of the nerve ring and two parallel nerve cords (dorsal and ventral), which are connected via commissural tracts. The majority of the neuron cell bodies are located in the head, forming large ganglia or the “brain” of the worm. Four sensory neurons are highlighted in the diagram above, the ASH, ASI, PHA and PHB neurons (the majority of the neurons found within C. elegans are paired, though for simplicity both the left and right neurons are referred to as a singular body). The two latter neurons are visible with the use of GFP expressed under a sensory neuron specific promoter (White et al., 1986, Li, 2001, Faber et al., 1999).

Following optimisation of the RNAi structure, Andrew Fire and Lisa Timmons sought to

simplify the RNAi delivery system by utilising the interference effect exhibited by the

dsRNA, i.e. the spreading of the interference from the initial site of injection, a

dispersal which was found to be more pronounced upon ingestion. Feeding worms

with an RNase III deficient E.coli strain containing the RNAi constructs produced large

quantities of the dsRNA following ingestion, which subsequently spread throughout

the animal’s body. The knock down of specific genes was comparable to that

90

generated by direct microinjection, and could be maintained over several generations

by the continued feeding of worms with these RNAi expressing bacteria (Fire et al.,

1998, Timmons et al., 2001, Timmons and Fire, 1998). This method of delivery has a

number of advantages over the more arduous injection technique, most notably the

cost and ease of performing genetic screens, as well as the capacity to alter RNAi

levels to uncover lesser phenotypes attributed to the knockout of pleiotropic genes

(Kamath et al., 2001).

The mechanism of RNAi induced post-transcriptional gene silencing (PTGS) in C.

elegans has yet to be fully characterised, and as such, many of the proteins involved

are unknown. Upon entering the organism, dsRNA is cleaved by the Dicer complex to

form small interference RNA (siRNA). These smaller structures are subsequently

unwound and bind to target mRNA, forming a platform for the assembly of the RNA

induced silencing complex (RISC), which catalyses degradation of mRNA, and in turn

generates genetic knock down (see Figure 3.3) (Grishok, 2005) .

91

Figure 3.3. Schematic of post transcriptional silencing in C. elegans. Upon ingestion dsRNA undergoes cleavage to form primary siRNA. The first step is catalysed by the Dicer complex, consisting of Dicer, a dsRNA specific RNAIII ribonuclease; Rde-1 (RNAi deficient 1), a dsRNA binding protein thought to facilitate transport of the siRNA to the downstream complex; Drh-1(Dicer related helicase 1); and Rde-4 (RNAi deficient 4), a dsRNA binding protein that works in concert with Rde-1 to retain dsRNA for cleavage by Dicer (Tabara et al., 2002, Grishok, 2005). Following cleavage, the siRNA can be degraded by eri-1 (enhanced RNAi 1), an exonuclease found in the gonads and certain neurons, thus antagonising knockdown(Kennedy et al., 2004). Conversely successful knockdown is achieved through unwinding of the siRNA and binding of the resulting single stranded siRNA to target mRNA. This double stranded structure provides an assembly site for the RNA induced silencing complex (RISC), comprising of several components (many unknown) including Tsn-1 (Tudor staphylococcal nuclease homolog 1), a protein containing 5 nuclease domains and Vig-1 (Vasa intronic gene 1), an RNA binding protein. Upon formation, the RISC complex is believed to catalyse the degradation of mRNA (Caudy et al., 2003, Grishok, 2005). In addition to the RISC complex, the single stranded siRNA act as primers for amplification catalysed by RNA dependent-RNA polymerases such as the germ line specific Ego-1 (enhancer of glip 1) or the somatic cell polymerase, Rrf-1 (RNA-dependent RNA polymerase 1). Either of these polymerases forms a complex along with other proteins, such as the polymerase β nucleotidyltransferase, Rde-3 (RNAi deficient 3), to reform the dsRNA, and provide a substrate for Dicer complex(Grishok, 2005). Adapted from (Grishok, 2005)

92

3.13 C. elegans and Huntington’s disease

C. elegans models have provided valuable insight into many neurodegenerative

diseases. The generation of transgenic strains expressing mutant variants of disease

causing genes can often recapitulate the human disease phenotype. Worms

expressing human β-amyloid peptide (Aβ-42), the toxic species attributed to AD, form

amyloid plaques and trigger cellular toxicity in muscle cells (Link, 1995). Over

expression of mutant human α-synuclein, a protein synonymous with PD, leads to loss

of dopaminergic neurons and subsequently motor deficits (Lakso et al., 2003). C.

elegans provides a means of studying the effects of such toxic insults in the context of

a less complex, yet well defined, nervous system compared to higher mammals.

Despite the lack of a worm HTT orthologue, Huntington’s disease (HD) has been

studied in C. elegans with the use of transgenic strains expressing the human gene.

This offers the capacity to dissect the toxic gain-of-function mechanisms underlying

HD pathology. Faber et al first generated such transgenic lines co-expressing GFP and

a human huntingtin fragment with a 150Q tract under a sensory neuron specific

promoter (OSM-10), and observed polyglutamine length-dependent formation of

protein aggregates and neurodegeneration of the ASH, ASI, PHA and PHB neurons (see

Figure 2.2). Toxicity in these strains was assessed through a simple nose touch

sensitivity test and by observing defects in the worm’s capacity to take up a lipophilic

dye via the exposed sensory neuron endings (Faber et al., 1999). The same group

performed genetic screens using these worm strains and found that mutations within

the polyQ enhancer (PQE-1) gene greatly exacerbated the HD phenotype, while

overexpression conveyed protection against neurotoxicity, through an unknown

mechanism (Faber et al., 2002).

Several subsequent genetic screens conducted using C. elegans, have highlighted the

use of this model as a tool for identifying candidates that directly modulate the gain-

of-function mechanisms responsible for mutant HTT toxicity. Such screens have been

made possible with advances in RNA interference (RNAi) technology, which was first

optimised and implemented in C. elegans, and has since proven to be invaluable in

reverse genetics (Brignull et al., 2006, Fire et al., 1998). Several groups have

93

capitalised on the use of this model coupled with RNAi, for example Nollen et al

performed a genome-wide RNAi screen that identified 186 candidate genes, whose

reduced expression enhanced aggregation in transgenic C. elegans strains expressing a

Htt40Q fragment fused to YFP (yellow florescent protein) in body wall muscle cells.

Interestingly many of these disease modifiers were involved in both RNA synthesis and

processing or to protein synthesis and folding. This led the authors to surmise that

aggregation may be attributable to a loss or imbalance of protein homeostasis. This

disequilibrium may be produced by a combination of translational defects and

dysfunctions of upstream biosynthetic processes (Nollen et al., 2004).

More recently, Lejeune et al conducted a similar RNAi screen, in which 6034 genes

were knocked down in a C. elegans strain expressing a 128Q N-terminal huntingtin

fragment in touch receptor neurons (under the mec-3p promoter). Expansion of the

polyglutamine tract resulted in progressive neuron dysfunction leading to an

impairment of the touch response. This screen identified 662 modifiers of neuronal

dysfunction involved in a variety of cellular processes including metabolism,

differentiation, homeostasis and the immune system. Many of these processes have

previously been linked to HD, demonstrating the strength and fidelity of such an

approach in identifying putative therapeutic targets. Interestingly, a third of these

genes had mouse orthologues, 49 of which are transcriptionally dysregulated in the

striatum of two widely used HD mouse models (CHDL knock-in and R6/2 mice). This

study therefore illustrates the use of C. elegans as a means of identifying new

therapeutic targets for HD, as well as validating other genetic models (Lejeune et al.,

2012).

In addition to genetic screens, C. elegans models provide a simple and cost effective

method for testing potential drug candidates. A recent drug screen revealed that

trichostatin A, lithium chloride and mithramycin, compounds previously tested in cell

culture, Drosophilia and mouse models respectively, all supressed polyQ induced

neurotoxicity in HD worms. To understand the mechanisms by which lithium chloride

and mithramycin conveyed neuroprotection, the group performed a starvation assay,

utilising the worm larvae’s capacity to undergo growth arrest in the absence of food.

Although this ability would ordinarily allow the worms to survive for longer periods of

94

time, the group found that this dauer state did not protect the sensory neurons from

the neurotoxic insult of mutant HTT fragments. Both Lithium chloride and

mithramycin promoted neuron survival, despite a lack of growth, suggesting that the

protective properties of the drugs could not be attributed to growth or developmental

pathways. In order to address the possible role of aging in the therapeutic effects of

these compounds, the group also screened the drugs in HD worm strains with a daf-16

null mutation (Voisine et al., 2007). DAF-16 is a transcription factor involved in the

insulin-like pathway, which plays a crucial role in aging. Previous research has shown

that knockdown of this transcription factor exacerbates mutant HTT aggregation and

reduces longevity of C. elegans models of HD, while overexpression extends the

lifespan, suggesting a common DAF-16 target in both aging and mutant HTT

aggregation pathways (Lin et al., 1997, Morley et al., 2002). Both lithium chloride and

mithramycin were found to be neuroprotective in the absence of daf-16, indicating a

daf-16 independent mode of action(Voisine et al., 2007).

Two of the salient features of HD are the formation of aggregates as well as an inverse

correlation between CAG repeats and age of onset. Many studies compare the effects

of very large polyQ tracts relative to shorter, less toxic repeats, with few studies

having examined the transitional cellular changes that occur as these CAG repeats

pass the toxicity threshold. Morley et al generated several worm strains expressing

polyQ-YFP fusion proteins containing varying length expansions (Q29, Q33, Q35, Q40,

Q44, Q64 and Q81) in the body wall muscle cells, and exploited the transparency of C.

elegans to investigate aggregation dynamics. The group found that below 35Q the

fusion proteins displayed a diffuse distribution within the cell, indicative of a soluble

form of the protein, and does not impair motility compared to non-transgenic strains.

Conversely the worms expressing greater than 44Q exhibited large focal aggregates,

correlating with a drastic decline in worm motility. Interestingly, the 40Q strains

displayed an amalgamation of cells, either with a diffuse distribution of the soluble

protein, or aggregates of the insoluble form. In this case motility deficits were

governed by the ratio of these two cell groups, with 40Q worm containing fewer of

the latter cells displaying a phenotype closer to the non-transgenic strain, thus

95

demonstrating the importance of the threshold in dictating a propensity of polyQ

proteins to aggregate (Morley et al., 2002).

With recent advances in RNAi technology coupled with the inherent advantages

associated with this invertebrate system, the use of C. elegans as a model for the

study of HD has only recently been fully realised. The combined work of many groups

has led to the discovery of pathways, genes and drugs that modulate the toxic effects

of mHTT, as well as confirming findings derived from in vitro studies, all of which has

provided new insight into this neurodegenerative disease.

3.14 Aims

Prior to further validation in HD murine models, we chose to refine the list of splicing

candidate genes that arose from the work conducted in HD mammalian cell lines.

Given the simplicity of the organism coupled with the ease of maintenance and

genetic manipulation, C. elegans provided a rapid intermediate to further examine

the mechanistic roots of the suppression, assessing whether knockdown of these

suppressors exacerbated HD-related phenotype, and therefore complement the our

previous work.

96

3.2 Materials and methods

3.21 Materials

3.211 Bacterial strains and nematode strains

Escherichia coli strain

A tetracycline resistant Ht115 (DE3) E.coli strain was used during the course of the

RNAi knockdown analysis, with the genotype: F-, mcrA, mcrB, IN(rrnD-rrnE)1, lambda-,

rnc14::Tn10, DE3 lysogen: lavUV5 promoter –T7 polymerase, IPTG-inducible T7

polymerase, RNase III (-), available from the Caenorhabditis Genetics Center (CGC).

Transgenic C. elegans strains

To assess the effects of depletion of our suppressor genes on HD phenotypes, we

utilised three C. elegans strains, an N2 wild type strain and two transgenic HD strains.

The first of these transgenic strains (osm-10p::GFP + osm-10p::HtnQ150 + Dpy-20(+))

coexpressed a huntingtin fragment containing 150 CAG repeats along with GFP, while

the second strain (osm-10p::GFP + osm-10p::HtnQ150 + Dpy-20(+) pqe-1) additionally

lacked expression of the polyQ enhancer-1 gene (PQE-1), a mutation which

exacerbated the aggregation of the mutant huntingtin fragment and consequently a

more aggressive phenotype. Huntingtin and GFP expression in both transgenic strains

were driven by the sensory neuron specific OSM-10 (osmosensory) promoter, limiting

expression to the ASH, ASHI, PHA and PHB bilateral sensory neurons. The two

transgenic strains were both generated by the Hart lab, and are available from the

CGC (strains HA659 and HA759 respectively) (Faber et al., 1999, Faber et al., 2002).

3.212 Constructs

L4440 Plasmids

The RNAi constructs were generated by Source Biosciences, genomic fragments were

cloned into the worm-specific L4440 vector (Timmons and Fire feeding vector). The

MCS is flanked by two T7 promoter sequences in an inverted orientation.

97

RNAi oligonucleotide

Generation of the RNAi constructs was carried out by Source Biosciences, using the

RNAi library compiled by the Ahringer lab (Fraser et al., 2000). cDNA was generated

from total worm RNA using the gene specific primers (see Table 3.1). The resulting

cDNA fragments were subsequently cloned into the L4440 plasmid via the EcoRV site.

Human

Gene

Worm

orthologue

Left Primer Right Primer cDNA

product

(bp)

CRNKL1 m03f8.3 TTTTCGCTAGTTTT

TGTCTGAGC

TCGTTGGAGTAG

GGTTTCATAGA

2231

EFTUD2 zk328.2 GCACTTTTGCCGA

AGAAGAC

TCAGCGTACTCGT

TTCGATG

1199

HNRNPF w02d3.4 ACATTGCAGCTCC

CAGAAGT

TTTCATTGTTTTT

GCCACGA

1179

HNRNPK f26b1.2 TTTTTCCCTTCAGT

TCCGTG

AGATCCACCGAA

TCGTTCAC

1185

TARDBP f44g4.4 ACCCTTGGATGAA

GTTAAGGAAA

CGTCTACTTTGTC

TGTGAGCCTT

1119

Table 3.1. Worm orthologues of suppressors and PCR primers. RNAi constructs were generated by amplifying worm orthologues of the splicing genes. Amplified genomic fragments were subsequently cloned in the L4440 plasmid and sequenced. All RNAi constructs were generated by Source Biosciences.

3.212 Media and agar

Nematode growth media (NGM)

C. elegans strains were maintained on NGM media containing 0.3% (w/v) sodium

chloride, 0.5% (w/v) peptone and 3.4% (w/v) agar. Media was autoclaved and

supplemented with: 1 mM calcium chloride; 5 µg/mL cholesterol in ethanol, 1mM

magnesium sulphate, 25 mM potassium phosphate, 1 mM isopropyl β-D-1-

thiogalactopyranoside, 100 units/mL Nystatin, 50 µg/ml of ampicillin, 2 µg/ml uracil,

and distributed into 58 mm x 15 mm petri dishes (Thermoscientific)

98

RNAi expressing bacteria

RNAi-expressing HT115 bacteria were cultivated on luria Bertani (LB) media

supplemented with: 50 μg/mL Ampicillin; 25 μg/mL carbenicillin and 10 μg/ml

tetracycline. Following selection on LB media, bacteria were cultured in LB broth until

an OD600 of 0.5 and seeded onto NGM plates. Plates were then incubated at 37 oC

overnight prior to the addition of the worm strains.

3.22 Methods

C. elegans husbandry

C. elegans strains were transferred onto NGM plates seeded with RNAi expressing

bacteria and incubated for 72 hours at 19 oC, at which point animals at the L4 stage of

the nematode life cycle were picked and placed onto a fresh bacterial lawn. These

worms were selected every two days and transferred onto fresh NGM plates to

prevent hatching and maturation of progeny, and maintain a group of animals of the

same age. Behavioural assays were conducted after 8 days of aging.

Sensitivity assay

The sensitivity of 8-day old adult hermaphroditic worms was assessed by stroking the

anterior body region corresponding to the location of the ASH sensory neurons. A

positive response to this mechanical stimulus was deemed to be a change in the

direction of movement or a break in locomotion. Animals were trailed ten times using

a Leica M165 FC microscope, and the number of positive responses recorded.

Body bend assay

The frequency of body bends during locomotion was assessed in 8-day old adult

hermaphroditic worms. The number of sine wave movements in a one minute period

was scored using Leica M165 FC microscope.

Dye filling assay

A 2 mg/ml stock of 1,1’-dioctadecylt-3,3,3’,3’-tetramethylindodicarbocyanine

perchlorate (DiD) in dimethyl formamide was diluted 1:200 in M9 buffer (0.3% KH2PO4

99

(w/v), 0.6% Na2HPO4 (w/v), 0.05% NaCl (w/v) and 0.1% NH4Cl (w/v)) and 3 day old

adult worms were incubated in the dilute dye solution for one hour. Animals were

then washed three times in M9 buffer and further destained on bacterial lawn for one

hour. Worms were mounted onto agar pads containing sodium azide, and ASH

neurons visualised at 40X magnification using a fluorescence microscope (Ziess

Axiovert 200M, Zeiss Colibri controller) with the Axiovision 4.8 software (Carl Zeiss).

RNA isolation

The N2 C. elegans strain was grown for 6 days on a bacterial lawn expressing each of

the RNAi constructs or an empty vector control. Following this initial incubation,

worms were pelleted and immersed in a bleach solution (5% (v/v) sodium

hypochlorite and 0.5 M sodium hydroxide) for 5 minutes. This bleach solution lysed all

adult and immature worms, while leaving the incipient eggs undamaged. These eggs

were washed several times in sterile distilled water, plated onto fresh bacterial lawns

and incubated for a further 72 hours at 19 oC. Following this second incubation,

mature 3 day old animals (L4 stage) were lysed in TRI reagent (Sigma), with the

resulting RNA precipitated in 75% ethanol and resuspended in RNase/DNase free

water (see Method and Material, Chapter 1). The integrity of the RNA was

subsequently determined using nanophotometer (Implen) (OD260/OD280 nm

absorption ratio >1.85).

Despite the behavioural assays having been performed on 8 day old worms, the RNA

isolation was carried out on 3 day old animals. After three days, worms mature to

adulthood and become capable of laying eggs. RNA samples taken after this period

would reflect knockdown in worms of varying ages, reducing the overall level of

knockdown observed, as 72hours is required for robust RNAi knockdown (Kamath et

al., 2001) . Conversely aged worms, as used in both of the behavioural tests, would

have yielded an inadequate amount of RNA. The bleaching method used in this study

generates a large number of worms of the same age which produce a sufficient

amount of the RNA for quantitative PCR.

100

Reverse transcription and quantitative PCR

Genetic knockdown of the splicing genes was assessed using real-time quantitative

PCR. 250 ng of RNA was reverse transcribed to form single stranded cDNA using the

ImProm-II™ Reverse Transcription System (Promega). The subsequent analysis was

performed using a Applied Biosystems StepOne™plus Real-Time PCR System coupled

with Fast SYBR® Green Master Mix (Applied Biosystems).Transcripts were amplified

using 0.4 µM forward and reverse primers (see Table 3.2), and data normalised to the

reference gene, which encodes alpha tubulin-4 (F44F4.11). Following optimisation, the

concentration of the primers to amplify the TARDBP transcipts were reduced to 0.1

µM, to lessen primer dimerization. Amplification consisted of a denaturation step at

95 oC for 10 minutes, followed by 40 cycles of 3 seconds at∆ 95 oC and 30 seconds at

60 oC, with fluorescence monitored at the 60 oC stage of each cycle. Subsequent

quantification of results was performed using the ∆∆Ct method as previously

described in (Pfaffl, 2001)

Prior to quantitative PCR, primer specificity and PCR efficiency was validated by

generating melt curves and standard curves respectively.

Primer Primer sequence (5’-3’) Tm (oC)

CRNKL1 C.ELEGANS FORWARD GATGTTACTGGAAGCATGGA 57

CRNKL1 C.ELEGANS REVERSE AGGCATCATAGTTTCAACTCTC 58

HNRNPF C.ELEGANS FORWARD ATGAAGTTCAATGTCGTGGA 57

HNRNPF C.ELEGANS REVERSE GCTTTCGATTCCATTATTTCCG 58

HNRNPK C.ELEGANS FORWARD CGTTTGGAAGATAATTTCTCGG 57

HNRNPK C.ELEGANS REVERSE GATAAGAGCTCCAGCATGTG 58

TARDBP C.ELEGANS FORWARD AGCTGTTCAAGTCGATCCCA 58

TARDBP C.ELEGANS REVERSE GCGGGGCAACAATAACGAAG 58

ALPHA TUBULIN FORWARD CGCCTGGACTACAAGTTTGAC 60

ALPHA TUBULIN REVERSE GCCTCAGTGAACTCTCCCTCT 60

Table 3.2. Quantitative PCR primers. The sequences and annealing temperatures of primers used to assess the knockdown of splicing genes.

101

3.3 Results

3.31 HD worms do not exhibit the dye-filling defect

As mutant HTT was only expressed within a small subset of sensory neurons, an assay

exploiting the inherent nature of these neurons to take up certain dyes was used to

assess neurodegeneration. As neurons degenerate in response to the toxic insult of

mutant huntingtin, their capacity to take up the lipophilic membrane dye, DiD,

becomes impaired. As demonstrated in Faber et al (1999), this impairment is

apparent in around 13% of 8 day old worms, with no noted defect in young adult

worms (Faber et al., 1999).

Following substantial optimisation of this assay, it became clear that the results were

variable, in both N2 wild-type and HtnQ150 strains. Staining of the ASH, ASI, PHB and

PHA neurons in particular proved to be problematic, with worms taking up the dye

into both neuronal cell as well as muscle cell. Longer incubation periods coupled with

additional washes, failed to ameliorate this issue. While some worms appeared to take

up the dye, other failed to display the dye (see Figures 3.4a and 3.4b). This variability

remained consistent in the Htn150Q transgenic strain (see Figures 3.4c and 3.4d). In

many cases, the red florescent signal of the DiD dye, failed to co-localise with the GFP

specifically expressed in these neurons under the OSM10 promoter. Given that a dye-

filling defect is not observed in 3 day old worms, this is unlikely to reflect degeneration

of four classes of sensory neurons.

102

103

Figure 3.4 Confocal images of worms following incubation within DiD lipophilic dye. Both N2 wild type and HtnQ150 transgenic worms were incubated in the red fluorescent dye DiD, and visualised with the use of florescence microscopy. Sensory neurons with nerve endings exposed to the environment hold the capacity to take up lipophilic dyes such as DiD, an ability abrogated upon neuronal degeneration (A) A subset wild type worms appeared to take up the red lipophilic dye into several of the neurons that make up the nerve ring in the head. However, many of the animals failed to incorporate the compound into the ASH or ASI neurons (the positions of which are indicated by the white boxes). (B) Some worms displayed clear uptake of the dye into these specific sensory neurons (highlighted in the second panel) (C) Similar to the wild type strain, many of the HtnQ150 transgenic worms failed to display uptake of the dye into either of the OSM10::GFP expressing ASH or ASI neurons. (D) Some worms (left worm) displayed a red fluorescence signal overlapping with the green florescence found within these neurons (All images taken at 10x magnification).

3.32 RNAi knockdown of splicing-related suppressors does not modulate HD

phenotypes

To determine the effects of RNAi knockdown of splicing genes in HD worms sensory

neuron functioning (sensitivity) and general movement (bondy bends) was assessed in

8 day old worms maintained on RNAi expressing bacteria. Worms were also grown on

bacteria expressing an empty vector as a control. Initial characterisation of the two HD

strains revealed a subtle sensitivity defect in the HD worms relative to the N2 wild

type strain and a much more prominent decline in the frequency of their movement

(see Figure 3.5 and 3.6).

RNAi constructs were generated by Source Biosciences and although the majority of

their RNAi expressing bacteria appeared to grow well on LB agar, those expressing

EFTUD2 failed to form colonies. After several attempts to grow this particular RNAi

expressing bacterial strain, Source Biosciences admitted they were unable to generate

further constructs, and consequently we were unable to continue with this portion of

the work.

Upon genetic knockdown of each of the remaining splicing genes, it became apparent

that a reduction in their expression brought about little or no change in the behaviour

of both the wild type and transgenic worms. However, prolonged knockdown of

CRNKL proved to be lethal in each of the three C.elegans strains, with the vast majority

104

of worms failing to mature past four to five days. As such these worms were excluded

from behavioural assays.

Results from the sensitivity assay revealed a small drop in sensitivity in the HtnQ150

strain grown on bacteria containing an empty vector, compared to the corresponding

wild type worms, with the HtnQ150 pqe-1KO strain displaying an even greater defect.

Although these results failed to gain statistical significance following an ANOVA, the

trend between the three nematode strains remained consistent following knock down

of each of our splicing genes.

Similar to the results obtained from the sensitivity assay, the HtnQ150 strain exhibited

fewer body bends during a one minute period, indicative of a movement defect. This

defect was once again more conspicuous in the HtnQ150 pqe-1KO strain, a trend

which remained constant despite knock down of each of the splicing genes.

Quantitative PCR was employed to validate the knockdown of our suppressor genes

following prolonged growth of N2 wild type animals on RNAi expressing bacteria. RNA

was extracted from worms after three days of ageing. In each case, changes in gene

expression were normalised to alpha tubulin, and subsequently compared to control

worms fed on bacteria containing an empty vector.

Knockdown of CRNKL and HNRNPF resulted in a fairly modest decrease in gene

expression, approximately 10% and 20% respectively. Conversely knockdown of

HNRNPK and TARDBP caused an unexpected increase in expression of 35-fold and 3-

fold respectively (see Figure 3.7). After performing a paired Wilcoxon test to compare

the RNAi and EV treated worms, the results we not deemed statistically significant.

105

Figure 3.5. Knockdown of target genes failed to exacerbate motor defects in HD nematode strains. Initial characterisation revealed a subtle motor defect in empty vector treated HDQ150 worms relative to the N2 wild type animals, a defect which was significantly more pronounced in the HDQ150 pqe-1 KO worms. The worms treated with the knock down constructs displayed no significant change in motor behaviour in relation to their corresponding empty vector treated controls (SEM error bars, two way ANOVA, ***P<0.001, N=25-35).

106

Figure 3.6. Knockdown of target genes does not alter sensitivity defects in HD nematode strains. Characterisation of empty vector treated worms revealed a subtle, though statistically insignificant, sensitivity defect in both HDQ150 and HDQ150 pqe-1 KO worms relative to the N2 wild type animals. The worms treated with the knock down constructs displayed no significant changes in their sensitivity in relation to their corresponding empty vector treated controls (SEM error bars, two way ANOVA, N=25-35).

107

Figure 3.7. Knockdown constructs brought about a moderate decrease in expression of CRNKL and HNRNPF, and an increase in expression of HNRNPK and TARDBP. Following propagation of N2 worms on RNAi expressing bacteria for six days, eggs were isoloated and allowed to mature for a further three days. RNA was isolated from three day old worms, and analysed with the use of quantitative PCR. The data was analysed using the ∆∆CT method as described in (Pfaffl, 2001), in which results were initially normalised to an internal control (alpha tubulin), and then to control worms fed bacteria containing an empty vector. Knockdown of both CRNKL and HNRNPF resulted in a drop in expression, 9% and 22% decrease respectively. Surprisingly knockdown of HNRNPK and TARDBP resulted correspondingly in a 32 and 3 fold increase in expression (SEM error bars, N=3, paired Wilcoxon test was performed to compared RNAi treated and EV treated worms).

108

3.4 Discussion

Given the results obtained from the overexpression of our mRNA splicing suppressor

genes in mammalian cells, we hypothesized that knockdown of these genes would

enhance/modulate HD phenotypes in worms. We aimed to knockdown the five genes

which conveyed protection against mutant huntingtin in our PC12 neuronal cell line

(CRNKL, EFTUD2, HNRNPF, HNRNPK and TARDBP). The consequences of knockdown

were to be assessed via sensitivity of the worms to a physical stimulus, and the uptake

of a flourescent dye via the exposed sensory nerve endings, both of which had been

previously been shown to be impaired in HD worms (Faber et al., 1999).

Unfortunately the dye filling assay was hampered by a lack of reproducibility, and

although Faber et al noted the same problem in regards to ASI, PHA and PHB neurons,

they found that the ASH neurons displayed a more consistent uptake of the dye in

wild type worms. However in our study the ASH sensory neurons failed to

reproducibly exhibit this phenotype, despite considerable optimisation, and as such

the assay was dropped.

While conducting the initial characterisation of the sensitivity defects in HD strains, it

became apparent that the HD worms exhibited a more erratic movement pattern

compared to their wild type counterparts. This was confirmed using a simple body

bends assay, which revealed fewer sine wave movements in the HD Q150 worm

strain, a motor defect which was significantly exacerbated in worms lacking the PQE1

gene. Although this particular phenotype has not been observed previously, loss of the

diverse functions of the ASH neurons, which include avoidance of toxic repellents,

osmotic changes and mechanical stimuli, may account for this motor defect (Hilliard et

al., 2005). The direction of nematode movement is reliant on the chemosensory

function of the ASH neurons, coupled with the antagonistic actions of PHA and PHB

neurons. These neurons enable the worms to sense it’s environment and direct it’s

movement away from repellents and towards more favourable conditions (Hilliard et

al., 2002). Given that mutant HTT is expressed in all three of these neurons, ablation

of this interaction may result in disorientation.

109

Work conducted by Fujiwara et al also demonstrated behavioural changes in worms

stemming from defects in their sensory perception. In this study, mutation in the

f38g1.1 genes resulted in restricted movement, known as dwelling behaviour, and

fewer body-bends during forward locomotion. This gene encodes CHE-2, a component

of the intraflagellar transport complex (IFT), which modulates behavioural responses

to sensory inputs. This complex is crucial in the biogenesis of the sensory cilium and

therefore the functioning of the sensory neurons. Mutations in CHE-2 are thought to

result in an increase in cGMP-dependent protein kinase (PKG), a protein implicated in

eliciting dwelling behaviour in both Caenorhabditis elegans and Drosophilia

melanogaster. (Fujiwara et al., 2002, Fitzpatrick and Sokolowski, 2004). Furthermore

HTT has been shown to interact indirectly with components of the IFT, and this in turn

may bring about changes in social behaviour (Haycraft et al., 2003, Gervais et al.,

2002)

Behavioural changes have also been noted following loss of the ASH and ASL neurons,

inw which group-forming worms transform into solitary feeders, a change

accompanied by a drop in locomotion. Although the results of this study

demonstrated the need to disrupt both sets sensory of neurons to bring about a

drastic behavioural change, it is feasible disruption of ASH neurons alone may result in

a more a subtle transition in behaviour (de Bono et al., 2002). Both of these studies

demonstrate the role of sensory neurons in the regulation of nematode behaviour,

and by extension the effects of altering neuron functioning, be it via specific mutations

or the introduction of toxic insults such as mutant HTT.

I was not able to study RNAi knockdown of CRNKL, as worms treated with RNAi

targeting this gene, showed developmental abnormalities, failing to mature beyond 8

days, and as such were excluded from study. Previous work has found that this genes

plays a crucial role in not only mRNA splicing, but embryogenesis and cell cycle

progression, and as such knockout of the gene has been shown to be embryonically

lethal in mice (Zhang et al., 1991, Chung et al., 2002). Genetic knockdown of CRNKL in

various cell lines has failed to result in such lethality, though the effects on an

organism as a whole may be significantly greater than those on a single cell line

(Nybakken et al., 2005, Paulsen et al., 2009).

110

Following a period of knockdown, sensitivity of worms to touch stimuli was assessed

using an assay previously shown to highlight sensory defects in the subset of sensory

neurons expressing mutant huntingtin. We noted a subtle difference in the sensitivity

of wild type worms relative to transgenic strains, with HtnQ150 and HDQ150 PQE-1

KO strains displaying a 5% and 8% decline in sensitivity respectively. These figures are

in stark contrast to those obtained by Faber et al, who noted almost a 50% drop in

sensitivity in HtnQ150 worms compared to N2 controls (Faber et al., 1999). While the

published work made use of around 200 worms, we used 30 to 40 animals during our

experiments. The discrepancy in sensitivity may therefore reflect the low sample size

in our study and by extension the greater polarising effects of the mosaicism inherent

to these transgenic strains (previous work suggests only approximately 75% of ASH

neurons expressed the GFP reporter in transgenic lines).

Depletion of potential mutant HTT suppressors (HNRNPF, HNRNPK and TARDBP) failed

to significantly alter the sensitivity or the motor function of the transgenic worms.

Again, the low animal numbers coupled with the mosaicism displayed by the

transgenic worms, may have diluted any subtle changes in phenotype. On the other

hand, knockdown of our genes simply may not exacerbate the HD phenotype, despite

the beneficial effects brought about by overexpression. Such lack of symmetry has

previously been demonstrated by our group, in which a genome-wide overexpression

suppressor screen performed in yeast, revealed many genes that largely failed to

overlap with those derived from a previous genome-wide deletion screen (Willingham

et al., 2003, Mason et al., 2013).

The validation of RNAi knockdown of the candidate genes using of quantitative RT-PCR

produced some surprising results. Ingestion of RNAi expressing bacteria targeting

CRNKL and HNRNPF resulted in a modest reduction in gene expression, but RNAi

targeting HNRNPK and TARDBP lead to an increase in expression. In both of the latter

cases, the apparent increase in gene expression can be attributed to the amplification

of ingested dsRNA through the actions of RNA-directed RNA polymerase (RdRP) (Sijen

et al., 2001). Detection of this amplified RNA using qPCR can ordinarily be avoided

using primers designed to anneal outside of the cDNA region ligated into the feeding

vector. Unfortunately, given the size of the cDNA regions in relation to the rest of the

111

gene, it was not possible to design primers that not only fell outside of this region but

also spanned exons, a caveat essential to avoid the amplification of unspliced

immature mRNA. Analysis of protein levels following knockdown, via Western blotting

may be a suitable alternative to qPCR.

Ultimately the negative results obtained from the behavioural assays following

depletion of our suppressors may reflect the modest decrease in expression brought

about by the RNAi. Other groups using this ingestion method of knockdown have

reported expression reductions in excess of 50% (Agger et al., 2007, Hansen et al.,

2005). This may be attributed to degradation of the RNAi inducing agent, IPTG, or

possibly insufficient levels of plated bacteria. On the other hand, given the lethal

phenotype brought about by CRNKL1 depletion, the knockdown can be reasoned to

have been somewhat effective.

112

3.5 Future Work

Although the genetic knockdown of our putative suppressor genes yielded largely

negative results, these results could be reinforced by expanding the number of

animals studied. This would diminish the effects of the mosaicism inherent to the

transgenic worm strains. Past studies using C.elegans as a model organism have

utilised hundreds of animals, which allows for the identificaton subtle phenotypes.

Repeating our experiment with a greater number of animals may uncover changes in

sensitivity and body bends following depletion of our suppressors.

The results were further weakened by the sole use of the subjective sensitivity and

body bends assays. Other assays could therefore be used alongside these behavioural

tests. These may include the use of chemical repellent or attractant tests, in which

worms are scored according to their avoidance of a toxic chemical, such as heavy

metals, alkaloids or detergents, or their movement towards attractants such as

chloride ions (Bargmann and Horvitz, 1991, Hilliard et al., 2002). Such assays rely on

the worm’s capacity to utilise it chemosensory, rather than mechanosensory

functions. Given the roles of ASH and ASHI neurons in chemical avoidance and

chemotaxis respectively, as well as the greater OSM-10 expression in ASH neurons, a

chemical repellent test would be expected to illustrate the greater impairment of

sensory neuron function (Faber et al., 1999). However, the mosaic nature of GFP and

HTT expression in the sensory neurons within these transgenic worms could still

potentially skew the results.

One possible solution to the issue regarding the mosaicism exhibited by this particular

transgenic C.elegans strain would be the use of an alternative HD nematode model,

for instance the strain generated by Wang et al expresses mutant HTT in muscle cells.

This model displays significant motor defects due to the large number of mutant HTT

expressing cells, with worms expressing Q74 HTT fragment displaying around a 60%

reduction in body bends in relation to N2 WT animals (Wang et al., 2006). These

worms may therefore exhibit a much greater response to RNAi targeting compared to

worms expressing of mutant HTT in a subset of the sensory neurons (Wang et al.,

2006).

113

Our results could also be explored with the aid of other HD nematode models,

including the ID448 strain developed by Lejune et al (Lejeune et al., 2012). In addition

to expressing a 128Q HTT N-terminal fragment in the nose touch receptors, worms

also carries a mutation in the rrf-3 gene, encoding an RNA-directed RNA polymerase,

and thus rendering the animal hypersensitive to the effects of RNAi (Simmer et al.,

2003, Simmer et al., 2002, Lejeune et al., 2012). Such nematode strains not only

display a significant touch-nose defect and axonal degeneration relative to animals

expressing a 19Q fragment, but exhibit enhanced effects of knock-down that could be

measured by this robust phenotypic read-out (Lejeune et al., 2012).

114

Chapter 4

Testing new behavioural paradigms in HD models mice

4.1 Introduction

4.11 The use of mouse models in the study of HD

Our understanding of human diseases has been greatly enriched by the use of murine

models. Genetically altering animals to express disease-causing genes or treating mice

with compounds to recapitulate disease phenotypes, has provided substantial insight

into the mechanisms underlying pathogenesis. Such advances have been particularly

prominent in the study of genetic neurodegenerative disorders, such as HD and PD

(Levine et al., 2004, Menalled and Chesselet, 2002, Menalled, 2005).

Since identification of the HTT gene in 1993, murine models have been used

extensively in the study of HD, providing a means of validating work conducted in vitro

and in model organisms such as Drosophilia melanogaster and Caenorhabditis

elegans. Given that mice possess far fewer CAG repeats within their own endogenous

copies of the HTT gene, sporadic forms of the condition do not exist, and as such

genetic modification is required to reproduce a phenotype similar to that observed in

human forms of the disease. A multitude of genetic mouse models have been

generated which exhibit some of the characteristics synonymous with HD, though no

one model exhibits the full spectrum of symptoms (Menalled and Chesselet, 2002,

Rubinsztein, 2002).

Genetic mouse models of HD display a range of phenotypes, including motor defects

such as clasping, resting tremors, abnormal gait and hypokinesis, and

neuropathological symptoms including brain atrophy and the formation of

intranuclear inclusions (see Figure 4.1). These features of HD can be assessed with the

aid of behavioural and histological assays.

4.12 HD mouse models

Genetic HD murine models can be divided into three groups: knockout (KO); knock-in

(KI); and transgenic, all of which have inherent advantages and limitations. A knockout

model is defined as an animal in which a gene has been deleted or interrupted,

115

preventing transcription. While insertion of a fragment mutation or a whole human

gene in place of the endogenous mouse gene is referred to as a knock-in model. Such

genes are positioned within the natural genomic context, and are therefore regulated

by endogenous promoters. Transgenic mice are generated by the insertion of a

mutant gene or fragment into the mouse genome at an unknown position. As such,

the animal expresses both the mutant and endogenous genes, with the former being

regulated by an independent promoter (Menalled and Chesselet, 2002)

4.13 Knockout mouse models

Knockout mouse models of HD were generated shortly after identification of the HTT

gene, and revealed a crucial role for HTT in embryogenesis. Mice lacking functional

copies of the HTT gene exhibited post-implantation embryonic lethality, with

developmental defects most evident in the brain (Duyao et al., 1995). Although such

nullzygous mice have provided little insight into HD pathogenesis, the lethal effects

suggest that mutant HTT retains some of the functionality of the wild-type protein,

(White et al., 1997, Duyao et al., 1995). Heterozygous knockout mice mature to

adulthood, although the reduced expression of HTT in these mice causes motor and

cognitive defects similar to those displayed by HD patients. These phenotypes suggest

HTT plays an important role in the correct functioning of the basal ganglia, and that a

reduction in the levels of the functional protein contributes to HD pathology (Duyao et

al., 1995, Nasir et al., 1995b).

Conditional knockout mouse models provide a means of regulating the temporal

and/or spatial expression of mutant HTT, and by extension the study of this protein at

varying point during disease progression. Suppression of mutant HTT expression in the

forebrain 5 days after birth, with the aid of a Camk2a promoter, was shown to result

in neuronal degeneration, early mortality and male sterility, suggestive of a role in not

only neuron maturation but also spermatogenesis (Dragatsis et al., 2000). In contrast,

elimination of mutant HTT in transgenic mice between 18 and 34 weeks of age

reverses HD disease symptoms. Such mice, expressing a chimeric human/mouse exon

one fragment containing a 94 residue polyQ tract driven by a tetracycline regulated

promoter, showed a dramatic loss of neuronal HTT inclusions coupled with a reversal

116

of the clasping phenotype, relative to their “gene-on” counterparts. Such study has

demonstrated that the continued expression of mutant HTT is required to maintain

the disease state (Yamamoto et al., 2000).

4.14 Knock-in mouse models

Although knock-in models of HD provide a faithful representation of the human

disease, phenotypes are often quite subtle, absent or manifest in late adulthood, with

mice having a life span comparable to their of wild-type counterparts (Menalled and

Chesselet, 2002). For example the HdhQ72-80 knock-in mouse lines display no other

phenotypic change in response to an increased polyQ tract other than an elevated

level of aggression, and exhibited no discernable neurological dysfunctions compared

to age-matched wild-type littermates. This was attributed to a greater CAG mutational

threshold in mice, and prompted others to examine much longer repeat lengths

(Shelbourne et al., 1999).

The generation of knock-in mice expressing 94 CAG repeats revealed the presence of

nuclear “microaggregates” in striatal neurons at 6 months and nuclear inclusions at 18

months. Animals also displayed subtle behavioural defects at 2 months in the form of

increased rearing at night and repetitive movement, giving way to reduced locomotion

at 4 months. These polarising behavioural changes from hyperkinesia to hypokinesia

have been observed in human HD sufferers, adding credibility to the model while

illustrating the importance of establishing the correct metrics to uncover subtle

phenotypes (Menalled et al., 2002).

Analysis of the HdhQ111 mouse line revealed nuclear inclusions at 12 months, late-

onset striatal degeneration as well as very subtle motor defects, most notably gait, at

24 months (Wheeler et al., 2000, Wheeler et al., 2002). A further expansion in the

number of CAG repeats in the HdhQ150 mice has provided the most promising of the

HD knock-in models, though different strains can often produce varying results

(Menalled, 2005). Many of these HdhQ150 mouse models typically display significant

motor defects between 16 and 24 months when assessed using behavioural tests such

as the rotarod, balance beam and clasping assays. This decline in motor function is

often correlated with an increased level of striatal atrophy and consequently a

117

reduction in striatal volume (Heng et al., 2007, Lin et al., 2001). These models also

display striatal and hippocampal nuclear inclusions at approximately 6 months, which

are later detected in other regions of brain, including the cerebellum and cortex (Heng

et al., 2007, Sathasivam et al., 2010).

More recently, other metrics have been utilised to study the progression of the

disease in HdhQ150 mouse models. For example, the use of the 3-stage water maze to

assess memory and the pre-pulse inhibition (sensorimotor gating) test to examine the

ability to respond to a stimuli following exposure to a similar, but weaker stimuli, has

uncovered deficits at 4 and 6 months respectively. In both cases, defects become

apparent long before the onset of any motor impairment, and again illustrates the

need to ascertain the most suitable metric to assess the manifestation of the disease

(Brooks et al., 2012).

4.15 Transgenic mouse models

Some of the first genetically manipulated animal models of HD were the transgenic

mice, and are considered to be an attractive model for preclinical trials, given their

short lifespan and well defined, fulminant phenotype (Carter et al., 1999). The R6/2

line, generated shortly after the identification of the HTT gene, is the most widely

used of these models and expresses an N-terminal exon 1 fragment with CAG repeats

ranging from 144 to 150 units (Mangiarini et al., 1996). These mice survive for a period

of up to 5 months, during which they exhibit an aggressive and ultimately debilitating

form of HD, akin to the human disease. Similar to the knock-in mice, this transgenic

line displays a progressive decline in motor function, with onset typically between 8

and 15 weeks, significant brain shrinkage at 12 weeks (with no significant neuronal

loss) and the appearance of nuclear inclusions by as early as 4 weeks (Carter et al.,

1999, Stack et al., 2005, Mangiarini et al., 1996). Furthermore the R2/6 mice display

defects in a myriad of cellular processes synonymous with the human disease,

including transcription, mitochondrial function and vesicle trafficking (Ferrante, 2009).

Such vast changes in cellular functions have been highlighted by microarray data

obtained at varying time points during disease progression, including early decreases

in the expression of components making up the calcium and retinoid signalling

118

pathways involved in synaptic plasticity, neural development and learning (Lane and

Bailey, 2005, Luthi-Carter et al., 2000). The same screen also revealed an early drop in

the expression of enkephalin, a neuropeptide selectively expressed in the medium

spiny neurons of the striatum, the most vulnerable cell type in HD (Luthi-Carter et al.,

2000).

Other HD transgenic lines include the N171-82Q mouse model, which expresses the

first 171 amino acids of mutant HTT (exons 1 and 2) under the mouse prion promoter,

thus restricting expression to the neurons of the CNS. Analogous to the R2/6 line, this

model displays phenotypic changes including failure to gain weight from 2 months,

motor defects from 4 months and a progressive increase in nuclear inclusions within

the striatum, cortex and hippocampus at 4 months, before ultimately succumbing to

the disease at approximately 5 months (Schilling et al., 1999). Unlike other HD mouse

models, the N171-82Q exhibits a significant level of neuronal degeneration in the

striatum, as well as a much higher degree of gliosis compared to that seen in the R6/2

and knock-in mice. This model is therefore more reminiscent of the human disease

and suggestive of a role played by the amino terminal sequence in mediating gliosis

and certain apoptotic pathways (Yu et al., 2003). Although the N171-82Q mouse line is

studied less extensively than the R2/6 mice, its robust phenotypic readouts have made

it an appealing model for a number of drug screens (Zádori et al., 2011, Saydoff et al.,

2006).

While many of the transgenic HD lines express a truncated form of the human gene,

the use of yeast artificial chromosomes (YACs) as a means of integrating a large

amount of genomic information into the mouse genome, has enabled the generation

of YAC HD mouse models, the most promising of which being the YAC128 line. This

mouse expresses full length human HTT under the control of the HTT promoter and

other regulatory sequences, in a spatial and temporal manner akin to the endogenous

mouse own protein (Slow et al., 2003). YAC128 mice generally develop motor defects

at 3 months, with sequential nuclear inclusions and striatal and cortical neuron loss at

around 12 months. This mouse model is predominantly used to examine the

mechanisms underlying disease pathogenesis, given the slower disease progression

relative to other transgenic lines, coupled with a strikingly similarity to the human

119

disease. Such similarities include selective vulnerability of the medium spiny striatal

neurons as well as dysfunction in apoptotic and mitochondrial pathways (Van

Raamsdonk et al., 2007, Fernandes et al., 2007)

While the various mouse models of HD each have intrinsic advantages and limitations,

they have provided a means of examining the many cellular and pathological facets of

HD, and have greatly enhanced our understanding regarding manifestation and

development of this condition (Menalled and Chesselet, 2002).

Figure 4.1. Timeline of behavioural and pathological symptoms exhibited by HD mouse

models. Timeline to illustrate onset of behavioural phenotypes and pathological

symptoms of various transgenic and knock-in mouse models, including rota-rod and

clasping phenotypes, the appearance of intranuclear inclusion, and changes in

transcriptional profiles. Adapted from (Crook and Housman, 2011, Menalled, 2005,

Menalled and Chesselet, 2002)

120

4.16 Aims

Following the work conducted in cellular and nematode models of HD, I wished to

examine our most promising candidate genes in a murine context. Initially work was to

be carried out in the HDhQ150 knock-in mouse model on a pure C57BL/6J background, a

generous gift from Professor Gillian Bates. This mouse line had yet to be fully

characterised in this strain background and as such were studied with the use of a

number of behavioural assays. While the rota-rod assay is used routinely to study

disease progression in HD models, the results are often variable, and as such we

sought to explore other more sensitive behavioural readouts. Ultimately the aim of

this portion of the study was to identify robust phenotypes to be used in the testing of

our candidate genes, and determine whether overexpression in the HD mouse model

conveyed the same suppression of mutant HTT observed in the previous HD models.

121

4.2 Materials and methods

4.21 Material

4.211 Bacterial strain

Escherichia coli strain

The DH5α bacterial strain was used during the course of the study, with the genotype:

F– Φ80lacZΔM15 Δ(lacZYA-argF) U169 recA1 endA1 hsdR17 (rK–, mK+) phoA supE44

λ– thi-1 gyrA96 relA1.

4.212 Constructs

pLENTI-CamKII-RTN3-RVS-GFP-Bsd

The 8991bp lentiviral plasmid consisted of the RTN3 (reticulon 3) gene under the

neuronal CAMKII promoter, as well as a GFP-Blastidicin fusion dual reporter under the

RSV (rous sarcoma virus) promoter. The plasmid was kindly donated by Dr Diego

Peretti, MRC Toxicology Unit, Leicester, UK.

4.213 Mouse strains

HdhQ150 C57Bl/6N knockin mice

Thsee HD mice were bred on a pure C57BL/6N background and were a generous gift

from Prof. Gillian Bates, Kings College, London. Mice expressed a 150 unit CAG repeat

within exon one of the mouse Hdh gene. All studies were conducted on homozygous

and wild type mice, while heterozygous mice were used for breeding. Homozygous

and WT mice were weighed every two weeks.

Tg(HD82GLn)81Dbo/J transgenic mice

These transgenic mice lines were bred on a C57BL/6N C3H mixed background, and

were purchased from the Jackson laboratory, Maine, USA. The mouse line expressed a

171 amino acid N-terminal fragment of human HTT cDNA containing 82 CAG repeats.

Expression was regulated by the mouse prion promoter, thus restricting HTT to the

122

neurons of the CNS. Given the rapid onset of disease phenotypes, hemizygous

offspring were used during the course of the study, which were bred from WT female

mated with male hemizygous mice.

All mice were reared and experimented upon in accordance with the Home Office

approved animal (scientific procedure) act of 1986. Mice were caged in groups of

three to five and provided with food and water ad libitum. The internal environment

was maintained at approximately 21oC on a 12-hour light/dark cycle. Ultimately mice

were sacrificed by cervical dislocation.

PPL 80/2340

PIL 40/10030

3.22 Methods

3.221 Genotyping

DNA was isolated from ear tissue (taken from newly born pups) using the RED Extract-

N-amp kit (Sigma). The isolations were carried our using the protocol supplied with

the kit. Tissue was incubated in 25 µL of extract/ tissue prep mix (4:1 ratio) for 10

minutes at room temperature and incubated at 95 oC for a further 3 minutes. 20 µL of

neutralisation buffer was subsequently added to samples, which were then stored at 4

oC prior to use.

HdhQ150 C57Bl/6N knockin mice

Following isolation of genomic DNA, samples were amplified using primers designed

to anneal to regions flanking the CAG repeats within the murine Htt gene. PCR

reactions were carried out in a thermocycler (Applied Biosystems 2720 thermocycler).

Each 20 µL reaction consisted of 2 µL of tissue digest, 1 µL of each of the forward and

reverse HdQ150 primers (see Table 4.1) (primers were generated by Sigma-Aldrich

and diluted to 10 µM stocks, with a final working concentration 0.05 µM) and 16µl of a

reaction mix. This latter buffer was made up of 8 µL of 5 M Betaine (Sigma-Aldrich), 2

µL of Detloff buffer (150 mM Tris-HCl pH 8.8, 150 mM Tris-HCl pH 9.0, 160 mM

ammonium sulphate, 25 mM magnesium chloride, 1.5 mg/mL BSA and 0.07% β-

123

mercaptoethanol), 1 µL of 2 mM dNTPs (NEB), 0.2 µL Herculase Taq (Agilent) and 4.8

µl of sterile distilled water.

The reaction consisted of an initialization step of 5 minute at 95 oC, followed by 30

cycles of 30 seconds at 94 oC, 30 seconds at 58oC and 30 seconds at 72 oC. After

cycling, amplification ended with a final 5 minute hold at 72 oC.

Following amplification, 3 µL of orange G running dye (Sigma-Aldrich) was added to

the 20 µL PCR reaction mix and subsequently loaded onto a 2% agarose/TAE gel,

alongside a 100 bp DNA ladder (Thermoscientific). Samples derived from WT mice

displayed bands of 278 bp in size, while those originating from knock-in mice gave rise

to bands of 707 bp. Samples obtained from heterozygous mice displayed both bands.

Tg(HD82GLn)81Dbo/J transgenic mice

Genotyping of the transgenic strain was performed using the MyTaqTM Extract-PCR kit

(Bioline), and in line with supplier instructions. Tissue samples were incubated with 20

µL of buffer A, 20 µL of buffer B and 70 µL of sterile DNase-free water at 75oC for 5

minutes. Samples were incubated for a further 10 minutes, at 95oC and subsequently

centrifuged at 13,000 r.p.m for 1 minute. The resulting supernatant was then used for

the PCR reaction, which consisted of 1 µL of genomic DNA containing supernatant, 0.5

µl of each of the four primers (20 µM stock) (see table 4.1), 12.5 µL of Taq HS red mix

and 9.5 µL of sterile DNase-free water.

The amplification reaction was made up of an initial denaturation step of 95 oC for 3

minutes, followed by 30 cycles of a 95 oC hold for 35 seconds, 58 oC for 30 seconds and

72 oC for 45 seconds. After the cycling stage, stages were heated to 72 oC for 2

minutes.

3 µL of orange G dye was added to each of the samples, which were then loaded onto

a 1.5% agarose gel. Samples from hemizygous transgenic mice gave rise to band

equivalent to 250 bp as well as a positive control band at 200 bp, while those from WT

mice displayed only the latter band.

124

Genotyping Primer Sequence (5’-3’)

Tm (oc)

HDQ150 forward CCCATTCATTGCCTTGCTG 61

HDQ150 reverse GCGGCTGAGGGGGTGA 63

HD82 forward GTGGATACCCCCTCCCCCAGCCTAGACC 79

HD82 reverse GAACTTTCAGCTACCAAGAAAGACCGTGT 70

HD82 positive control forward GTCAGTCGAGTGCACAGTTT 60

HD82 positive control reverse CAATGTTGCTTGTCTGGTCG 62

4.222 Lentiviral construction

Several methods were employed to clone the most promising of candidate genes into

the pLENTI plasmid. This required the removal of the RTN3 gene and subsequent

insertion of either the murine EFTUD2, HNRNPF or HNRNPK cDNA. The pLENTI plasmid

was transformed into E.coli, incubated on selective media, and isolated via plasmid

mini prep (as previously described).

Inverse PCR

Given the close proximity of the CAMKI promoter relative to the RTN3 gene, it was not

possible to simply digest the latter out of the pLENTI construct. As such, an inverse

PCR method was initially employed to remove the RTN3 gene from the pLENTI vector

and subsequently introduce restriction sites (XhoI and PacI), to facilitate the

subsequent cloning of murine cDNA into the vector. Both primers also contained an

additional SpeI restriction site upstream of the XhoI/PacI site (see Table 4.2). These

SpeI sites were added to the primer sequences to allowing re-ligation of the plasmid

following amplification. The circular plasmid was then transformed into E.coli,

isolated, and digested with XhoI and PacI. Both the PCR and restriction digests were

performed as previously described, with an annealing temperature of 65 oC and an

elongation period of 8 minutes.

Table 4.1. List of genotyping primers and melting temperatures. Knock-in and

transgenic mice were genotyped with the use of several primer sets. Genotyping of

the HDQ150 mice required the use of primers flanking the CAG repeat tract, while the

genotyping of the transgenic line required two sets of primers, one pair annealing to

the prion promoter (HD82 forward) and the human HTT cDNA (HD82 reverse), and

second pair annealing to T-cell receptor delta gene used as a positive control.

125

Gibson Assembly

After some difficulty with the inverse PCR method, Gibson assembly was employed to

generate lentiviral constructs. This method is based on the assembly of multiple DNA

fragments in an order governed by overlapping regions and was performed in

accordance with supplier instructions (New England Biolabs Inc.).

Sequences for the CAMKII promoter and candidate genes were amplified from the

pLENTI plasmid or the DNA constructs generated in Chapter 2. Primers were designed

with the aid of the NEBuilder for Gibson assembly web tool (www.NEBGibson.com),

and consisted of a gene specific sequence and an overlap sequence of approximately

15-25 bp, homologous to the adjacent fragment (see Table 2).

The pLENTI construct was initially digested with BamHI and XbaI, releasing both the

CAMKI promoter and RTN3 gene. The digested plasmid was loaded onto a 0.8%

agarose gel, and the vector backbone isolated as described previously.

Amplification of the CAMKII promoter sequences made use of the same forward

primer (Primer 3, GA PLENTI-CAMKII FOR) and a gene specific reverse primer

containing an overlapping region homologous to the 5’ end of each candidate genes.

Amplification of the EFTUD2, HNRNPF and HNRNPK required the use of forward

primers consisting of an annealing sequence matching the 5’ end of the cDNA, as well

as an overlap region homologous to the 3’ end of the CAMKI promoter sequence. The

reverse primer contained sequences homologous to the 3’ end of each the cDNA

genes and an overlap region matching the pLENTI plasmid.

Amplification of fragments was performed as previously described. The CAMKII

promoter was amplified using an annealing temperature of 63 oC and an elongation

time of 30 seconds. Amplification of EFTUD2, HNRNPF and HNRNPK required

annealing temperatures and elongation times of: 60 oC and 3 minutes; 64 oC and 1

minute; 64 oC and 1 minute respectively.

Following amplification, PCR samples were digested with DpnI to remove methylated

template DNA. The reaction mix consisted of 8 µL of PCR product, 1 µL of NEBuffer 4

126

and 1 µl of DpnI (New England Biolabs Inc.). Samples were incubated at 37 oC for 30

minutes, and heated to 80 oC for 20 minutes to deactivating the enzyme.

Prior to assembly samples were analysed by agarose gel electrophoresis and the

concentration of each sample estimated by comparison to molecular weight markers

(Hyperladder I). The 20 µL assembly mixture consisted of 25 ng of vector backbone,

150 ng of the CAMKII promoter fragment, 50 ng of the cDNA fragments, 10 µl of 2X

Gibson assembly master mix, and deionised water making up the remainder of the

volume. Samples were incubated at 50 oC for 15 minutes and then stored on ice,

before being transformed into E.coli and grown on a selective LB media.

To confirm the correct integration of fragments into the pLENTI vector, plasmids from

positive colonies were isolated and digested with either NruI or PstI, which cut both

the cDNA and CAMKII promoter fragments. Plasmids containing the HNRNPF or

HNRNPK cDNA were expected to yield bands equating to 1676 bp and 7839 bp, or

1820 bp and 7839 bp when cut with NruI respectively. Constructs containing EFTUD2

were expected to produce bands of 2782 bp and 8404 bp when digested with PstI.

Colonies were further validated via sequencing with the use of primers stated on

Table 4.2.

127

Primer Sequence (5’-3’) Tm (oc)

pLENTI XHO1 AGCTTGATCACTCGAGTCTAGAGTCGACCATAGTGA

67

pLENTI PAC1 GATCACTAGTTTAATTAACTGCCCCCAGAACTAGGGGC

67

GA PLENTI-CAMKII FOR

AAAATTCAAAATTTTATCGGACTTGTGGACTAAGTTTGTTCG

61

GA HNRNPF-CAMKI REV

CAGCATCATCTGCCCCCAGAACTAGGG 66

GA CAMKII-HNRNPF FOR

GGGGGCAGATGATGCTGGGCCCTGAG 65

GA PLENTI-HNRNPF REV

GTCACTATGGTCGACTCTAATCATATCCGCCCATGC

61

GA HNRNPK-CAMKI REV

GTCTCCATCTGCCCCCAGAACTAGGG 66

GA CAMKII-HNRNPK FOR

GGGGGCAGATGGAGACCGAACAGCCA 64

GA PLENTI-HNRNPK REV

GTCACTATGGTCGACTTTAGAAAAACTTTCCAGAATACTGCTTC

61

GA EFTUD-CAMKII REV

AGTATCCATCTGCCCCCAGAACTAGGG 66

GA CAMKI-EFTUD FOR

GGGGGCAGATGGATACTGACTTGTATGATG 57

GA PLENTI-EFTUD REV

GTCACTATGGTCGACTTCACATGGGATAATTGAGC

57

pLENTI seq F GGTACAGTGCAGGGGAAAGAATAG 59

pLENTI seq R GTACAAGAAAGCTGAACGAGAAACG 58

pLENTI HNRNPF seq F

GAAAGCATGGGACACCGGTATATTG 59

pLENTI HNRNPK seq F

AGTGTTTCAGTCCCAGACAGCAGTG 62

Table 4.2. List of primers used cloning of EFTUD2, HNRNPF and HNRNPK into pLENTI

vector. Several methods were employed during the cloning of candidate genes into the

pLENTI plasmid, including inverse PCR with the use of primers 1 and 2. The Gibson

assembly technique was also used to generate the overexpression constructs. Each Gibson

assembly required the amplification of fragments and the addition of overlapping regions.

Each primer consisted of two of three regions, a plasmid specific region (in green), a

CAMKII promoter region (in red) or a gene specific region (in blue) (Primers 3-12).

Sequencing of the constructs was performed with the aid of primers 13 and 14, as well as

primers 15 and 16 for constructs containing the HNRNPF and HNRNPK gene inserts.

128

4.223 Behavioural tests

Rota-rod

Mice were tested on an accelerating rota-rod to assess the decline in motor

coordination during disease progression. All experiments were performed with the aid

of the Mouse Rota-Rod 47600 (Ugo Basille) as described in (Dunham and Miya, 1957,

Jones and Roberts, 1968). Mice were trained at a constant rotational speed of 4 r.p.m.

for 5 minutes, followed by a rest of 30 minutes. Animals subsequently underwent

three trials on the accelerating function, with the speed increasing from 4 to 40 r.p.m

over a 5 minute period. The latency to fall was recorded and mice allowed a 30 minute

recovery time between trials. Each mouse was tested on three consecutive days, and

the average of the 2nd and 3rd day trials plotted.

Burrowing

The innate burrowing behaviour of mice, associated with hippocampal functioning,

was assessed as previously described (Deacon, 2006). Mice were placed in cages with

a clear plastic tube, 20cm long and 6.8cm diameter, raised at one end by 0.3cm and

closed at the opposite end, filled with 140 g of dried food pellets. Tubes were placed

at the far left of the cage and animals were provided with water ad libitum as well as

nesting material. Cages were placed in incubators at a constant temperature of 21oC.

After 2 hours, the weight of the pellets displaced was calculated based on those

remaining in the tube. The tubes, containing the remainder of the pellets were

returned to the cages, and mice were subsequently allowed to burrow overnight, after

which a second reading was recorded.

Activity cage

General locomotion of mice was assessed using AM548 Standard (Dual Layer)

locomotor activity monitors (Linton instrumentation). The apparatus consisted of two

layers (46 cm by 25 cm), each with 24 infrared beams, positioned 4 cm apart. The

lower level monitored X and Y movement, while the upper level was used to measure

Z movement. All movement was monitored and configured by AMONLITE software

129

(MJS technology Limited), which provided data for general movement, speed and

rearing.

Mice were placed in the testing room, and allowed to acclimatize for one hour in

darkness. Animals were then placed into cages, with access to a water source, and

movement recorded every ten minutes over a one hour period in darkness.

130

4.3 Results

Characterisation of HdhQ150 C57Bl/6N knock-in mice

4.31 Knock-in mice exhibit a rota-rod defect at 14 months

Given the background of the HdhQ150 knock-in mouse model, it was necessary to

characterise the mice prior testing of candidate genetic modifiers in a murine context.

One of the most widely used assays to monitor the decline in motor coordination seen

in rodent models of neurodegenerative disease, is the rota-rod. Animals were initially

trained at a constant speed of 4 r.p.m for 5 minutes, and subsequently trialled on an

accelerating rod, from 4 r.p.m to 40 r.p.m over 5 minutes. Mice were assessed on the

apparatus three times a day on three consecutive days, with the average of the 2nd

and 3rd day’s trials being noted. Homozygous HdhQ150 and wild-type mice faired

comparatively well up until around 14 months, after which HdhQ150 mice displayed an

approximate 40 second decrease in the latency to fall to their wild-type counterparts

(see Figure 4. 2).

Figure 4.2 Homozygous HdhQ150 knock-in mice display a rota-rod defect at 14

months. Wild-type and homozygous HdhQ150 mice were tested on an accelerating

rota-rod, with the latter displaying a defect at 14 months. Mann-Whitney comparison

at each time point, P<0.05 (*) (N= 5-17).

131

4.32 Activity of HdhQ150 knock-in mice begins to deviate from wild-type after 12

months

The activity of mice was assessed via a number of metrics with the use of activity

cages. Animal were placed into the activity cage for one hour and monitored with the

aid of infrared sensors. One of the read-outs obtained from apparatus was the general

locomotion of mice, i.e. the number of beams breaks detected by the infrared sensors.

Mice aged 6 and 9 months displayed the same level of activity during the one hour

test period, with an initial hyperactivity, indicative of exploratory behaviour, followed

by a gradual decline in movement. However HdhQ150 mice aged 12 months, exhibited

a small but non-significant decline in early activity (see Figure 4.3).

The activity cage apparatus monitored several other metrics, including speed, back to

front traverses of the cage and rearing. Although there was no difference in speed and

back to front traverses between the two genotypes, 12 month old knock-in mice

exhibited a significant drop in rearing compared to WT animals (see Figure 4.4).

132

Figure 4.3 Homozygous HdhQ150 knock-in mice display defects in locomotion at 12

months. Mice were assessed in terms of their general locomotion, the number of

beam breaks were recorded every 10 minutes. Both 6 and 9 month homozygous mice

displayed similar activity profiles compared to age-matched WT, on the other hand, 12

month old knock-in mice appeared to display a much lower initial activity. Mann-

Whitney comparison at each time point (N= 3-6).

A. B.

C.

133

4.33 Homozygous HdhQ150 knock-in mice failed to gain weight during their life span

Both WT and homozygous HdhQ150 mice were weighed every two weeks from 9

months until around 18 months, after which the latter mice often displayed

debilitating HD-like symptoms (resting tremors, hypoactivity etc.) and were

consequently culled. Homozygous HdhQ150 mice weighed on average, approximately

30% less than their WT equivalents consistently during this 9 month period (see Figure

4.5). Furthermore there was no significant change in the weight of these homozygous

mice over this period.

Figure 4.4 Homozygous HdhQ150 knock-in mice display defects in rearing at 12 months.

In addition to general locomotion, the movement of mice was analysed in terms of

speed, cage traverses and rearing. Only the latter metric revealed any significant

difference between the two genotypes, which became apparent at 12 months. Mann-

Whitney comparison at each time point (N=3-6).

Figure 4.5 Homozygous HdhQ150 knock-in mice fail to gain weight. Mice were weighed

every two weeks from 9 to 18 months. Homozygous knock-in mice were found to

consistently weigh much less than WT mice over this period. Unpaired T-test of Linear

regression, P<0.0001 (****) (N= 6-16)

134

4.34 Homozygous HdhQ150 mice display a burrowing defect at 9 month

Given the late onset of defects in the rota-rod test, we sought explore other more

sensitive and robust behavioural read-outs to characterise our HdhQ150 mice. Mice

were assessed in terms of their innate burrowing behaviour, a prominent phenotype

in prion disease mouse models, and an assay used extensively within the Mallucci

group (Deacon et al., 2001). Mice were placed in cages containing a pellet-filled tube,

and allowed to burrow into the tube for 2 hours. Knock-in animals displayed a defect

in this behaviour from 9 months of age, displacing approximately 30% fewer pellets

than WT mice. Furthermore, this defect remained consistent until 15 months (see

Figure 4.6).

Figure 4.6 Homozygous HdhQ150 knock-in mice develop a burrowing defect at 9

month. Mice were assessed in terms of their burrowing phenotype, with knock-in

animal exhibiting a decline in this behaviour at 9 months relative to WT mice. Mann-

Whitney comparison at each time point, P<0.05 (*) (N= 5-17).

135

4.35 CAG tract in HdhQ150 displayed genomic instability

Initially the CAG repeat tract of the HTT gene found in our HD knock-in mice was

stably transmitted from generation to generation. Over the course of my studies, we

found the CAG tract became increasingly unstable (see Figure 4.6), in some cases

expanding by almost 150 bp. After attempts to stabilise, and restore the CAG repeat

length, we were forced to retire the knock-in mouse line, in lieu of the transgenic

HD82GLn model.

Figure 4.7. The CAG tract found in the HTT gene carried by the HdhQ150 mouse models

was found to be incredibly unstable. Upper gel. Initially the CAG tract of the HTT gene

exhibited intergenerational stability, with the knock-in HTT gene corresponding to a

single band at 707bp and the WT HTT gene of 278bp. Lower gel. After approximately 2

years, the CAG tract displayed increasing levels of instability, expanding by around

150bp.

136

4.4 Discussion

The aims of this portion of the project involved the initial characterisation of murine

models of HD, and the subsequent lentiviral-mediated overexpression of our most

promising candidate genes (see Chapter 2) via stereotaxic injection or intranasal (see

Chapter 5). We employed several behavioural methods during the initial

characterisation of the HdhQ150 knock-in model of HD, including the rota-rod test to

assess motor coordination, widely utilised to study the of degeneration of the basal

ganglia. We also monitored the burrowing behaviour of mice, a phenotype typically

observed in the study of prion disease and other neurological disordered stemming

from the degeneration of the hippocampus. Finally, we employed the open field

locomotor cages to monitor general movement of mice. Given the duration of each

test, all behavioural work was conducted on female mice, as male animals would have

required single housing to prevent sibling aggression.

The results derived from the rota-rod revealed defects in motor coordination fairly

late during disease progression, becoming apparent at around 14 months. This

particular test is especially sensitive to degeneration of the cerebellar and motor

neurons, and while cerebellar degeneration is largely absent in the HD, the latter

occurs late in the disease. Therefore, defects in motor coordination are often subtle in

knock-in models of HD, manifesting much later than other phenotypes (Kennedy et al.,

2003, Hickey and Chesselet, 2011, Hickey et al., 2008).

Although the rota-rod is widely used in the study of HD mouse models, the reliability

of the test is marred by a number of disadvantages and confounding variables. This

has led to the increasing use of more robust assays in place of, or alongside the rota-

rod. Potential problems with the rota-rod assay include the use of an accelerating

function and the need to conduct the study nine times over three consecutive days

(excluding the prerequisite training period), which may introduce the effects of fatigue

and learning to the results (Brooks and Dunnett, 2009). The design of the apparatus

itself may also present further problems, for example, grooves present on the rotating

cylinder can allow mice to grip the rod during the assessment, though this can be

remedied by simply covering these ridges. The most influential factor determining a

137

mouse’s capacity to complete the task is the weight and size of the mouse, with

smaller lighter mice being able to negotiate the rotations much easier than larger

mice. In our study HD knock-in mice were much smaller than their WT counterparts,

likely impacting greatly on our results and possibly masking earlier onset motor

dysfunction (Brooks and Dunnett, 2009, Hickey and Chesselet, 2011). However, the

weight of the homozygous knock-in mice remained unchanged between 8 and 15

months, suggesting the decline in motor co-ordination at 14 months is independent of

weight at this time point.

Given the disadvantages associated with the rota-rod test we also made use of other

behavioural assays to determine whether an early phenotype could be observed. The

results obtained from the burrowing assay, for instance, were somewhat surprising,

with homozygous knock-in mice displaying a clear behavioural defect at 9 months

relative to WT animals. Such innate behaviour has been demonstrated in the context

of prion disease, with prion infection of the hippocampus diminishing this phenotype

(Deacon et al., 2001, Mallucci et al., 2007). Here we demonstrate that burrowing is

impaired much earlier than any motor defect, indicative of a hippocampal deficit in

the early stages of the disease in the HD knock-in model (Deacon et al., 2002).

Although this particular phenotype is novel in the context of HD murine models, early

cognitive dysfunction has been noted in the R2/6 line, manifesting prior to any motor

impairments (Lione et al., 1999, Murphy et al., 2000).

Again, the size of the mouse may be a confounding variable in the interpretation of

our results, with smaller mice unable to displace the pellets given the greater energy

expenditure relative to their size. However, the constant weight of the homozygous

knock-in mice between 9 and 15 months, suggests the burrowing deficit was

independent of weight.

Following the burrowing phenotype, we assessed animal movement with the aid of

open field locomotor cages. This behavioural test holds a number of advantages, most

notably the large amount of data generated from a single one hour exercise. Previous

activity cage studies performed on R6/2 transgenic mice have demonstrated early

defects in rearing and climbing in both male and female mice (Zarringhalam et al.,

138

2012).Our preliminary data demonstrated possible lower levels of initial movement

and fewer acts of rearing in 12 month homozygous knock-in mice compared to age-

matched WT animals, suggestive of reduced exploratory behaviour. Both metrics have

been documented previously in other HD models (Bolivar et al., 2004, Steele et al.,

2007). Unfortunately the small number of mice used in these assay prevented us from

drawing firm conclusions from the data.

This portion of the study was suspended due to intergenerational expansion of the

CAG repeat length. Such mutational instability is fairly common in HD knock-in mice,

with expansion of the CAG tracts length typically being attributed to male

transmission of the mutation, while contraction is observed during female

transmission (Wheeler et al., 1999, Shelbourne et al., 1999). We attempted to

maintain and possibly restore the CAG repeat length through the breeding of

homozygous females expressing the correct mutation with WT animals. The resulting

heterozygous mice were mated to produce further homozygous mice, which

unfortunately also displayed a much larger mutation. Eventually the knock-in mouse

line was retired and substituted with the transgenic HD82Gln line.

The instability associated with the CAG repeat has been linked to the mismatch DNA

repair genes, MSH3 and MSH6. While the former gene is required for the increase in

CAG length, MSH6 prevents contraction from generation to generation, though the

mechanisms surrounding these effects are unclear (Dragileva et al.,

2009).Polymorphisms in the MSH3 gene present in different mouse backgrounds have

been shown to govern the stability of the CAG repeats. For example, animals of a

C57BL/6J background contain polymorphisms in the MSH3 coding region that lead to

much higher levels of the MSH3 protein compared to mice of a BALB/cByJ background

(Tomé et al., 2013). Given that our knock-in mice line was bred on a pure C57BL/6J

background, the changes in CAG repeat length between generations may be

attributed to the expression of higher levels of this DNA repair protein. Although the

knock-in model was unsuitable for our work, it may be of some use in the study of

intergenerational instability.

139

After retiring the HdhQ150 mouse line, we sought to conduct similar behavioural

characterisation using the HD82Gln transgenic model. However, such work was

delayed initially due to re-derivation issues following transfer off all in vivo work to the

University of Leicester’s Central Research Facility. This work was further postponed

due to breeding of transgenic mice with wild-type mice infected with Klebsiella.

However after almost 16 months, behavioural work using this transgenic model was

commenced in June 2014, and although the preliminary results are promising, they

have not been included in this report.

Ultimately the aim of the research project was to overexpress several of our candidate

genes in the HD mouse model. We intended to clone three of our most promising

gene hits, EFTUD2, HNRNPF and HNRNPK into the pLENTI vector, which was to be

packaged into lentiviruses and delivered into our murine models via stereotaxic

injection or intranasal delivery (see Chapter 5). The pLENTI construct (kindly donated

by Dr Diego Peretti) contained the RTN3 gene under the regulation of the CAMKI

promoter. Given the close proximity of the promoter and gene, it was not possible to

simply digest the latter from the plasmid and insert our own genes. Consequently we

attempted an inverse PCR method to amplify the plasmid backbone along with the

promoter sequence, using primers designed to attach additional restriction sites to

the ends of the resulting fragment, facilitating subsequent cloning. However after

several attempts to amplify such a large fragment (approximately 8 kb), including the

use of gradient PCRs and alternative polymerases, we decided to use Gibson assembly

to generate the constructs.

Using the Gibson Assembly method, we amplified our genes of interest and the

CAMKII promoter, generating fragments with overlapping regions. Following

restriction digestion of the pLENTI plasmid to remove the RTN3 gene and adjoining

promoter, these fragments were assembled into the digested plasmid. Assembly

mixes were subsequently transformed into E. coli and plated on selective media. Our

first attempt resulted in a lack of colonies, while our second attempt yielded several

colonies, which contained the intact pLENTI construct containing the RTN3 gene. Due

to time constraints of the project, further attempts to clone our candidate genes into

the pLENTI plasmid could not be made.

140

In conclusion, our characterisation data suggests that the burrowing assay may be a

robust and simple test to determine early behavioural dysfunction. This defect was

shown to become apparent almost half a year prior to the any significant rota-rod

impairment.

4.6 Future work

As stated previously, the ultimate goal of this portion of the study was the lentiviral-

mediated overexpression of our most promising candidate genes in our HD murine

models. Given the time constraints of the project, we were unable to complete the

cloning of EFTUD2, HNRNPF and HNRNPK into the pLENTi plasmid. Were time less of a

factor, we would endeavour to generate our overexpression lentiviral constructs,

either via the Gibson assembly method or another cloning method. Such constructs

would subsequently be administered into the transgenic mice via stereotactic

injection or intranasal delivery (see Chapter 5)

Finally with additional time, we could complete the initial characterisation of the

transgenic mouse model, and determine whether these mice exhibit the same

burrowing defect as seen in our knock-in mouse line.

141

Chapter 5

Validating intranasal delivery in prion infected mice

5.1 Introduction

As mentioned in Chapter 4, the ultimate goal of this project was the identification and

validation of genes found to suppress mutant HTT in yeast. The most promising

candidate genes were to be overexpressed in our HD murine models using a lentiviral

system. Although our group have previously made use of stereotaxis as a means of

delivering such viruses to the CNS, we have sought to explore and optimise alternative

routes, including intranasal delivery.

5.11 Delivery to the CNS

As highlighted in Chapter 3, RNAi technology has become of great importance in the

study of genetics. When utilising RNAi in multicellular organisms, the route by which

constructs are administered can dictate the efficacy of gene expression changes. In C.

elegans, genetic knockdown is achieved through feeding, leading to expression

changes comparable to that brought about via gonadal microinjection. This route of

delivery is not feasible in a murine model, and indeed humans, due to degradation of

constructs in the gastrointestinal tract (Timmons and Fire, 1998, Kamath et al., 2001).

Unfortunately the non-invasive delivery of plasmids, as well as drugs and other

therapeutics compounds, to the CNS of mice and humans is often met with a number

of obstacles, including the blood brain barrier (BBB), the blood cerebrospinal fluid

barrier (BCSF), systemic modification, targeting of specific neurons, and the possibility

of inflammatory responses (Costantini et al., 2000, Alam et al., 2010).

Many non–invasive methods of delivery have been optimised to negotiate these

obstacles, including chemical modification (such as the use of prodrugs), biological

methods (such as antibody conjugation), and carriers systems (for example, the

encapsulation of therapeutic compound in a lipid vesicle). Although these methods

have been shown to be effective in the targeted delivery of therapeutics, they each

have inherent limitations that prevent their widespread use, and consequently

invasive delivery techniques are still often used in scientific research. Such methods

142

include the direct injection of compounds intracranially, intracerebral implants or the

disruption of the BBB, all of which are rarely applied in humans (Pathan et al., 2009,

Alam et al., 2010).

5.12 Intranasal delivery

Several non-invasive alternative routes to oral administration have yielded promising

results, one of which is intranasal delivery. This method was first investigated as a

potential drug delivery pathway almost two decades ago, and exploits the presences

of olfactory and trigeminal nerves fibres in the nasal epithelium to rapidly transport

compounds to the CNS (Figure 5.1) (Thorne et al., 1995, Lochhead and Thorne, 2012).

Studies have shown that intranasal delivery is an effective route of administration for

drugs, peptides and nucleic acids in both animal models and humans, and as such this

technique has been successfully used in clinical trials (Lochhead and Thorne, 2012).

The intranasal delivery of insulin to diabetic patients was once viewed as a promising

alternative to the cumbersome and inconvenient daily bouts of subcutaneous

injections. (Dhuria et al., 2010). Although both routes of administration were equally

effective, the intranasal route often produced several side effects including nasal

irritation and destruction of the nasal mucosa, and was therefore abandoned as a

means of treatment for diabetes sufferers (Owens et al., 2003). More recently,

intranasally administered insulin has been shown to improve memory and mood in

healthy individuals and patients with Alzheimer’s disease (AD) (Benedict et al., 2004,

Reger et al., 2008). This increase in cognition was mirrored by a rapid rise in insulin

levels in the brain and the periphery, and although unclear, authors postulated this

increase may have reversed abnormalities in insulin and insulin-like growth factors 1

and 2 signalling synonymous with AD (Steen et al., 2005, Reger et al., 2008). These

latter studies have highlighted the use of intranasal delivery as a means of rapidly

targeting therapeutics to the brain for the treatment of neurological disorders.

143

Upon inhalation, compounds must penetrate several barriers to reach to the CNS, the

first of which is the epithelial layer lining the nasal passage (see Figure 5.2).

Compounds can traverse this initial barrier in several ways, including transcytosis

through the epithelial cells and paracellular diffusion. The latter of these processes is

hindered by the presence of tight junctions, though transient gaps formed between

adjacent cells, as a result of the continuous turnover of the epithelial cells, can

facilitate paracellular diffusion. Both pathways lead to the transport of substances to

the lamina propria, and subsequent absorption into either the blood vessels (and

systemic circulation) and lymph nodes or diffusion into the olfactory and trigeminal

nerve bundles(Lochhead and Thorne, 2012, Mathison et al., 1998, Kristens.K and

Olsson, 1971). Substances can also enter the olfactory and trigeminal nerve fibres

found interspersed between nasal epithelial cells, via endocytosis (absorptive, fluid

phase and receptor-mediated), resulting in retrograde transport to the olfactory bulbs

and brainstem respectively (Thorne et al., 1995, Lochhead and Thorne, 2012, Baker

and Spencer, 1986).

Figure 5.1. Intranasal Delivery. Following inhalation, substances including drugs, peptides and nucleic acids are capable of traversing the nasal mucosa. Substances can then travel along the olfactory or trigeminal nerve bundles to reach the olfactory bulbs and brainstem respectively, and diffuse throughout the brain. Adapted from (Hashizume et al., 2008)

144

Compounds that are able to cross the nasal epithelium, and avoid absorption into the

general circulation or the lymphatic system can enter the CNS via several potential

routes. The first of these pathways is intracellular transport within the olfactory and

trigeminal nerves, which has been demonstrated with the use of tracers such as

horseradish peroxidase conjugated wheat germ agglutinin (WGA-HRP) (Broadwell and

Balin, 1985, Anton and Peppel, 1991). This form of transport has a speed ranging from

130mm/day to 36mm/day and is largely governed by length and diameter of the axon

(Buchner et al., 1987).

Other routes of entry into the brain include extracellular diffusion along

compartments associated with the olfactory and trigeminal pathways, such as the

connective tissue sheath or via small perineural spaces maintained by tight junction

proteins for nerve fibre regrowth (Jansson and Bjork, 2002, Li et al., 2005). This form

of passive diffusion can lead to transport rates ranging from approximately 67mm/day

to 4mm/day and is largely dictated by the diffusion coefficient of the compound

(Thorne et al., 2004). Substances capable of diffusing into the connective tissue may

subsequently pass into small fluid filled spaces, such as the perineural and

perivascular spaces, running parallel to the olfactory and trigeminal nerves. Whilst

contained within these spaces, compounds are prone to convective bulk flow, which

can rapidly bring them into close proximity to the brain, at a rate of approximately

200μm/min (Bradbury et al., 1981, Ichimura et al., 1991). For instance the perineural

spaces surrounding the olfactory nerve passes through the cribiform plate, draining

into the cerebrospinal fluid (CSF) surrounding the brain. Given the speed at which

many substances administered intranasally are detected in the brain, it believed that

bullk flow is the predominant mode of transport for many compounds administered

via the nasal route (Lochhead and Thorne, 2012, Dhuria et al., 2010).

Upon reaching the olfactory bulbs and brainstem, compounds may spread throughout

the brain intracellularly from the peripheral olfactory and trigeminal neurons to the

second order neurons located within the brain, or travel extracellularly utilising the

bulk flow system within the cerebral perivascular and interstitial spaces(Dhuria et al.,

2010, Lochhead and Thorne, 2012). Recently the rostral migratory stream, the route

by which neuroblasts travel from the subventricular region of the brain to the

145

olfactory bulbs, has also been linked to the retrograde distribution of compounds

entering the olfactory bulbs. Disruption of this stream was found to reduce the levels

of intranasally administered calcitonin and erythropoietin in the brain (Scranton et al.,

2011).

Figure 5.2. Transport of intranasally administered compounds from the nasal passage to the olfactory and trigeminal nerve bundles. Compounds entering the nasal passage can traverse the nasal mucosa, via: 1. Intracellular transport within the olfactory and trigeminal nerve fibres; 2. Transcellular transport through the epithelial cells; 3. Paracellular transport between adjacent epithelial cells. Upon passing through this first barrier, compounds may reach the brain via: 4. Diffusion along the compartments lining the olfactory and trigeminal pathways; 5. Axonal transport within the olfactory and trigeminal neurons; 6. Bulk flow within the perivascular and perineural spaces. Adapted from (Lochhead and Thorne, 2012).

146

5.13 Factors affecting intranasal delivery

A number of factors must be taken into consideration, when administering

compounds intranasally, the first of which being the technique or application method.

Many studies utilising the nasal route to convey compounds to the brain have simply

used a pipette to place a small droplet of liquid onto the entrance of the nasal passage

of anesthetised mice or rats, allowing the sedated animal to inhale the solution. This is

generally performed over a 20 minute period, placing droplets onto alternate nostrils

and allowing around 2 minutes between doses for the liquid to penetrate the nasal

mucosa (Dhuria et al., 2010). Using this method has an inherent limitation, the

olfactory mucosa is a small region located at the top of the nasal cavity, and can be

difficult to reach simply by applying a small volume of liquid to the nostril. Head

position can therefore dictate the efficacy of this technique. Intranasal studies

performed on rats has demonstrated the importance of head position. Rodents placed

in the supine position, either at 70o or 90o, lead to a greater intranasal delivery of

steroid hormones into the CSF, compared to that achieved with a 90o upright position.

Although the supine position at 70o and 90o were both shown to be lead to transport

to the CSF, several studies have used a 0o position to avoid build-up of compounds in

the oesophagus and trachea (van den Berg et al., 2002).

Another factor influencing the efficiency of intranasal delivery is the volume of the

compound administered relative to the capacity of the nasal cavity. For example, the

size of an adult mouse’s nasal cavity is approximately 0.032 cm3, with the ideal volume

for delivery of compounds being between 3 and 4µL. High volumes of compound can

accumulate in the nasopharynx, and cause respiratory difficulty, while lower volumes

may be insufficient to reach the brain and elicit the desired effect (Dhuria et al., 2010,

Gross et al., 1982).

Given the role of the nasal mucosa in the defence against inhaled toxins, intranasally

administered compounds must overcome the body’s protective systems, including the

mucocilliary clearance mechanisms, efflux transport proteins, and the presence of

tight junctions. All of these protective mechanisms can be surmounted by altering the

formulation of the intranasally administered mixture (Pires et al., 2009).

147

The mucocilllary clearance (MCC) system significantly hinders the transport of

intranasally delivered compounds by reducing their residence time in the nasal

mucosa. The MCC consists of a mucus layer covering the nasal mucosa, which traps

foreign agents and subsequently transports them along the nasal pharynx to the

gastrointestinal tract with the aid of cilia. The action of the MCC is dependent on the

abundance and movement frequency of the cilia and the viscoelasticity of the mucus.

A greater number of cilia coupled with a viscous mucus lining results in a greater

clearance of foreign substances (Pires et al., 2009, Dhuria et al., 2010). The actions of

the MCC can be delayed with the use of mucoadhesives, which can increase the

contact time of administered compounds with the nasal epithelia, allowing for greater

levels of diffusion across this first membrane barrier. One such mucoadhesive is

chitosan, a cationic polysaccharide that forms electrostatic interactions with the

negatively charged epithelial cells, allowing compounds administered intranasally to

adhere to the nasal epithelial lining for a greater period of time. Chitosan is also

believed to disrupt the tight junctions in place between adjacent epithelial cells, thus

promoting paracellular transport (Yu et al., 2004, Vaka et al., 2009). The addition of

chitosan has been shown to particularly effective in the transport of naked plasmid

DNA. Given its positive charge, chitosan induces a conformation change in the plasmid

structure, forcing the DNA to adopt a compact structure that facilitates absorption

through the mucosa (MacLaughlin et al., 1998, Iqbal et al., 2003).

Other approaches to counteract the actions of the MCC, includes the use of

engineered nanoparticles conjugated to ligands with the capacity to bind to specific

cell types. Formulations containing nanoparticles attached to Ulex europeus

agglutinin I (UEA I), significantly increased delivery to the brain (Gao et al., 2007). This

ligand group binds to receptors found in the olfactory epithelium and facilitate

penetration of this initial barrier, therefore enhancing delivery to the CNS via the

olfactory route (Gao et al., 2007). Nanoparticles attached to wheat germ agglutinin

(WGA), have been found to bind receptors found throughout the nasal epithelia,

resulting in enhanced transport of compounds to both the CNS and systemic

circulation (Gao et al., 2006a).

148

The limited permeability of the nasal mucosa can greatly hamper the transport of

intranasally administered compounds to the CNS. One strategy to overcome this

barrier is the use of non-ionic surfactants that transiently disrupt the nasal epithelial

membrane, putatively via membrane fluidization, increasing transcellular and possibly

paracellular transport. For example alkylglucoside tetradecylmaltoside (TDM),

increases nasal uptake of both low molecular weight heparin (LMWH) and

fluoresceine-isothiocyanate labelled insulin (Arnold et al., 2002, Arnold et al., 2004).

The latter experiment noted a greater number of internalised vesicles upon addition

of TDM, and attributed the enhanced absorption of insulin to a higher rate of

endocytosis, leading to both a greater uptake and the removal of cilia from the apical

membrane(Arnold et al., 2004).

Blood flow surrounding the nasal passage purportedly influences the uptake of

intranasally administered compounds. While blood flow is crucial in maintaining the

diffusion gradient at the nasal epithelium and thus facilitating the diffusion of these

compounds across this initial barrier, a strong blood flow can result in increased level

of systemic clearance (Dhuria et al., 2010). Treatment with vasoconstrictors such as

phenylephrine reduces absorption into the blood, and increases transport via the

olfactory nerve pathway (Dhuria et al., 2009).

Although the intranasal route has several advantages over systemic delivery, the vast

majority of compounds administered by this method travels to the brain via

extracellular pathways, so the BBB can still poses a major obstacle (Graff and Pollack,

2003). The P-glycoprotein efflux system found within the BBB attenuates the

movement of a wide range of substances such as drugs, antibiotics and antivirals.

Studies have shown that intranasal pre-treatment with P-glycoprotein inhibitors, such

as rifampin, can greatly enhance the brain uptake (Padowski and Pollack, 2010).

5.14 Gene therapy and Neurodegenerative diseases

The use of intranasal delivery as a means of inducing changes in gene expression

within the brain has been studied by few groups. Despite the scarcity of research,

these recent studies have highlighted the potential use of this technique for gene

therapy of CNS disorders (Kim et al., 2009a, Renner et al., 2012).

149

Gene therapy is defined as the delivery of nucleic acid to treat or prevent disease, and

has often been touted as a potential treatment for inherited disorders (Nanou and

Azzouz, 2009). Although preliminary work suggests great promise, progress has been

hampered by a number of issues, including the stability and expression duration of

transgenes, cell specific targeting, possible toxicity and the delivery route. Many of

these matters have been addressed by utilising the appropriate vector system and

altering such vectors to minimise unintentional side effects (Ralph et al., 2006). Of the

various vehicles used to convey nucleic acids to target tissue, the lentiviral system, a

form of retrovirus, has been used extensively. Advantage of this system over others,

include the large capacity for cloning, targeting of both dividing and non-dividing cells

as well as stable long-term expression (Azzouz, 2006). Lentiviral vectors can also be

pseudotyped to a variety of envelope proteins, thus manipulating the tissue specificity

or tropism of the vector (Nanou and Azzouz, 2009).

Work is underway to determine the therapeutic implications of gene therapy in the

treatment of neurodegenerative diseases, with several lines of research currently

entering clinical human trials. One particularly promising preclinical study was that

carried out by Azzouz et al, who made use of a lentiviral vector based on the equine

infectious anemia virus to treat a toxin-induced rat model of PD, a condition

predominantly affecting the dopaminergic neurons found within the striatum. The

vector expressed tyrosine hydroxylase, aromatic amino acid dopa decarboxylase and

GTP cyclohydrolase, enzymes involved in dopamine synthesis, in a single transcript.

Following stereotaxic injection, the construct was found to be stably expressed in

striatal tissue, leading to increased dopamine synthesis and reversal of behavioural

defects (Azzouz et al., 2002).

The potential use of gene therapy in the treatment of HD has been illustrated in

several studies. For example, de Almeida et al found overexpression of ciliary

neurotrophic factor (CNTF), a nerve growth factor involved in inflammation, to be

neuroprotective in a toxin induced HD rat model. Animals injected stereotaxically with

lentiviral construct displayed improvements in behavioural defects coupled with

significantly reduced striatal lesions following administration of quinolinic acid, which

produced HD-like symptoms (de Almeida et al., 2001). Further studies conducted in

150

the transgenic YAC72 HD mouse model revealed improvements in behaviour, as well

as a stable expression of CNTF almost a year after stereotaxic injection of the lentiviral

construct (Zala et al., 2004).

The most striking example to illustrate the potential of gene therapy comes from

studies of amyotrophic lateral sclerosis (ALS), a neurodegenerative condition

characterised by the loss of motor neurons of the spinal cord, brainstem and motor

cortex. Genetic forms of the disease caused by mutations in the gene encoding Cu/Zn

superoxide dismutase I (SOD), an enzyme that catalyzes the conversion of superoxide

radicals to oxygen and hydrogen peroxide. Simultaneous research conducted by both

Ralph et al and Raoul et al, found that RNAi knock down of SOD via lentiviral-mediated

injections into muscle and spinal tissue of ALS mouse models respectively,

substantially delayed onset of the condition, with the former group reporting an 80%

increase in survival (Ralph et al., 2005, Raoul et al., 2005).

As of 2013 over 1800 gene therapy clinical trials were underway or had been

completed. The number of studies in phase II/III of clinical trials has steadily increased

by over 25% since 2004, with two studies relating to haemophilia D and lipoprotein

lipase deficiency, in phase IV. Although the use of gene therapy as a means of treating

or preventing neurodegenerative diseases is in its infancy, the progress made to date,

is promising (Ginn et al., 2013).

5.15 Aims

Many gene therapy studies have involved the injection of viral vector into target

tissue, which although effective, is inappropriate for human treatment. The aims of

this portion of the project centre on the validation of intranasal delivery as a means of

conveying lentiviral constructs to the brain, with the ultimate goal of delivery of

constructs to overexpress candidate HD modifiers. While my own research focuses on

HD, this work was conducted in the context of prion disease, using a well-established

mouse model as a paradigm for neurodegeneration.

151

Prion disease, or transmissible spongiform encephalopathy, is a rare

neurodegenerative condition characterised by the formation of large vacuoles, or

spongiforms within the brain, leading to neuronal loss and a mirrored decline in

mental and physical capabilities (see Table 5.1) (Wadsworth and Collinge, 2011). The

disease is attributed to conformational changes in the prion protein (PrP), which arise

sporadically, accounting for ~85% of human cases, or through autosomal dominant

mutations in the PRNP gene. Such changes gives rise to the infectious agent, PrPSc, a

protease-resistant protein which holds the capacity to convert the host-encoded PrPC

into toxic prion species, and leading to the formation of aggregates inclusions

(Wadsworth and Collinge, 2011, Prusiner, 1991). The mechanisms by which prions are

able to propagate are contentious, though the continued presence of the

endogeneous PrPC species appears to be crucial, with prion-null mice displaying a

resistant to prion infection (Büeler et al., 1993). This has been further illustrated

following generation of a transgenic mouse line expressing PrpC under control of the

Cre-lox system. Depletion of the neuronal prion protein following inoculation with

infected brain homogenate, was found to halt disease progression and substantially

extend lifespan relative to those mice with a sustained expression of the protein

(Mallucci et al., 2003). Such work provided the foundation for further therapeutic

research using lentiviral-mediated RNAi constructs against PRNP. Single stereotaxic

treatment with RNAi constructs into the hippocampus of infected mice was found to

reduce spongiosis and neuronal loss and extended lifespan by 24% compared to

untreated mice (White et al., 2008a).

152

Symptom type Description

Behavioural Aggression; anxiety; depression; loss of inhibition and judgement

Communication Slurred speech; repetition; loss of reading and writing ability

Memory Loss of both short and long term memory; loss of recognition

Cognition Reduction in intellect; lack of awareness

Movement Ataxia; loss of gait; jerky movements; tremors; rigidity

Perception Double vision; cortical blindness; hallucinations;

Seizures Late onset seizures

Table 5.1. Clinical signs of prion disease. Adapted from

http://www.prion.ucl.ac.uk/clinic-services/information/signs-and-symptoms/

Here I aimed to the conduct similar experiments to determine whether the intranasal

delivery may prove to be a viable alternative method to stereotaxic injection, using

the prion disease as a paradigm for neurodegeneration. Ultimately the technique was

to be employed in the lentiviral-mediated overexpression of our suppressor in HD

mouse models.

153

5.2 Material and Methods

5.21 Materials

5.211 Lentiviral constructs

MW1/EV

Lentiviral constructs were generated as described in (White et al 2008). The pLL3.7

lentiviral plasmid consisted of an siRNA sequence (21 nucleotide inverse repeat and 9

nucleotide loop) against PRNP, downstream of the U6 promoter, as well as the EGFP

encoding gene under the control of a CMV promoter. A construct, lacking the siRNA

sequence was also used during this study as a control. Virus production was carried

out by GenTarget Inc. The company provided packaged virus particles, pseudotyped

with VSV-G (vesicular stomatitis virus glycoprotein) proteins, at a titer of 107 IFUs

(infection function units).

5.212 Mouse models

C57BL/6N mice

Wild-type mice were purchased from Charles River Laboratories, and reared and

experimented upon in accordance with the Home Office approved animal (scientific

procedure) act of 1986. Mice were caged in groups of three to five and provided with

food and water ad libitum. The internal environment was maintained at approximately

21 oC on a 12-hour light/dark cycle. Mice were sacrificed by cervical dislocation.

Tg37 homozygous/hemizygous mice

Tg37 mice over-expressing the Prnp gene were used during the course of the study

(Mallucci, Ratte et al., 2002). These mice had been generated via the microinjection

the mouse Prnp transgene flanked by loxP sites into FVB Prnp0/0 (prion null) mice. The

transgenic mouse line expressed approximately 5-7 copies of the transgene, leading to

a Prnp expression of around 3-4 fold greater than wild type mice.

154

5.22 Methods

Prion inoculation

Animals were inoculated with 30 µL of 1% RML (Rocky Mountain Laboratories) brain

homogenate, derived from the scrapies prion strain PDG 586, or NBH (normal brain

homogenate), which served as a control. Inoculations were performed intracerebrally,

with hemizygous and homozygous mice succumbing to the prion infection at around 7

and 12 weeks post inoculation respectively (Chandler, 1961, Mallucci et al., 2003).

Inoculations were carried out by the research assistant Nicholas Verity.

Stereotaxic injection

Animals were anesthetised with the use of isoflurane (3% ATM for approximately 2

minutes in a closed chamber), and injected with 2 µL of 2 x 107 IFU into the CA1 region

of the hippocampus (2 mm posterior to bregma, 1.5-3 mm lateral and 1.8-2 mm

ventrals). Injections were performed by the research assistant Nicholas Verity.

Intranasal lentivirus Delivery

Four month old wild type mice (C57BL/6N mice) were initially anesthetised with the

use of isoflurane and remained under a lower dose during intranasal delivery (2%

ATM, via a hose system). Mice were placed into cushioned cradles with their heads

elevated to 70o in the supine position. Body temperature was maintained with the aid

of a heat pad at 37 oC. A total of 84 µL of either the MW1 or EV lentivirus (106 TU/mL)

was administered in 12 µL doses in alternate nostrils, with a two minute break

between each dose. Mice were administered with the lentivirus either on a single day,

or received the same dose on a weekly basis for three consecutive weeks. Mice were

sacrificed 14 days after their final lentiviral dose, and subsequently dissected for tissue

samples in order to analyse PRNP levels via quantitative PCR.

A similar experiment was conducted on tg37 mice, which were inoculated with either

RML or NBH at 3 weeks of age. Mice were administered with the MW1/EV lentivirus

intranasally at various time points post inoculation, and as above received a varied

number of doses (single day, weekly). Mice were examined carefully until they

155

developed the debilitating signs of prion disease, most notably an impairment of the

righting reflex, at which point mice were culled and dissected.

As a means of optimising the technique, several variations were made to the above

procedure, including changes to the dose amount, and the addition of nasal-uptake

enhancing compounds. These compounds included chitosan (Sigma-Aldrich), TMD

(Sigma-Aldrich) and rifampin (Sigma-Aldrich) (see Table 5.2), and were either

administered along with the lentiviral particles, or prior to the first dose. Each of these

compounds was also tested in the absence of lentiviral constructs, with a saline

control.

Compound Stock solution Working

concentration

Notes

Chitosan 10% in saline 1% Added to virus particles prior to use

TMD 1% in saline 0.125% Added to virus particles prior to use.

Rifampin 5mg/ml in 5% v/v methanol in saline

5 mg/mL 5 µL administered intranasally per nostril, 7 minutes prior to first lentiviral dose.

RNA isolation

Following dissection of the various brain regions (hippocampi, midbrain, cerebellum,

brainstem, cortex, olfactory bulbs and trigeminal nerves), samples were stored in

RNAlater solution (Ambion) at 20 oC to prevent further degradation. Samples were

homogenized with the use of the lysing matrix D tube (Fisher), and total RNA isolated

with Trizol reagent (Ambion) as previously described. RNA samples were quantitated

with the use of the NanoDrop ND-2000c Spectrophotometer (Thermo Scientific).

Table 5.2. Intranasal enhancers. Several compounds previously shown to increase

uptake at the nasal epithelium were used to enhancer uptake of the MW1/EV lentiviral

constructs. Stock solutions of the chitosan and TMD were made and added to the

lentiviral particles to achieve the working concentrations stated. 5 µL of the stock

solution of rifampin was administered per nostril to mice 7 minutes before their first

intranasal lentiviral dose.

156

Reverse transcription and quantitative RT-PCR

Prion protein levels were assessed with the use of real-time PCR, as previously

described. Analysis was conducted with the use of the Applied Biosystems

StepOne™plus Real-Time PCR System, coupled with the Fast SYBR® Green Master Mix

(Applied Biosystems). Amplification was carried out using 0.4 µM PRNP forward and

reverse primers (see Table 5.3), and data normalised to a reference gene, RP114 (0.9

µM forward primer and 0.3 µM reverse primer). Amplification consisted of a

denaturation step at 95 oC for 10 minutes, followed by 40 cycles of 3 seconds at 95 oC

and 30 seconds at 60 oC, with fluorescence monitored at the 60 oC stage of each cycle.

Subsequent quantification of results was performed using the δδCT method as in

Pfaffl (2001)

Primer Sequence (5’-3’) Tm (oC)

PRNP Forward primer GAGCCAAGCAGACTATCAGTC 60.6

PRNP Reverse primer TCAGTCCACATAGTCACAAAGAG 60.8

RP114 Forward primer GGCCGGTCTCTCGTTCTCA 68.0

RP114 Reverse primer TTACAGAAAGTCCTTCGGGTTTTT 64.6

Immunofluorescence (paraffin embedded tissue sections)

Following dissection, half of the brain was fixed in 10% formalin for five days, before

being dehydrated in 70% ethanol, cleared in xylene (Fisher), and embedded in

paraffin. Samples were cut sagittally into 7 µm thick slices, with the use of a

microtome. These samples were deparaffinised in xylene, rehydrated via decreasing

concentrations of ethanol, and underwent antigen retrieval through boiling in 0.01M

sodium citrate. Histological preparation of brain tissue was carried out by Jennifer

Edward, MRC toxicology, Leicester, UK. Once prepared, slides were washed in 0.05M

TBS and blocked in TrisA (0.05 M TBS and 10% Triton X) and 2% goat serum for 1 hour.

Slides were subsequently incubated in 1:1000 (diluted in 0.05 M TBS and 2% goat

Table 5.3. List of RT-PCR primer sequences. The sequences and annealing

temperatures of primers used to analyse PRNP expression levels relative to the

endogenous control, RP114, a ribosomal protein.

157

serum) rabbit anti-GFP (Abcam ab290) overnight at 4oC. Slides were then washed 3

times in TBST and incubated in 1:1000 dilution of secondary antibody, an AlexaFlour

488 conjugated goat anti-rabbit antibody (Invitrogen) for one hour in the dark. Finally

slides were washed and mounted with the aid of mounting media (VECTASHIELD®).

Images were taken with the use of a fluorescence microscope (Ziess Axiovert 200M,

Zeiss Colibri controller), through a 10X magnification, using Axiovision 4.8 software

(Carl Zeiss)

Haematoxylin and eosin staining

Paraffin-embedded brain slices were stained with the nuclear dye, haematoxylin and

cytoplasmic protein dye, eosin. Slides were prepared and rehydrated as stated above.

Slides were then washed in distilled water for 3 minutes, and incubated in Harris’

Hematoxylin (Sigma-Aldrich) for 15 minutes. Slides were further washed in distilled

water for 1 minutes, incubated in 1% HCl for 15 seconds, and washed in water for 9

minutes. This was followed by a 3 minutes incubation in 1% aqueous Eosin Y (Sigma-

Aldrich), a 2 minute wash in water, a 5 minute incubation in 70% IMS, and finally a 7

minute incubation in xylene. Slide were mounted in the synthetic resin, DPX

(distyrene, plasticizer and xylene) (Sigma-Aldrich), and allowed to dry. Histology work

was performed by Jennifer Edward, MRC toxicology, Leicester, UK.

Brain homogenisation

Following dissection, hippocampal tissue was isolated and stored at -80 oC until

required. Prior to immunoblotting, hippocampal tissue was homogenised with the aid

of a hand-held micro-grinder in 200 µL of 2X homogenisation buffer (100 mM Tris

pH8.0; 300 mM sodium chloride; 4 mM EDTA; 2mM magnesium chloride; 200mM

sodium fluoride (Sigma-Aldrich); 20% (v/v) glycerol; 2% (v/v)Triton X-100; 2% (w/v)

sodium deoxycholate (Sigma-Aldrich); 0.2% SDS; and 0.25 M sucrose). Following

homogenisation, 200 µL of lysis buffer (0.25 M sucrose, 50 mM TRIS pH 7.4, 1 mM

EDTA, phos-STOP tablet (Roche) (one tablet per 10 mL of buffer), 1X complete

protease inhibitor (Roche)) was added to samples. Protein samples were stored on ice

for 10 minutes, centrifuged at 13,000 r.p.m for 10 minutes at 4 oC, and the

supernatant stored at -80 oC for further use.

158

Protein quantitation

Protein concentration was assayed using 10X Coomassie blue based Bradford reagent

(Bio-Rad). 1 µL of protein samples was mixed with 1 mL of 1X Bradford reagent to

form a homogenous blue solution. Samples were placed into clear polystyrene

cuvettes (Bio-Rad) and the absorbance at 595 nm determined using of the

Biophotometer Spectrometer (Eppendorf).

Preparation of protein samples for immunoblotting

To analyse the levels of PrpC, PrpSc and several synaptic proteins in the hippocampus

of RML/NBH infected mice, an immunoblot technique was employed. Samples were

prepared specifically for each of the three blots, for example 20 µg of samples

intended to determine synaptic protein levels, were diluted in 15 µL of sterile distilled

water. For the analysis of prion levels ~20 µg protein was diluted in 20 µL of

homogenisation buffer for PrpC and PrpSc blots, with the samples for the latter blot

being treated with 1µl of µg/µL proteinase K and incubated at 37 oC for 1 hour. 4 µL of

5x SDS loading dye (0.25% (w/v) bromophenol blue; 10% (w/v) SDS; 10 mM DTT; 20%

(v/v) glycerol; 0.2 M Tris-HCl pH 6.8), was added to all samples and heated at 95oC

prior polyacrylamide gel eletrophoresis.

SDS-Polyacrylamide gel electrophoresis and transfer

Following denaturation, protein samples were separated by gel electrophoresis.

Samples to determine synaptic proteins levels were separated on a 10%

polyacrylamide gel, while PrP were separated on 12% gels. The 10% resolving gel

solution was made up of: 4ml of distilled water; 2.5 mL 1.5 M Tris pH 8.8; 3.3 mL of

30% acrylamide/bisacrylamide 37.5:1 (Geneflow); 100 µLof 10% (w/v) SDS; 100 µL of

10% (w/v) ammonium persulphate (APS) (Sigma-Aldrich); 10 µL

Tetramethylethylenediamine (TEMED) (National Diagnostics). The 12% resolving gel

solution was made up of: 3.3 mL of distilled water; 2.5 mL 1.5M Tris pH 8.8; 4 mL of

30% acrylamide/bisacrylamide 37.5:1; 100 µL of 10% (w/v) SDS; 100 µL of 10% (w/v)

APS; 10 µL TEMED. In both cases the stacking gel solution comprised of: 3 mL of

distilled water; 1.26 mL of 0.25 M Tris pH6.8; 660 µL 30% acrylamide/bisacrylamide

159

37.5:1; 50 µL of 10% (w/v) SDS; 50 µL of 10% (w/v) APS; 5 µL TEMED (the APS and

TEMED were added prior to use).

Once set, the gel was secured into the electrode assembly, which was in turn placed

into the running tank. Sufficient 1X Running buffer (25 mM Tris; 192 mM glycine; 0.1%

(w/v) SDS), was poured into the tank, to cover the top of the wells. Samples pipetted

(long stemmed tips) into the wells, along with 5 µL of Spectra Multicolour broad range

protein ladder (Thermoscientific). An electrical current of 120V was applied to the gel

for 90 minutes to facilitate separation of proteins.

After sufficient separation of proteins, the electrode assembly was dismantled. The gel

was carefully layered on to the gel holder cassette (Bio-Rad) along with foam pads,

blotting paper and nitrocellulose membrane (Whatman) with a pore size of 0.2 μm

(Amersham) (layer consisted of a foam pad, blotting paper, gel, nitrocellulose

membrane, 2nd blotting paper, 2nd foam pad). Once assembled, the cassette was

placed into the electrode assembly (with current flowing from the negative electrode

through the gel and in turn the membrane). The electrode assembly was placed into

the transfer tank, which was subsequently filled with 1X transfer buffer, made up of:

12 mM Tris, 96 mM glycine (Fisher) and 20% methanol. An electrical current of 150V

was applied to the transfer cassette for 90 minutes.

Immunoblotting

Following transfer of protein from the gel to nitrocellulose, membranes were

incubated for 2 minutes in 0.1% (v/v) Ponceau S (Acros) in 5% acetic acid, to ensure

transfer of proteins. Membranes were then washed thoroughly in distilled water.

Synaptic protein blots were blocked in 5% milk in TBST for 1 hour at room

temperature, and then incubated overnight at 4 oC with constant rocking, in 5% milk

supplemented with the primary antibodies, 1 in 10,000 mouse anti-GAPDH

(glyceraldehyde 3-phosphate dehydrogenase) (monoclonal, Santa Cruz 32233), 1 in

10,000 rabbit anti-SNAP-25 (synaptosomal-associated protein 25) (polyclonal, Abcam

5666) and 1 in 1000 mouse anti-PSD-95 (post synaptic density-95) (polyclonal, BD

transduction laboratories 610496). Membranes were rinsed three times with TBST,

160

and then washed a further three times, each for 5 minutes. The membranes were

subsequently probed with 1 in 10,000 goat anti-rabbit and anti-mouse horse radish

peroxidase (HRP) conjugated secondary antibodies (Dako), for 1 hour at room

temperature, followed by five 5 minute washes in TBST. Finally membranes were

covered with ECL Primer Western blotting detection reagent (Amersham) and

incubated for 5 minutes. The membranes were subsequently exposed to radiographic

film (Fuji) for varying lengths of time in dark conditions (typically 2 minutes). Film was

developed in developer agent (RG Universal) until bands became visible, immersed in

stop solution (0.3% acetic acid, Fisher Scientific), and finally fixed with the use of

fixation fluid (RG Universal).

Both of the PrP blots were initially incubated overnight in PBST, and then probed with

1 in 10,000 mouse anti-prion ICSM35 primary antibody (monoclonal, D-gen 0130-

03501) in PBST for 1 hour at room temperature. Membranes were the washed three

times in PBST, and incubated with PBST containing 1 in 10000 goat anti-mouse HRP

conjugated secondary antibody for a further hour at room temperature. Membranes

were washed 5 times with PBST, and incubated in Pierce ECL Western blotting

substrate (Thermoscientific) for 5 minutes. Blots were developed as stated above.

Image J was used to measure the average pixel density of each protein band, and in

the case of the synaptic proteins blot, samples were normalised to the GAPDH levels.

PrPC blots were stripped (see below), probed with tubulin, and normalised to this

internal control.

Stripping of immunoblot membranes

PrPC blots were incubated in stripping buffer (2% SDS; 100 mM 2-mercaptoethanol;

and 62.5 mM Tris pH6.8), at 50 oC for 30 minutes with occasional shaking. Membranes

were washed three times in TBST, and subsequently blocked in 3% BSA in TBST for an

hour at room temperature. The membranes were incubated with 1 in 5000 mouse

anti-tubulin antibody in 3% BSA over night at 4 oC. Following incubation, blots were

washed three times in TBST, and probed with1 in 10000 goat anti-mouse HRP

conjugated secondary antibody in 3% BSA for a further hour at room temperature.

Finally membranes were washed a further 5 times in TBST, incubated in ECL Primer

161

immunoblotting detection reagent (Amersham) for 5 minutes and developed as stated

above.

162

5.3 Results

5.31 GFP expression detected in both the cortex and hippocampus following

intranasal delivery of MW1 or EV into prion infected hemizygous mice

Previous work conducted in prion infected tg37 models of prion disease, revealed a

reversal of the prion disease phenotype in animals treated stereotactically with the

MW1 PRNP knockdown lentivirus (White et al., 2008b). We sought to determine

whether the therapeutic benefits of treating prion inoculated hemizygous mice with

this construct could be achieved when administered via a nasal route. Animals were

administered either the MW1 or EV lentivirus at 6, 7 and 8 weeks post inoculation

(w.p.i). Mice were culled via cervical dislocation upon manifestation of prion disease

symptoms, notably a loss of the righting reflex. Interestingly immunohistochemistry

against the GFP gene marker found in both lentivirus, revealed significant expression

of the constructs in both the hippocampal and cortical cells of terminally ill mice

treated intranasally at these timepoints compared to untreated RML only animals (see

Figures 5.3 and 5.4). As illustrated in the hippocampal images, greater expression of

GFP was detected in CA3 the regions of the hippocampus relative to the CA1 and

dentate gyrus. Images of the cortex showed a more even spread of GFP containing

cells throughout this region of the brain. Given that the shRNA and GFP expression are

regulated by different promoters, expression of the latter is only surrogate indicator of

knockdown, though these results indicate that both MW1 and EV are able to reach

these brain regions following intranasal delivery.

163

Figure 5.3. GFP is expressed in the hippocampi of mice treated with three doses of either EV or MW1. Prion infected mice were intranasally treated with either the MW1 or EV lentiviral construct at 6, 7 and 8 w.p.i. Both constructs expressed GFP, which was detected using of rabbit anti-GFP and goat anti-rabbit (conjugated to an AlexaFluor 488). Images were taken at 10X magnification of the whole hippocampus, as well as a 40X magnification of the CA3 region.

164

5.32 The MW1 Prnp knockdown lentiviral constructs failed to significantly reduce

prion levels in C57BL6N wild type mice

To determine the effectiveness of the intranasally administered MW1 lentiviral

construct in eliciting knock down of mouse Prnp expression, we intranasally

administered this virus or an empty vector (EV) viral control into C57BL6N WT mice.

Such work was conducted in parallel to that of prion infected transgenic mice, though

given the period of prion incubation, expression work was carried out in these WT

mice. Animals were sacrificed 14 days after treatment and the brainstem, cortex,

Figure 5.4 GFP is expressed in the cortex of mice treated with three doses of either EV or MW1. Prion infected mice were intranasally treated with either the MW1 or EV lentiviral construct at 6, 7 and 8 w.p.i. Both constructs expressed GFP, which detected using rabbit anti-GFP and goat anti-rabbit (conjugated to an AlexaFluor 488). Images were taken at both a 10X and 40X magnification.

165

hippocampus and olfactory bulbs dissected out. PrP mRNA levels were analysed via

quantitative PCR, which revealed only a moderate drop in expression in the brainstem

and hippocampus, no change in the olfactory bulbs and a slight increase in the cortex,

though in each case results were statistically insignificant (see Figure 5.5). The results

were suggestive, with a downward trend in both in the brainstem and hippocampus.

We analysed expression after only 14 days, it may be that later analysis would have

shown further reduction.

Figure 5.5. The MW1 PRNP knockdown construct failed to significantly reduce prion levels brain regions of wild type C57BL6N mice. Quantitative PCR revealed no significant drop in PRNP levels following the intranasal delivery of MW1 lentiviral construct in the cortex (B) or the olfactory bulbs (D), though we noted a moderate, but not significant, drop in expression in brainstem (A) and hippocampus (C). Mann-Whitney non parametric test, (Untreated N=3, EV treated N=7, MW1 treated N=3)

166

5.33 Prnp knockdown increased lifespan and a delayed onset of spongiosis in prion

infected homozygous mice.

Previously in the Mallucci group, analysis of PRNP expression changes via qPCR

following delivery of the MW1 lentivirus was been found to be very variable in

individual mice, while the collective in vivo effects of knockdown, such as survival and

burrowing were unequivocal. Therefore despite no significant changes in PRNP

expression following intranasal delivery of the MW1 lentivirus into C57BL6N WT mice,

we continued to administer the construct into the tg37 prion model. MW1 or an EV

was administered to homozygous mice at 3, 4 and 5 w.p.i of prion infected brain

homogenate (MW1 was administered only twice during this period). Mice were

subsequently monitored daily until they exhibited confirmatory symptoms of prion

disease. Although not statistically significant, we noted an increase in the survival of

mice treated with the MW1 lentivirus at 3 and 4 w.p.i relative to those delivered EV

(see Figure 5.6). Furthermore, haematoxylin and eosin staining of brain slices derived

from terminally ill mice treated at 3 and 4 w.p.i, as well as those animals culled at 7

w.p.i, revealed a reversal of spongiosis in mice administered MW1 relative to those

animals treated with EV (see Figure 5.7).

Figure 5.6. Treatment of homozygous prion inoculated mice with MW1 lentivirus at 3 and 4 weeks post inoculation extends life span of some mice. Mice were treated with a single, double or triple of the MW1, between 3 and 5 weeks post inoculation. Several of the mice administered the lentivirus at 3 and 4 w.p.i inoculation appeared to survive much longer than those of given the empty vector (EV). Mantel-Cox test (EV 3, 4, 5 w.p.i N=3, MW1 3 and 4 w.p.i N=3)

167

5.34 Early intranasal treatment of prion infected hemizygous mice revealed a delay

in spongiosis though no change in synaptic protein level changes.

Following the promising work conducted in the homozygous tg37 mice, we performed

the same study in hemizygous mice. Animals were treated with three doses of EV or

MW1 between 2 and 7 w.p.i with mice. Those mice treated at 2, 3 and 7 w.p.i

appeared to survive significantly longer than prion infected mice (RML only), though

exhibited a survival comparable to EV treated animals (see Figure 5.8).

Figure 5.7. Prion infected homozygous mice treated with MW1 at 3 and 4 weeks post inoculation displayed a similar degree of spongiosis at 7 weeks compared to EV treated mice, while terminally ill mice exhibited a reversal of spongiosis. Prion infected mice treated with either MW1 or EV were sacrificed at 7 w.p.i or upon exhibiting confirmatory signs of prion sickness. The brains of animals were dissected, sectioned, and stained with haematoxylin and eosin. At 7 w.p.i, the hippocampus, especially the CA3 region, of the mice treated with MW1 revealed a comparable level of spongiosis to those animals treated with EV. In contrast, brain sections derived from terminally ill mice displayed a reversal in spongiosis, with some mice administered MW1 exhibiting no visible vacuoles. Images taken at 5X magnification.

168

Haematoxylin and eosin staining of brain slices derived from the terminally ill mice

administered lentiviral constructs at 2, 3 and 7 weeks revealed a postponement of

spongiosis in mice treated with MW1, exhibiting an largely intact hippocampal ribbon

structure compared to untreated or EV treated animals. (see Figure 5.9).

To further determine the effectiveness of the intranasal delivery of MW1 at these

time points, we sacrificed mice at 9 w.p.i, dissected the hippocampal tissue and

analysed the levels of the synaptic proteins, SNAP-25 and PSD-95 (see Figure 5.10).

Our results demonstrated no significant change in PSD-95 levels in mice treated with

MW1 relative to tissue derived from animals inoculated with NBH or RML only and

then culled at 9.w.p.i, with both of the latter displaying similar protein levels. All three

of these groups exhibited much higher levels of PSD-95 than terminally ill RML only

treated mice. Levels of SNAP-25 were inconsistent with previous work conducted in

the Mallucci group , while MW1 treated mice exhibited much higher levels of this

synaptic protein than untreated prion infected mice taken at both 9 w.p.i and

Figure 5.8. The early treatment of prion inoculated hemizygous mice with MW1 construct at 2, 3 and 7 weeks post inoculation appears to extend life span of some mice. Mice were treated with three doses of the MW1 or EV between 2 and 7 weeks post inoculation. While mice treated with MW1 at 2, 3 and 7 w.p.i appeared to survive much longer than untreated animals (RML only), their survival was comparable to that of mice administered EV. Mantel-Cox test (RML only N=7, EV 2, 3 and 7 w.p.i N=3, MW1 2, 3 and 7 w.p.i N=5)

169

terminally ill animals, all three groups displayed protein levels much higher than that

of NBH inoculated mice.

Next we set out to assess total prion level coupled with PrPSc levels (see Figure 5.11).

While the prion levels in all four groups appeared to be similar, the PrPSc levels were

paradoxically found to be much greater in MW1 treated mice compared RML only

treated mice taken at 9 w.p.i, though much lower than RML only infected animals

taken when terminally ill.

170

Figure 5.9. Prion infected hemizygous mice treated with MW1 at 2, 3 and 7 weeks post inoculation displayed a delay in spongiosis relative to EV treated and RML only, untreated mice. Prion infected mice treated with either MW1, EV or left untreated were sacrificed once confirmatory signs of prion sickness became apparent. The brains of animals were dissected, sectioned, and stained with haematoxylin and eosin. The hippocampal ribbon (CA1 and CA3 regions) appeared to be intact in some mice administered MW1 compared to EV and RML only treated mice. Images taken at 5X magnification.

171

Figure 5.10. Mice treated with MW1 at 2, 3 and 7 w.p.i displayed no significant increase in PSD-95 relative to NBH or RML only treated mice at 9 w.p.i., though they exhibited a greater level of SNAP-25. A. Prion infected mice were intranasally treated with the MW1 lentiviral construct at 2, 3 and 7 w.p.i, and subsequently sacrificed at 9 w.p.i and hippocampal tissue dissected. Synaptic proteins levels were assessed via immunoblotting B. Blots revealed no significant change in PSD-95 levels following MW1 treatment compared to tissue derived from RML and NBH only inoculated mice. Untreated terminally ill RML infected animals displayed a significantly reduced level of this synaptic protein. C. MW1 treated mice also displayed a higher level SNAP-25 relative to RML and NBH only inoculated mice, the latter of which being comparable to terminally ill mice. Mann-Whitney non parametric test, (NBH N=3, RML only N=3, MW1 N=3, RML only term ill N=2)

172

5.35 Lowered lentiviral dose, or the use of chitosan, rifampin and TMD did not

extend survival of prion infected mice

In order to optimise the intranasal delivery of the MW1 lentivirus, we attempted to

lower the dose of the virus from 12 µL per nostril, with a total volume of 96 µL, to 4 µL

per nostril and a total volumeof 32 µL, as this had previously be shown to prevent

drainage into the trachea and increase uptake at the nasal epithelium (Dhuria et al.,

2010). We administered EV and MW1 intranasally into prion infected hemizygous

mice at 4, 5 and 6 w.p.i, and also stereotactically injected 96µl into mice at 4 w.p.i to

provide comparison. Such optimisation failed to extend the lifespan of hemizygous

prion infected mice relative to RML only treated mice or those administered EV

intranasally, while those mice injected with the MW1 lentivirus stereotaxically

displayed increased survival, though this increase was not statistically different to RML

only treated animals (see Figure 5.12).

Figure 5.11- PrPC levels in mice administered the MW1 lentiviral construct at 2, 3 and

7 w.p.i appeared to be similar to those displayed by RML only treated mice at 9 w.p.i,

while PrPSc

levels seemed to be elevated. A. Mice were inoculated with either NBH or RML, with the latter administered the MW1 or left untreated (RML only). All mice were sacrificed 9 w.p.i, and hippocampal tissue dissected and analysed via western blotting. The mouse prion protein was identified with the aid of the ICSM35 antibody. B. All RML infected mice displayed prion levels equivalent than those inoculated with NBH. C. Protein samples were treated with proteinase K (PK) prior to loading, with PrpSc species being resistant to such treatment. Mice administered with MW1 exhibited

higher levels PrPSc

, compared to both RML only treated mice and NBH inoculated animals. Mann-Whitney non parametric test, (NBH N=3, RML only N=3, MW1 N=3, RML only term ill N=2)

173

My second attempt at optimisation involved delivering the lentivirus together with

compounds believed to enhance uptake at the nasal epithelium. These included the

mucoadhesive chitosan, the permeation enhancer, TMD and the P-glycoprotein efflux

inhibitor rifampin. While delivery of MW1 with chitosan and rifampin failed to extend

the lifespan of homozygous prion infected mice, TMD combined with the lentivirus

was initially found to extend survival by several days relative to control animals

treated with the compound and EV (see Figure 5.13). However a follow up experiment

making use of TMD failed to yield a consistent result, and when combined with the

previous data, any significant extension in survival was lost.

Figure 5.12- Prion infected hemizygous mice treated with low doses of MW1 at 4, 5 and 6 weeks post inoculation displayed no significant increase in survival compared to EV and RML only treated animals. Prion infected mice were administered with a lower dose (4 µL per nostril, with a total of 32 µL) of either the MW1 or EV. In parallel, a small cohort of mice was stereotaxically injected with either of the constructs as a means of comparison. Animals injected with the MW1 virus displayed the longest survival, while intranasal delivery of the same construct led to no extension in survival compared to EV treated and RML only untreated mice. Mantel-Cox test (MW1 intranasal N=10, EV intranasal N=4, RML only N=5, MW1 injection N=8, EV injection N=3)

174

5.36 Treatment of prion infected hemizygous mice with MW1 at 6, 7 and 8 w.p.i

increased survival.

Following the promising results obtained in the previous set of experiments, we aimed

to determine whether a later dose of the MW1 knockdown lentivirus, could bring

about a greater therapeutic effect than that seen when administered between 2 and 7

w.p.i. Animals were treated with MW1 or EV between 6 and 8 w.p.i, receiving either

one to two doses. As in the previous experiment, mice were culled upon displaying

confirmatory prion disease signs. Mice treated with MW1 at 6, 7 and 8 w.p.i survived

appreciably longer than those left untreated or administered the EV, with some mice

Figure 5.13 Treatment of homozygous mice with chitosan, rifampin and TMD failed to improve the survival of mice intranasally administered with MW1. Mice were treated with either A. chitosan, B.rifampin, or C. TMD prior to, or alongside the intranasal delivery of the MW1 or EV lentiviral constructs at 2 and 3 w.p.i. In all cases, those treated with the MW1 construct failed to display any significant increase in survival relative to mice given EV or saline. Mantel-Cox test (chitosan + saline N=2, chitosan + EV-3, chitosan + MW1 N=3, rifampin + saline N=2, rifampin + EV N=3, rifampin + MW1 N=2, TMD + saline N=5, TMD + EV N=5, TMD + MW1 N=7)

175

surviving on average 5 days longer than RML only infected mice (see Figure 5.14).

Furthermore, there appeared to be no marked difference between mice receiving two

or three doses of the MW1 lentivirus. In contrast to the study conducted on

homozygous mice, haematoxylin and eosin staining of brain slices derived from

terminally ill hemizygous mice revealed similar levels of spongiosis in EV, MW1 treated

and untreated mice (Figure 5.15).

Figure 5.14. Treatment of hemizygous prion inoculated mice with MW1 at 6, 7 and 8 weeks post inoculation extends life span. Mice were treated with either a double or triple doses of the MW1 construct between 6 and 8 weeks post inoculation. Mice administered the lentivirus at 6, 7 and 8 weeks post weeks inoculation survived significantly longer than those of given the empty vector (EV) as well as RML only treated animals. Mantel-Cox test, P< 0.05 (*) (EV 6, 7 and 8 w.p.i N=3, RML only N=2, MW1 6, 7 and 8 w.p.i N=3)

176

Figure 5.15- Prion infected hemizygous mice treated with MW1 and EV at 6, 7 and 8 weeks post inoculation displayed comparable levels of spongiosis relative to RML only, untreated mice. Prion infected mice treated with either MW1 or EV were sacrificed once signs of prion sickness became apparent. The brains of animals were dissected, sectioned, and stained with haematoxylin and eosin. The hippocampal regions of mice administered MW1 or EV displayed no visible differences in the degree of spongiosis. Images taken at 5X magnification.

177

5. 4 Discussion

Intranasal delivery has recently grown in popularity as a means of conveying

compounds and gene vectors to the brain. The key advantages of this technique over

more invasive methods include the penetration of the BBB coupled with the ease of

administration, as well as the potential for long term therapy (Kulkarni et al., 2013).

Several lines of research have sought to combine this method with gene therapy, the

results of which hold great promise in the treatment of CNS disorders (Jiang et al.,

2012, Kim et al., 2009a, Kanazawa et al., 2013).

The aim of this portion of the research centred on the optimisation of intranasal

delivery as a means of delivering lentiviral constructs to the CNS, providing an

alternative route than the more invasive stereotaxic approach. Such work was

conducted in the context of a model of the neurodegenerative condition prion

disease, with the ultimate goal of using the same technique to overexpress our

splicing suppressor genes in the HD82Gln transgenic model of HD.

The study was designed to mimic the work carried out by White et al (2008), who

noted an 80% knockdown in PRNP levels following the stereotaxic injection of the

MW1 lentiviral construct into the hippocampus of wild-type mice (White et al.,

2008b). We did not observe a significant knockdown of Prnp by qPCR following

intranasal delivery of the construct, which may be due to the more diffuse spread of

the lentivirus associated with this delivery method, in contrast to the highly targeted

stereotaxic technique. As previously demonstrated, this widespread effect will have

reduced the number of transfected cells, resulting in a diminished level of vector

integration copy number and in turn a lower expression of the shRNA (Zielske et al.,

2004).

Despite the lack of statistical significance, the results derived from qPCR data

illustrated a modest drop in PRNP expression in both the hippocampus and the

brainstem following MW1 treatment, but no such change in the olfactory bulbs. This

may be indicative of the principal route of entry into the brain, with a greater influx of

lentiviral particles via the trigeminal route. This is supported by the work conducted by

Thorne et al, who found insulin-like growth factor levels to be 10 fold greater in the

178

trigeminal nerves relative to the olfactory bulbs, following intranasal delivery of the

compound into adult rats (Thorne et al., 2004). However given the shorter distance

from the nasal mucosa to the olfactory bulbs, the majority of studies making use of

this technique have noted a substantially larger concentration of intranasally

administered compounds in the olfactory bulbs compared to the brainstem (Dufes et

al., 2003, Johnson et al., 2010). This suggests that differences in the administered

compound, such as lipophilicity and molecular weight, may influence the route by

which they gain entry into the brain (Kulkarni et al., 2013).

Stereotaxic injection of MW1 into the hippocampus of prion infected tg37 hemizygous

mice had previously been shown to extend lifespan by ~24% relative to the EV treated

counterparts (White et al., 2008b). We performed similar experiments on both

homozygous and hemizygous mice, with the aim of testing whether this approach

emulated these results using the intranasal delivery method.

The work conducted in the prion infected tg37 homozygous mice revealed some,

albeit statistically insignificant, extension in survival following intranasal treatment

with the MW1 lentivirus at 3 and 4 w.p.i, with mice surviving on average 4.2 days

longer than those animals treated with EV. Some mice survived over a week longer

than control mice, a considerable extension given that untreated infected

homozygotes generally survive for 9 w.p.i before succumbing to prion sickness.

Furthermore, increased survival was coupled with a delayed onset of hippocampal

spongiosis in terminally ill mice, suggesting that intranasal administration of the MW1

lentivirus may have hindered disease.

We next sought to determine whether protection could be achieved in the

hemizygous transgenic mouse line. These mice displayed a less fulminant phenotype

in response to prion infection, allowing us to examine the best times points to

administer the MW1 virus in relation to disease progression. While mice treated with

MW1 between 2 and 7 w.p.i survived much longer than RML only treated animals,

these mice displayed a survival comparable to those animals administered EV. Upon

examination of brains derived from terminally ill mice we largely observed a delayed

spongiosis in mice treated with MW1, relative to RML only and EV treated animals.

179

PrPSc and synaptic proteins levels were similar in both EV and MW1 treated and

untreated mice at 9 w.p.i.. Surprisingly these PrPSc and synaptic protein levels were

comparable to those animals inoculated with normal brain homogenate (NBH),

suggesting further repeats of this experiment are warranted.

These results are in contrast to previous research conducted by the Mallucci group, in

which prion infection brought about a 50% decline in both SNAP-25 and PSD-95 levels

at 9 w.p.i compared to NBH treated animals. This deficit was abolished by stereotaxic

injection of the MW1 lentivirus into the hippocampus (Moreno et al., 2012). Such

findings were attributed to the induction of the unfolded protein response (UPR), a

direct consequence of prion protein accumulation. The UPR was found to be mediated

by the eukaryotic translation initiation factor, eIF2α-P. Upon stimulation of the UPR,

global translation is repressed, leading to a reduction in synaptic proteins and in turn a

loss of synaptic plasticity. Such synaptic dysfunction was restored through the

inhibition of eIF2α-P or knockdown of the prion protein with the MW1 lentivirus. Our

results suggest that intranasal delivery of the MW1 virus into the brain fails to

sufficiently knockdown PRNP to elicit the therapeutic effects seen when administered

via a stereotactic approach.

To enhance the uptake of the lentivirus, we employed several strategies to either

transiently alter the morphology of the nasal mucosa or increase the residence time of

intranasally administered compounds within the nasal cavity. One such method

included the use of a lower volume of intranasally delivered lentiviral constructs, as

previous research has demonstrated doses above 3-4µL would often lead to drainage

into the nasalphaynx and consequently respiratory difficulties (Dhuria et al., 2010). We

reduced our doses to 4µL, with a total volume of 32µL per day, but failed to observe

any significant change in survival or burrowing behaviour in hemizygous mice treated

with MW1 at 4, 5 and 6 w.p.i. Next we made use of compounds which increase

uptake at the nasal mucosa. These compounds included chitosan, a mucoadhesive

shown reduce mucocillary clearance while temporarily disrupting tight junctions at the

nasal epithelium; TMD, a permeation enhancer thought to alter the fluid state of the

membrane; and rifampin, an inhibitor of the efflux system found at the nasal mucosa

(Illum, 2007, Iqbal et al., 2003, Arnold et al., 2002, Arnold et al., 2004, Graff and

180

Pollack, 2003, Padowski and Pollack, 2010). Homozygous mice administered with the

MW1 virus alongside chitosan, rifampin or TMD at 2 and 3 w.p.i, all failed to display

any extension in life span. We can thus conclude that these compounds failed to

enhance uptake of our lentivirus to the brain.

We also sought to determine whether a later treatment of prion infected tg37

hemizygous with MW1 could produce survival results similar to though seen in the

homozygous transgenic model. Such work yielded promising results, with mice

treated with MW1 at 6, 7 and 8 w.p.i displaying a life span extended by 3 days, though

no reduction in spongiosis was observed. Additionally, immunohistochemistry

performed on tissue derived from these mice, revealed considerable GFP expression in

the hippocampus and cortex, illustrating the successful spread of both the MW1 and

EV constructs upon entering the brain.

Our results indicate a greater penetrance of virus into the brain, when administered at

later time points of disease progression. This may be attributed to several

observations relating to the widespread effects of prion infection. One of the key

features of prion particle infectivity is the capacity to spread via neural pathways.

Intracerebral prion infection has previously been shown to spread to the olfactory

bulbs, and sequentially to the olfactory epithelium via retrograde transport (DeJoia et

al., 2006). Prion deposits have been observed in the olfactory receptor neurons,

leading to degeneration and loss of these infected neurons, an effect coupled with

shedding of prion particles and further transmission (Corona et al., 2009, Bessen et

al., 2010) Additionally, recent structural studies of the prion protein has revealed a

membrane binding capacity, which can disrupt membrane stability and consequently

increase permeability (Sonkina et al., 2010). It is therefore feasible that prion induced

degeneration of olfactory and trigeminal nerve endings, or other neurons along these

pathways, may be conducive to nasal uptake. This may be brought about by an

increased permeability or through the transient gaps generated within the nasal

mucosa following loss of the olfactory nerves fibres.

Furthermore, work conducted by Van Den Pol et al and Lemiale et al showed

retrograde transport of intranasally administered viruses via the olfactory nerves, and

181

while both groups found evidence of their respective viruses reaching the olfactory

bulbs, neither reported further transport past this region of the brain (van den Pol et

al., 2002, Lemiale et al., 2003). The latter group found that the GFP-expressing

rhabdovirus was incapable of escaping the olfactory bulbs, and was eventually

expunged after 8 days post infection. This effect was attributed to activation of the

innate and specific immune response within the olfactory system (Reiss et al., 1998,

van den Pol et al., 2002).

In prion disease, the immune system, especially dendritic cells of the peripheral

lymphoid tissue are vulnerable to infection, given their high expression of the prion

protein and inability to mount a defensive attack against the infectious particles

(Brown et al., 1999, Glatzel and Aguzzi, 2000). Infected cells not only harbour the toxic

PrPSc particles, which induce conversion of host cell PrPC peptides, but also migrate

into the CNS (Aucouturier et al., 2000, Aguzzi et al., 2003, Aucouturier and Carnaud,

2002, Rosicarelli et al., 2005). This mechanism is crucial for prion entry into the brain

following infection and is observed throughout the course of disease pathogenesis,

including in those animals inoculated intracerebrally (Aguzzi et al., 2003, Bradford and

Mabbott, 2012). The migration of these cells has been postulated as a potential route

of entry for therapeutics into the prion-infected brain, and may facilitate the entry of

our intranasally administered virus (Rosicarelli et al., 2005).

Our results suggest that the lentivirus administered intranasally into hemizygous mice

at early points during prion disease progression, may fail to traverse the nasal

epithelium and/or enter the brain due to the healthy state of neuron endings and

early infection of the immune system. At later time points these neurons may have

been sufficiently degenerated to facilitate penetrance of our lentiviral constructs. The

subsequent entry into the brain may have been enhanced by the migratory influx of

infected dendritic cells in the CNS. Despite expression of the GFP reporter within the

hippocampus and cortex, we failed to observe any substantial extension in survival or

a reduction in spongiosis in hemizygous mice treated with MW1 at a later time point,

suggesting Prnp was not appreciably knocked down. This most likely reflects the lower

concentration of lentivirus reaching these regions of the brain compared to that

182

achieved with the aid of stereotaxis. This is consistent with the low level of PRNP

knockdown seen in wild type mice administered the MW1.

Recently intranasal delivery has been extensively investigated due to the multitude of

advantages stated earlier, and has been used to convey drugs, plasmids and other

compounds to the CNS. Delivery of our lentivirus appeared to be incredibly variable,

and although we attempted to optimise the technique within this study, our results

indicate that further optimisation is still required to obtain a much greater degree of

consistency between experiments. Regardless, our results demonstrate much promise

for the technique as a viable alternative to more invasive methods. Unfortunately

given the time constraints of the project, it was not possible to further investigate this

technique.

5.5 Future work

Since starting this project almost four years ago, intranasal delivery has grown in

popularity, especially as a means of targeting the brain (Kulkarni et al., 2013). During

this time, we have endeavoured to integrate current findings and variations regarding

the technique into our research. In many cases such variations did not increase

delivery of our lentivirus to the CNS, or more specifically alter the prion disease

phenotype.

In order to optimise this technique, we must gain a greater understanding of the

various routes from the nasal cavity to the brain. Such knowledge could be achieved

by analysing GFP reporter levels in the olfactory and brainstem regions at various time

points following intranasal delivery. By assessing the speed at which the lentivirus

reaches these regions, we could deduce not only the principal route by which it travels

to the brain, but also the form of transport. For example intraneuronal transport is

believed to be much slower than passive diffusion along the connective tissue

ensheathed around the olfactory and trigeminal neurons, which in turn conveys

compounds at a much slower rate than the convective bulk flow along the fluid filled

perineural spaces that run parallel to these neurons (Thorne et al., 2004). By

establishing the predominant route of entry into the brain, it may be possible to

modify the technique to enhance transport along such route.

183

Upon reaching the brain, the olfactory immune system has been shown to hamper the

progress of viruses past the olfactory bulbs. Immunodeficient mice virally infected via

the nasal route display a much greater degree of infectivity (Reiss et al., 1998,

Huneycutt et al., 1993). Treatment with immunosuppressants prior to intranasal

delivery of our lentivirus may therefore significantly increase the spread from the

olfactory bulbs. Although use of immunosuppressant drugs has been shown to be

therapeutic in some prion mouse models, such treatment fails to alter disease

progression in animals inoculated intracerebrally with prion infected brain

homogenate, and thus no confounding affects should occur by this approach (Aguzzi,

2003, Aucouturier and Carnaud, 2002, Aguzzi and Sigurdson, 2004).

Despite our failed attempts to enhance intranasal delivery with the aid of compounds

including chitosan, rifampin and TMD, formulations of the intranasally administered

substances and delivery conditions are crucial in obtaining optimal uptake at the nasal

mucosa. A multitude of compounds have been demonstrated to enhance nasal

uptake, these include methylcellulose and poly (acrylic acid), emulsifying agents that,

like chitosan, increase the residence time of nasal administered compounds in the

nasal cavity, and in turn impede mucociliary clearance (Vyas et al., 2006, Dhuria et al.,

2010). Compounds such as cyclodextrins have also been shown to increase nasal

uptake by solubilising nasal membrane proteins and enhancing permeability (Shao et

al., 1992, Nonaka et al., 2008). In contrast, the vasoconstrictor phenylephrine, reduces

blood flow to the nasal mucosa, thus limiting absorption of intranasally administered

substances into systemic circulation (Dhuria et al., 2009). Other compounds have also

been shown to impede uptake at the brain, including isoflurane gas, the anaesthetic

used during the delivery of our lentivirus. Animals are reported to display a 25%

reduction in brain uptake while in the anesthetised state (Toyama et al., 2004). In

future experiments we could assess such variations, not only in formulation, but also

by our choice of anaesthetic.

Currently much of the work regarding intranasal delivery and genetic knockdown has

made use of siRNA. Although siRNA has a much more transient effect than shRNA,

several groups have administered covalently modified siRNA with improved stability

and knockdown duration into rodents via the nasal passage. For example, the work

184

conducted by Renner et al and Kim et al have both demonstrated the efficacy of

intranasal delivery of naked siRNA, with the latter study reporting 70% knockdown in

both the amygdala and hypothalamus 12 hours after administration (Kim et al., 2009a,

Renner et al., 2012). More recently, siRNA was found to form stable complexes with

copolymers of poly (ethylene glycol), poly (ɛ-caprolactone) and the cell-penetrating

peptide Tat (MPEG-PCL-Tat), derived from HIV-Tat. Uptake of siRNA at the nasal

mucosa was significantly enhanced owing to the unique penetrating capacity of the

Tat peptide (Kanazawa et al., 2013). In addition to siRNA, the MPEG-PCL-Tat

copolymers enhance uptake of plasmid DNA, drugs, and combinations of all three.

(Tanaka et al., 2010, Kanazawa et al., 2013). As such, a similar system may be utilised

to enhance delivery of our PRNP knockdown construct, though the size of the

lentivirus may hinder formation of the complexes.

One of the main issues arising from our study was the diffuse expression of our

intranasally administered shRNA compared to results obtained with the more targeted

stereotaxic approach. This wider expression likely brought about a low copy number

integration, mirrored by a lower expression of the shRNA. This may have produced an

insufficient level of knockdown to elicit significant changes in our behavioural and

analytical metrics. Although increased doses of our lentivirus would increase

knockdown, it would not be financially feasible given the expense of generating

sufficient lentivirus particles. A cheaper alternative is the use of siRNA, which, as

stated above can yield robust knockdown (Renner et al., 2012, Kim et al., 2009a).

Although siRNA produces a transient knockdown effect, typically lasting only a few

days, chemical and covalent modifications, such as those seen in Dharmacon siSTABLE

constructs, can greatly enhance the stability and duration of the knockdown (Rao et

al., 2009, Renner et al., 2012). Furthermore, siRNA can be administered intranasally in

conjunction with the MPEG-PCL-Tat copolymers system mentioned above. Such work

has been conducted in rats with malignant glioma, whereby siRNA against the

oncogene Raf-1 was administered intranasally daily for one week, and thereby

increased in survival by almost 50% (Kanazawa et al., 2014). In light of these results,

daily doses of siRNA against the PRNP gene in our transgenic prion lines could

185

represent an easy and cost effective method of achieving sufficient knockdown to

alter disease progression.

186

Chapter 6

Concluding Remarks

During my doctoral research, I have worked to validate and characterise nine mRNA

splicing genes, previously shown to suppress mutant HTT in a yeast genetic modifier

screen. After generating overexpression constructs for each of these gene, I stably

transfected the constructs into our mammalian PC12 cell line. Subsequent cytotoxicity

assays revealed five of these candidate genes (CRNKL, EFTUD2, HNRNPF, HNRNPK and

TARDDP) to be protective when expressed alongside mutant HTT. Two of the gene hits

did not alter toxicity (HNRNPQ and SNRNB), while overexpression of the remaining

candidate genes exacerbated toxicity (HNRNPU and SFRS3).

The five protective genes were further investigated in order to determine the

mechanisms by which they suppress mutant HTT phenotypes. The suppressors were

found to modulate aggregation dynamics – while overexpression of HNRNPF, HNRNPK

and TARDDP acceleration aggregation, overexpression of CRNKL and EFTUD2 reduced

aggregation. Such changes in aggregation were mirrored by changes in mutant HTT

expression. It is feasible that overexpression of our candidate genes may alter global

transcription, mRNA stability or simply transcription from our inducible system,

though further investigation is required.

In addition to aggregation, I examined the role of mRNA splicing as a possible

mechanism to account for the protection observed in our mammalian cells. Although

it appeared that this post transcriptional process played no role in the protection

conveyed by our candidate genes, our splicing efficiency assay uncovered a polyQ

length dependent splicing defect - a result consistent with emerging research

implicating such defect in HD pathogenesis.

The more promising candidates were further interrogated in C. elegans models of HD.

RNAi mediated knockdown of our suppressor genes in nematodes failed to alter

mutant HTT induced behavioural phenotypes, though knockdown of CRNKL was

found to be lethal.

187

Ultimately I aimed to overexpress the most promising candidate genes in the brain of

HD mouse models via a lentiviral expression system, unfortunately this was not

possible due to time constraints. However, while characterising our HD knock-in

model mice, we identified a burrowing defect that manifested much earlier than

impairments on the more conventional rota-rod assay. As such, the burrowing

phenotype could be a robust and simple test in the study of HD in mice models.

Finally I explored an alternative, non-stereotactic method of administering lentiviruses

to the murine brain. To this end, I employed an intranasal method, in which the

viruses are delivered to the nasal mucosa, and travel in a retrograde manner within or

alongside the trigeminal and olfactory nerves, ultimately entering the brain via the

brainstem and olfactory bulbs respectively. I performed this work using a well-

established prion disease paradigm of neurodegeneration. This study made use of a

lentivirus to knockdown the PRP gene, an effect previously shown to stave, and to

some extent, reverse progression of prion disease. Although our results were more

variable compared to the stereotactic injection of the lentivirus, we showed

unequivocally that the virus had indeed reached the brain with the aid of a GFP

reporter. Interestingly, the results appeared to be dependent on the stage of prion

disease. Although these results were promising, further optimisation is required to

determine whether intranasal delivery could be a viable alternative to stereotactic

injection.

In summation, my work has provided insight into the modulatory potential of splicing

factors in HD, and thus contributing to the growing body of research linking splicing

and the pathogenesis of this devastating disorder.

188

References

ACHARYA, K. K., GOVIND, C. K., SHORE, A. N., STOLER, M. H. & REDDI, P. P. 2006. < i> cis</i>-Requirement for the maintenance of round spermatid-specific transcription. Developmental Biology, 295, 781-790.

AGGER, K., CLOOS, P. A. C., CHRISTENSEN, J., PASINI, D., ROSE, S., RAPPSILBER, J., ISSAEVA, I., CANAANI, E., SALCINI, A. E. & HELIN, K. 2007. UTX and JMJD3 are histone H3K27 demethylases involved in HOX gene regulation and development. Nature, 449, 731-U10.

AGUZZI, A. 2003. Prions and the immune system: a journey through gut, spleen, and nerves. Advances in immunology, 81, 123-171.

AGUZZI, A., HEPPNER, F. L., HEIKENWALDER, M., PRINZ, M., MERTZ, K., SEEGER, H. & GLATZEL, M. 2003. Immune system and peripheral nerves in propagation of prions to CNS. British medical bulletin, 66, 141-159.

AGUZZI, A. & SIGURDSON, C. J. 2004. Antiprion immunotherapy: to suppress or to stimulate? Nature Reviews Immunology, 4, 725-736.

ALAM, M. I., BEG, S., SAMAD, A., BABOOTA, S., KOHLI, K., ALI, J., AHUJA, A. & AKBAR, M. 2010. Strategy for effective brain drug delivery. European Journal of Pharmaceutical Sciences, 40, 385-403.

ALKAN, S., MARTINCIC, K. & MILCAREK, C. 2006. The hnRNPs F and H2 bind to similar sequences to influence gene expression. Biochem. J, 393, 361-371.

AMADA, N., TEZUKA, T., MAYEDA, A., ARAKI, K., TAKEI, N., TODOKORO, K. & NAWA, H. 2003. A novel rat orthologue and homologue for the Drosophila crooked neck gene in neural stem cells and their immediate descendants. Journal of Biochemistry, 133, 615-623.

ANCKAR, J. & SISTONEN, L. 2011. Regulation of HSF1 Function in the Heat Stress Response: Implications in Aging and Disease. In: KORNBERG, R. D., RAETZ, C. R. H., ROTHMAN, J. E. & THORNER, J. W. (eds.) Annual Review of Biochemistry, Vol 80. Palo Alto: Annual Reviews.

ANTON, F. & PEPPEL, P. 1991. CENTRAL PROJECTIONS OF TRIGEMINAL PRIMARY AFFERENTS INNERVATING THE NASAL-MUCOSA - A HORSERADISH-PEROXIDASE STUDY IN THE RAT. Neuroscience, 41, 617-628.

APOSTOL, B. L., ILLES, K., PALLOS, J., BODAI, L., WU, J., STRAND, A., SCHWEITZER, E. S., OLSON, J. M., KAZANTSEV, A., MARSH, J. L. & THOMPSON, L. M. 2006. Mutant huntingtin alters MAPK signaling pathways in PC12 and striatal cells: ERK1/2 protects against mutant huntingtin-associated toxicity. Human Molecular Genetics, 15, 273-285.

APOSTOL, B. L., KAZANTSEV, A., RAFFIONI, S., ILLES, K., PALLOS, J., BODAI, L., SLEPKO, N., BEAR, J. E., GERTLER, F. B., HERSCH, S., HOUSMAN, D. E., MARSH, J. L. & THOMPSON, L. M. 2003. A cell-based assay for aggregation inhibitors as therapeutics of polyglutamine-repeat disease and validation in Drosophila. Proceedings of the National Academy of Sciences of the United States of America, 100, 5950-5955.

ARNOLD, J., AHSAN, F., MEEZAN, E. & PILLION, D. J. 2002. Nasal administration of low molecular weight heparin. Journal of Pharmaceutical Sciences, 91, 1707-1714.

ARNOLD, J. J., AHSAN, F., MEEZAN, E. & PILLION, D. J. 2004. Correlation of tetradecylmaltoside induced increases in nasal peptide drug delivery with morphological changes in nasal epithelial cells. Journal of Pharmaceutical Sciences, 93, 2205-2213.

ARRASATE, M., MITRA, S., SCHWEITZER, E. S., SEGAL, M. R. & FINKBEINER, S. 2004. Inclusion body formation reduces levels of mutant huntingtin and the risk of neuronal death. Nature, 431, 805-810.

ASH, P. E., ZHANG, Y.-J., ROBERTS, C. M., SALDI, T., HUTTER, H., BURATTI, E., PETRUCELLI, L. & LINK, C. D. 2010. Neurotoxic effects of TDP-43 overexpression in C. elegans. Human Molecular Genetics, 19, 3206-3218.

189

ATWAL, R. S., DESMOND, C. R., CARON, N., MAIURI, T., XIA, J. R., SIPIONE, S. & TRUANT, R. 2011. Kinase inhibitors modulate huntingtin cell localization and toxicity. Nature Chemical Biology, 7, 453-460.

ATWAL, R. S., XIA, J., PINCHEV, D., TAYLOR, J., EPAND, R. M. & TRUANT, R. 2007. Huntingtin has a membrane association signal that can modulate huntingtin aggregation, nuclear entry and toxicity. Human Molecular Genetics, 16, 2600-2615.

AUCOUTURIER, P. & CARNAUD, C. 2002. The immune system and prion diseases: a relationship of complicity and blindness. Journal of leukocyte biology, 72, 1075-1083.

AUCOUTURIER, P., CARP, R. I., CARNAUD, C. & WISNIEWSKI, T. 2000. Prion diseases and the immune system. Clinical Immunology, 96, 79-85.

AUERBACH, W., HURLBERT, M. S., HILDITCH-MAGUIRE, P., WADGHIRI, Y. Z., WHEELER, V. C., COHEN, S. I., JOYNER, A. L., MACDONALD, M. E. & TURNBULL, D. H. 2001. The HD mutation causes progressive lethal neurological disease in mice expressing reduced levels of huntingtin. Human Molecular Genetics, 10, 2515-2523.

AZZOUZ, M. 2006. Gene Therapy for ALS: progress and prospects. Biochimica et Biophysica Acta (BBA)-Molecular Basis of Disease, 1762, 1122-1127.

AZZOUZ, M., MARTIN-RENDON, E., BARBER, R. D., MITROPHANOUS, K. A., CARTER, E. E., ROHLL, J. B., KINGSMAN, S. M., KINGSMAN, A. J. & MAZARAKIS, N. D. 2002. Multicistronic lentiviral vector-mediated striatal gene transfer of aromatic L-amino acid decarboxylase, tyrosine hydroxylase, and GTP cyclohydrolase I induces sustained transgene expression, dopamine production, and functional improvement in a rat model of Parkinson's disease. Journal of Neuroscience, 22, 10302-10312.

BAKER, H. & SPENCER, R. F. 1986. TRANS-NEURONAL TRANSPORT OF PEROXIDASE-CONJUGATED WHEAT-GERM-AGGLUTININ (WGA-HRP) FROM THE OLFACTORY EPITHELIUM TO THE BRAIN OF THE ADULT-RAT. Experimental Brain Research, 63, 461-473.

BANNAI, H., FUKATSU, K., MIZUTANI, A., NATSUME, T., IEMURA, S., IKEGAMI, T., INOUIE, T. & MIKOSHIBA, K. 2004. An RNA-interacting protein, SYNCRIP (heterogeneous nuclear ribonuclear protein Q1/NSAP1) is a component of mRNA granule transported with inositol 1,4,5-trisphosphate receptor type 1 mRNA in neuronal dendrites. Journal of Biological Chemistry, 279, 53427-53434.

BARGMANN, C. I. & HORVITZ, H. R. 1991. CHEMOSENSORY NEURONS WITH OVERLAPPING FUNCTIONS DIRECT CHEMOTAXIS TO MULTIPLE CHEMICALS IN C-ELEGANS. Neuron, 7, 729-742.

BARTELS, C., KLATT, C., LUHRMANN, R. & FABRIZIO, P. 2002. The ribosomal translocase homologue Snu114p is involved in unwinding U4/U6 RNA during activation of the spliceosome. EMBO Reports, 3, 875-880.

BENEDICT, C., HALLSCHMID, M., HATKE, A., SCHULTES, B., FEHM, H. L., BORN, J. & KERN, W. 2004. Intranasal insulin improves memory in humans. Psychoneuroendocrinology, 29, 1326-1334.

BENNETT, M. J., SAWAYA, M. R. & EISENBERG, D. 2006. Deposition diseases and 3D domain swapping. Structure, 14, 811-824.

BERGET, S. M. 1995. EXON RECOGNITION IN VERTEBRATE SPLICING. Journal of Biological Chemistry, 270, 2411-2414.

BESSEN, R. A., SHEARIN, H., MARTINKA, S., BOHARSKI, R., LOWE, D., WILHAM, J. M., CAUGHEY, B. & WILEY, J. A. 2010. Prion shedding from olfactory neurons into nasal secretions. PLoS pathogens, 6, e1000837.

BIRMAN, S. 2008. Neurodegeneration: RNA turns number one suspect in polyglutamine diseases. Current Biology, 18, R659-R661.

BIRNBOIM, H. C. & DOLY, J. 1979. RAPID ALKALINE EXTRACTION PROCEDURE FOR SCREENING RECOMBINANT PLASMID DNA. Nucleic Acids Research, 7, 1513-1523.

190

BOCHAROVA, N., CHAVE-COX, R., SOKOLOV, S., KNORRE, D. & SEVERIN, F. 2009. Protein aggregation and neurodegeneration: Clues from a yeast model of Huntington's disease. Biochemistry-Moscow, 74, 231-234.

BODNER, R. A., OUTEIRO, T. F., ALTMANN, S., MAXWELL, M. M., CHO, S. H., HYMAN, B. T., MCLEAN, P. J., YOUNG, A. B., HOUSMAN, D. E. & KAZANTSEV, A. G. 2006. Pharmacological promotion of inclusion formation: A therapeutic approach for Huntington's and Parkinson's diseases. Proceedings of the National Academy of Sciences of the United States of America, 103, 4246-4251.

BOLIVAR, V. J., MANLEY, K. & MESSER, A. 2004. Early exploratory behavior abnormalities in R6/1 Huntington's disease transgenic mice. Brain Research, 1005, 29-35.

BOMSZTYK, K., DENISENKO, O. & OSTROWSKI, J. 2004. HnRNP K: One protein multiple processes. Bioessays, 26, 629-638.

BOSE, J. K., WANG, I. F., HUNG, L., TARN, W. Y. & SHEN, C. K. J. 2008. TDP-43 Overexpression Enhances Exon 7 Inclusion during the Survival of Motor Neuron Pre-mRNA Splicing. Journal of Biological Chemistry, 283, 28852-28859.

BOSSY-WETZEL, E., SCHWARZENBACHER, R. & LIPTON, S. A. 2004. Molecular pathways to neurodegeneration. Nature Medicine, 10, S2-S9.

BRADBURY, M. W. B., CSERR, H. F. & WESTROP, R. J. 1981. DRAINAGE OF CEREBRAL INTERSTITIAL FLUID INTO DEEP CERVICAL LYMPH OF THE RABBIT. American Journal of Physiology, 240, F329-F336.

BRADFORD, B. M. & MABBOTT, N. A. 2012. Prion disease and the innate immune system. Viruses, 4, 3389-3419.

BRENNER, S. 1974. GENETICS OF CAENORHABDITIS-ELEGANS. Genetics, 77, 71-94. BRIGNULL, H. R., MORLEY, J. F., GARCIA, S. M. & MORIMOTO, R. I. 2006. Modeling polyglutamine

pathogenesis in C-elegans. In: KHETERPAL, I. & WETZEL, R. (eds.) Amyloid, Prions, and Other Protein Aggregates, Pt B. San Diego: Elsevier Academic Press Inc.

BROADWELL, R. D. & BALIN, B. J. 1985. ENDOCYTIC AND EXOCYTIC PATHWAYS OF THE NEURONAL SECRETORY PROCESS AND TRANSSYNAPTIC TRANSFER OF WHEAT-GERM AGGLUTININ-HORSERADISH PEROXIDASE INVIVO. Journal of Comparative Neurology, 242, 632-650.

BROOKS, S., HIGGS, G., JONES, L. & DUNNETT, S. B. 2012. Longitudinal analysis of the behavioural phenotype in Hdh((CAG)150) Huntington's disease knock-in mice. Brain Research Bulletin, 88, 182-188.

BROOKS, S. P. & DUNNETT, S. B. 2009. Tests to assess motor phenotype in mice: a user's guide. Nature Reviews Neuroscience, 10, 519-529.

BROWN, K., STEWART, K., RITCHIE, D., MABBOTT, N., WILLIAMS, A., FRASER, H., MORRISON, W. & BRUCE, M. 1999. Scrapie replication in lymphoid tissues depends on prion protein-expressing follicular dendritic cells. Nature Medicine, 5, 1308-1312.

BUCHNER, K., SEITZTUTTER, D., SCHONITZER, K. & WEISS, D. G. 1987. A QUANTITATIVE STUDY OF ANTEROGRADE AND RETROGRADE AXONAL-TRANSPORT OF EXOGENOUS PROTEINS IN OLFACTORY NERVE C-FIBERS. Neuroscience, 22, 697-707.

BÜELER, H., AGUZZI, A., SAILER, A., GREINER, R.-A., AUTENRIED, P., AGUET, M. & WEISSMANN, C. 1993. Mice devoid of PrP are resistant to scrapie. Cell, 73, 1339-1347.

BUGG, C. W., ISAS, J. M., FISCHER, T., PATTERSON, P. H. & LANGEN, R. 2012. Structural Features and Domain Organization of Huntingtin Fibrils. Journal of Biological Chemistry, 287, 31739-31746.

BURATTI, E. & BARALLE, F. E. 2011. TDP-43: new aspects of autoregulation mechanisms in RNA binding proteins and their connection with human disease. Febs Journal, 278, 3530-3538.

BURKE, K. A., GODBEY, J. & LEGLEITER, J. 2011. Assessing mutant huntingtin fragment and polyglutamine aggregation by atomic force microscopy. Methods, 53, 275-284.

191

BURKE, K. A., KAUFFMAN, K. J., UMBAUGH, C. S., FREY, S. L. & LEGLEITER, J. 2013. The Interaction of Polyglutamine Peptides with Lipid Membranes Is Regulated by Flanking Sequences Associated with Huntingtin. Journal of Biological Chemistry, 288, 14993-15005.

BUSCH, A. & HERTEL, K. J. 2012. Evolution of SR protein and hnRNP splicing regulatory factors. Wiley Interdisciplinary Reviews-Rna, 3, 1-12.

CACERES, J. F. & KORNBLIHTT, A. R. 2002. Alternative splicing: multiple control mechanisms and involvement in human disease. Trends in Genetics, 18, 186-193.

CACERES, J. F., SCREATON, G. R. & KRAINER, A. R. 1998. A specific subset of SR proteins shuttles continuously between the nucleus and the cytoplasm. Genes & Development, 12, 55-66.

CAMPESAN, S., GREEN, E. W., BREDA, C., SATHYASAIKUMAR, K. V., MUCHOWSKI, P. J., SCHWARCZ, R., KYRIACOU, C. P. & GIORGINI, F. 2011. The Kynurenine Pathway Modulates Neurodegeneration in a Drosophila Model of Huntington's Disease. Current Biology, 21, 961-966.

CARON, N. S., DESMOND, C. R., XIA, J. R. & TRUANT, R. 2013. Polyglutamine domain flexibility mediates the proximity between flanking sequences in huntingtin. Proceedings of the National Academy of Sciences of the United States of America, 110, 14610-14615.

CARPENTER, B., MCKAY, M., DUNDAS, S. R., LAWRIE, L. C., TELFER, C. & MURRAY, G. I. 2006. Heterogeneous nuclear ribonucleoprotein K is over expressed, aberrantly localised and is associated with poor prognosis in colorectal cancer. British Journal of Cancer, 95, 921-927.

CARTER, R. J., LIONE, L. A., HUMBY, T., MANGIARINI, L., MAHAL, A., BATES, G. P., DUNNETT, S. B. & MORTON, A. J. 1999. Characterization of progressive motor deficits in mice transgenic for the human Huntington’s disease mutation. The Journal of neuroscience, 19, 3248-3257.

CASSADA, R. C. & RUSSELL, R. L. 1975. DAUERLARVA, A POST-EMBRYONIC DEVELOPMENTAL VARIANT OF NEMATODE CAENORHABDITIS-ELEGANS. Developmental Biology, 46, 326-342.

CATTANEO, E., ZUCCATO, C. & TARTARI, M. 2005. Normal huntingtin function: An alternative approach to Huntington's disease. Nature Reviews Neuroscience, 6, 919-930.

CAUDY, A. A., KETTING, R. F., HAMMOND, S. M., DENLI, A. M., BATHOORN, A. M. P., TOPS, B. B. J., SILVA, J. M., MYERS, M. M., HANNON, G. J. & PLASTERK, R. H. A. 2003. A micrococcal nuclease homologue in RNAi effector complexes. Nature, 425, 411-414.

CHA, J.-H. J. 2007. Transcriptional signatures in Huntington's disease. Progress in Neurobiology, 83, 228-248.

CHANDLER, R. 1961. Encephalopathy in mice produced by inoculation with scrapie brain material. The Lancet, 277, 1378-1379.

CHATTERJEE, N. & SINHA, S. 2008. Understanding the mind of a worm: hierarchical network structure underlying nervous system function in C. elegans. In: BANERJEE, R. & CHAKRABARTI, B. K. (eds.) Models of Brain and Mind: Physical, Computational and Psychological Approaches. Amsterdam: Elsevier Science Bv.

CHELLGREN, B. W., MILLER, A. F. & CREAMER, T. P. 2006. Evidence for polyproline II helical structure in short polyglutamine tracts. Journal of Molecular Biology, 361, 362-371.

CHEN, H. H., CHANG, J. G., LU, R. M., PENG, T. Y. & TARN, W. Y. 2008. The RNA Binding Protein hnRNP Q Modulates the Utilization of Exon 7 in the Survival Motor Neuron 2 (SMN2) Gene. Molecular and Cellular Biology, 28, 6929-6938.

CHEN, H. H., YU, H. I., CHIANG, W. C., LIN, Y. D., SHIA, B. C. & TARN, W. Y. 2012. hnRNP Q Regulates Cdc42-Mediated Neuronal Morphogenesis. Molecular and Cellular Biology, 32, 2224-2238.

CHOMCZYNSKI, P. & SACCHI, N. 1987. SINGLE-STEP METHOD OF RNA ISOLATION BY ACID GUANIDINIUM THIOCYANATE PHENOL CHLOROFORM EXTRACTION. Analytical Biochemistry, 162, 156-159.

CHUNG, S. Y., MCLEAN, M. R. & RYMOND, B. C. 1999. Yeast ortholog of the Drosophila crooked neck protein promotes spliceosome assembly through stable U4/U6.U5 snRNP addition. Rna-a Publication of the Rna Society, 5, 1042-1054.

192

CHUNG, S. Y., ZHOU, Z. L., HUDDLESTON, K. A., HARRISON, D. A., REED, R., COLEMAN, T. A. & RYMOND, B. C. 2002. Crooked neck is a component implicated in the of the human spliceosome and splicing process. Biochimica Et Biophysica Acta-Gene Structure and Expression, 1576, 287-297.

CONSORTIUM, C. E. S. 1998. Genome sequence of the nematode C-elegans: A platform for investigating biology. Science, 282, 2012-2018.

COOPER, T. A., WAN, L. & DREYFUSS, G. 2009. RNA and disease. Cell, 136, 777-793. CORONA, C., PORCARIO, C., MARTUCCI, F., IULINI, B., MANEA, B., GALLO, M., PALMITESSA, C.,

MAURELLA, C., MAZZA, M. & PEZZOLATO, M. 2009. Olfactory system involvement in natural scrapie disease. Journal of virology, 83, 3657-3667.

COSTANTINI, L. C., BAKOWSKA, J. C., BREAKEFIELD, X. O. & ISACSON, O. 2000. Gene therapy in the CNS. Gene Therapy, 7, 93-109.

CRAMER, P., PESCE, C. G., BARALLE, F. E. & KORNBLIHTT, A. R. 1997. Functional association between promoter structure and transcript alternative splicing. Proceedings of the National Academy of Sciences of the United States of America, 94, 11456-11460.

CRICK, S. L., JAYARAMAN, M., FRIEDEN, C., WETZEL, R. & PAPPU, R. V. 2006. Fluorescence correlation spectroscopy shows that monomeric polyglutamine molecules form collapsed structures in aqueous solutions. Proceedings of the National Academy of Sciences of the United States of America, 103, 16764-16769.

CROOK, Z. R. & HOUSMAN, D. 2011. Huntington's disease: can mice lead the way to treatment? Neuron, 69, 423-435.

DARNELL, G. D., DERRYBERRY, J., KURUTZ, J. W. & MEREDITH, S. C. 2009. Mechanism of Cis-Inhibition of PolyQ Fibrillation by PolyP: PPII Oligomers and the Hydrophobic Effect. Biophysical Journal, 97, 2295-2305.

DE ALMEIDA, L. P., ZALA, D., AEBISCHER, P. & DEGLON, N. 2001. Neuroprotective effect of a CNTF-expressing lentiviral vector in the quinolinic acid rat model of Huntington's disease. Neurobiology of Disease, 8, 433-446.

DE BONO, M., TOBIN, D. M., DAVIS, M. W., AVERY, L. & BARGMANN, C. I. 2002. Social feeding in Caenorhabditis elegans is induced by neurons that detect aversive stimuli. Nature, 419, 899-903.

DEACON, R. M. 2006. Burrowing in rodents: a sensitive method for detecting behavioral dysfunction. Nature Protocols, 1, 118-121.

DEACON, R. M., CROUCHER, A. & RAWLINS, J. N. P. 2002. Hippocampal cytotoxic lesion effects on species-typical behaviours in mice. Behavioural brain research, 132, 203-213.

DEACON, R. M., RALEY, J. M., PERRY, V. H. & RAWLINS, J. N. P. 2001. Burrowing into prion disease. Neuroreport, 12, 2053-2057.

DEHAY, B. & BERTOLOTTI, A. 2006. Critical role of the proline-rich region in Huntingtin for aggregation and cytotoxicity in yeast. Journal of Biological Chemistry, 281, 35608-35615.

DEJOIA, C., MOREAUX, B., O'CONNELL, K. & BESSEN, R. A. 2006. Prion infection of oral and nasal mucosa. Journal of virology, 80, 4546-4556.

DHURIA, S. V., HANSON, L. R. & FREY, W. H. 2009. Novel Vasoconstrictor Formulation to Enhance Intranasal Targeting of Neuropeptide Therapeutics to the Central Nervous System. Journal of Pharmacology and Experimental Therapeutics, 328, 312-320.

DHURIA, S. V., HANSON, L. R. & FREY, W. H. 2010. Intranasal Delivery to the Central Nervous System: Mechanisms and Experimental Considerations. Journal of Pharmaceutical Sciences, 99, 1654-1673.

DLUGOSZ, M. & TRYLSKA, J. 2011. Secondary Structures of Native and Pathogenic Huntingtin N-Terminal Fragments. Journal of Physical Chemistry B, 115, 11597-11608.

DOI, H., OKAMURA, K., BAUER, P. O., FURUKAWA, Y., SHIMIZU, H., KUROSAWA, M., MACHIDA, Y., MIYAZAKI, H., MITSUI, K. & KUROIWA, Y. 2008. RNA-binding protein TLS is a major nuclear

193

aggregate-interacting protein in huntingtin exon 1 with expanded polyglutamine-expressing cells. Journal of Biological Chemistry, 283, 6489-6500.

DRAGATSIS, I., LEVINE, M. S. & ZEITLIN, S. 2000. Inactivation of Hdh in the brain and testis results in progressive neurodegeneration and sterility in mice. Nature Genetics, 26, 300-306.

DRAGILEVA, E., HENDRICKS, A., TEED, A., GILLIS, T., LOPEZ, E. T., FRIEDBERG, E. C., KUCHERLAPATI, R., EDELMANN, W., LUNETTA, K. L. & MACDONALD, M. E. 2009. Intergenerational and striatal CAG repeat instability in Huntington's disease knock-in mice involve different DNA repair genes. Neurobiology of Disease, 33, 37-47.

DREDGE, B. K., POLYDORIDES, A. D. & DARNELL, R. B. 2001. The splice of life: Alternative splicing and neurological disease. Nature Reviews Neuroscience, 2, 43-50.

DUFES, C., OLIVIER, J.-C., GAILLARD, F., GAILLARD, A., COUET, W. & MULLER, J.-M. 2003. Brain delivery of vasoactive intestinal peptide (VIP) following nasal administration to rats. International Journal of Pharmaceutics, 255, 87-97.

DUGAICZYK, A., BOYER, H. W. & GOODMAN, H. M. 1975. LIGATION OF ECORI ENDONUCLEASE-GENERATED DNA FRAGMENTS INTO LINEAR AND CIRCULAR STRUCTURES. Journal of Molecular Biology, 96, 171-&.

DUJON, B. 1996. The yeast genome project: what did we learn? Trends in Genetics, 12, 263-270. DUNAH, A. W., JEONG, H., GRIFFIN, A., KIM, Y. M., STANDAERT, D. G., HERSCH, S. M., MOURADIAN,

M. M., YOUNG, A. B., TANESE, N. & KRAINC, D. 2002. Sp1 and TAFII130 transcriptional activity disrupted in early Huntington's disease. Science, 296, 2238-2243.

DUNHAM, N. & MIYA, T. 1957. A note on a simple apparatus for detecting neurological deficit in rats and mice. Journal of the American Pharmaceutical Association, 46, 208-209.

DUNING, K., BUCK, F., BAMEKOW, A. & KREMERSKOTHEN, J. 2008. SYNCRIP, a component of dendritically localized mRNPs, binds to the translation regulator BC200 RNA. Journal of Neurochemistry, 105, 351-359.

DUYAO, M., AMBROSE, C., MYERS, R., NOVELLETTO, A., PERSICHETTI, F., FRONTALI, M., FOLSTEIN, S., ROSS, C., FRANZ, M., ABBOTT, M., GRAY, J., CONNEALLY, P., YOUNG, A., PENNEY, J., HOLLINGSWORTH, Z., SHOULSON, I., LAZZARINI, A., FALEK, A., KOROSHETZ, W., SAX, D., BIRD, E., VONSATTEL, J., BONILLA, E., ALVIR, J., CONDE, J. B., CHA, J. H., DURE, L., GOMEZ, F., RAMOS, M., SANCHEZRAMOS, J., SNODGRASS, S., DEYOUNG, M., WEXLER, N., MOSCOWITZ, C., PENCHASZADEH, G., MACFARLANE, H., ANDERSON, M., JENKINS, B., SRINIDHI, J., BARNES, G., GUSELLA, J. & MACDONALD, M. 1993. TRINUCLEOTIDE REPEAT LENGTH INSTABILITY AND AGE-OF-ONSET IN HUNTINGTONS-DISEASE. Nature Genetics, 4, 387-392.

DUYAO, M. P., AUERBACH, A. B., RYAN, A., PERSICHETTI, F., BARNES, G. T., MCNEIL, S. M., GE, P., VONSATTEL, J.-P., GUSELLA, J. F. & JOYNER, A. L. 1995. Inactivation of the mouse Huntington's disease gene homolog Hdh. Science, 269, 407-410.

EDENFELD, G., VOLOHONSKY, G., KRUKKERT, K., NAFFIN, E., LAMMEL, U., GRIMM, A., ENGELEN, D., REUVENY, A., VOLK, T. & KLAMBT, C. 2006. The splicing factor crooked neck associates with the RNA-binding protein HOW to control glial cell maturation in Drosophila. Neuron, 52, 969-980.

EHRNHOEFER, D. E., DUENNWALD, M., MARKOVIC, P., WACKER, J. L., ENGEMANN, S., ROARK, M., LEGLEITER, J., MARSH, J. L., THOMPSON, L. M., LINDQUIST, S., MUCHOWSKI, P. J. & WANKER, E. E. 2006. Green tea (-)-epigallocatechin-gallate modulates early events in huntingtin misfolding and reduces toxicity in Huntington's disease models. Human Molecular Genetics, 15, 2743-2751.

EIDELBERG, D. & SURMEIER, D. J. Brain networks in Huntington disease. Journal of Clinical Investigation, 121, 484-492.

FABER, P. W., ALTER, J. R., MACDONALD, M. E. & HART, A. C. 1999. Polyglutamine-mediated dysfunction and apoptotic death of a Caenorhabditis elegans sensory neuron. Proceedings of the National Academy of Sciences of the United States of America, 96, 179-184.

194

FABER, P. W., VOISINE, C., KING, D. C., BATES, E. A. & HART, A. C. 2002. Glutamine/proline-rich PQE-1 proteins protect Caenorhabditis elegans neurons from huntingtin polyglutamine neurotoxicity. Proceedings of the National Academy of Sciences of the United States of America, 99, 17131-17136.

FABRIZIO, P., LAGGERBAUER, B., LAUBER, J., LANE, W. S. & LUHRMANN, R. 1997. An evolutionarily conserved U5 snRNP-specific protein is a GTP-binding factor closely related to the ribosomal translocase EF-2. Embo Journal, 16, 4092-4106.

FERNANDES, H. B., BAIMBRIDGE, K. G., CHURCH, J., HAYDEN, M. R. & RAYMOND, L. A. 2007. Mitochondrial sensitivity and altered calcium handling underlie enhanced NMDA-induced apoptosis in YAC128 model of Huntington's disease. The Journal of neuroscience, 27, 13614-13623.

FERNANDEZ-FUNEZ, P., NINO-ROSALES, M. L., DE GOUYON, B., SHE, W. C., LUCHAK, J. M., MARTINEZ, P., TURIEGANO, E., BENITO, J., CAPOVILLA, M., SKINNER, P. J., MCCALL, A., CANAL, I., ORR, H. T., ZOGHBI, H. Y. & BOTAS, J. 2000. Identification of genes that modify ataxin-1-induced neurodegeneration. Nature, 408, 101-106.

FERRANTE, R. J. 2009. Mouse models of Huntington's disease and methodological considerations for therapeutic trials. Biochimica et Biophysica Acta (BBA)-Molecular Basis of Disease, 1792, 506-520.

FIRE, A., ALBERTSON, D., HARRISON, S. W. & MOERMAN, D. G. 1991. PRODUCTION OF ANTISENSE RNA LEADS TO EFFECTIVE AND SPECIFIC-INHIBITION OF GENE-EXPRESSION IN C-ELEGANS MUSCLE. Development, 113, 503-514.

FIRE, A., XU, S. Q., MONTGOMERY, M. K., KOSTAS, S. A., DRIVER, S. E. & MELLO, C. C. 1998. Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature, 391, 806-811.

FITZPATRICK, M. J. & SOKOLOWSKI, M. B. 2004. In search of food: Exploring the evolutionary link between cGMP-dependent protein kinase (PKG) and behaviour. Integrative and Comparative Biology, 44, 28-36.

FORMAN, M. S., TROJANOWSKI, J. Q. & LEE, V. M. 2004. Neurodegenerative diseases: a decade of discoveries paves the way for therapeutic breakthroughs. Nature Medicine, 10, 1055-1063.

FRASER, A. G., KAMATH, R. S., ZIPPERLEN, P., MARTINEZ-CAMPOS, M., SOHRMANN, M. & AHRINGER, J. 2000. Functional genomic analysis of C-elegans chromosome I by systematic RNA interference. Nature, 408, 325-330.

FRAZER, L. N., LOVELL, S. C. & O'KEEFE, R. T. 2009. Analysis of Synthetic Lethality Reveals Genetic Interactions Between the GTPase Snu114p and snRNAs in the Catalytic Core of the Saccharomyces cerevisiae Spliceosome. Genetics, 183, 497-515.

FREIBAUM, B. D., CHITTA, R. K., HIGH, A. A. & TAYLOR, J. P. 2010. Global Analysis of TDP-43 Interacting Proteins Reveals Strong Association with RNA Splicing and Translation Machinery. Journal of Proteome Research, 9, 1104-1120.

FUENTEALBA, R. A., UDAN, M., BELL, S., WEGORZEWSKA, I., SHAO, J., DIAMOND, M. I., WEIHL, C. C. & BALOH, R. H. 2010. Interaction with Polyglutamine Aggregates Reveals a Q/N-rich Domain in TDP-43. Journal of Biological Chemistry, 285, 26304-26314.

FUJIKAKE, N., NAGAI, Y., POPIEL, H. A., OKAMOTO, Y., YAMAGUCHI, M. & TODA, T. 2008. Heat shock transcription factor 1-activating compounds suppress polyglutamine-induced neurodegeneration through induction of multiple molecular chaperones. Journal of Biological Chemistry, 283, 26188-26197.

FUJIWARA, M., SENGUPTA, P. & MCINTIRE, S. L. 2002. Regulation of body size and behavioral state of C-elegans by sensory perception and the EGL-4 cGMP-dependent protein kinase. Neuron, 36, 1091-1102.

GAFNI, J. & ELLERBY, L. M. 2002. Calpain activation in Huntington's disease. Journal of Neuroscience, 22, 4842-4849.

195

GAFNI, J., HERMEL, E., YOUNG, J. E., WELLINGTON, C. L., HAYDEN, M. R. & ELLERBY, L. M. 2004. Inhibition of calpain cleavage of huntingtin reduces toxicity - Accumulation of calpain/caspase fragments in the nucleus. Journal of Biological Chemistry, 279, 20211-20220.

GAO, X. L., CHEN, J., TAO, W. X., ZHU, J. H., ZHANG, Q. Z., CHEN, H. Z. & JIANG, X. G. 2007. UEA I-bearing nanoparticles for brain delivery following intranasal administration. International Journal of Pharmaceutics, 340, 207-215.

GAO, X. L., TAO, W. X., LU, W., ZHANG, Q. Z., ZHANG, Y., JIANG, X. G. & FU, S. K. 2006a. Lectin-conjugated PEG-PLA nanoparticles: Preparation and brain delivery after intranasal administration. Biomaterials, 27, 3482-3490.

GAO, Y. G., YAN, X. Z., SONG, A. X., CHANG, Y. G., GAO, X. C., JIANG, N., ZHANG, Q. & HU, H. Y. 2006b. Structural insights into the specific binding of huntingtin proline-rich region with the SH3 and WW domains. Structure, 14, 1755-1765.

GARDINA, P. J., CLARK, T. A., SHIMADA, B., STAPLES, M. K., YANG, Q., VEITCH, J., SCHWEITZER, A., AWAD, T., SUGNET, C. & DEE, S. 2006. Alternative splicing and differential gene expression in colon cancer detected by a whole genome exon array. Bmc Genomics, 7, 325.

GARNEAU, D., REVIL, T., FISETTE, J. F. & CHABOT, B. 2005. Heterogeneous nuclear ribonucleoprotein F/H proteins modulate the alternative splicing of the apoptotic mediator Bcl-x. Journal of Biological Chemistry, 280, 22641-22650.

GAUTHIER, L. R., CHARRIN, B. C., BORRELL-PAGES, M., DOMPIERRE, J. P., RANGONE, H., CORDELIERES, F. P., DE MEY, J., MACDONALD, M. E., LESSMANN, V., HUMBERT, S. & SAUDOU, F. 2004. Huntingtin controls neurotrophic support and survival of neurons by enhancing BDNF vesicular transport along microtubules. Cell, 118, 127-138.

GERVAIS, F. G., SINGARAJA, R., XANTHOUDAKIS, S., GUTEKUNST, C. A., LEAVITT, B. R., METZLER, M., HACKAM, A. S., TAM, J., VAILLANCOURT, J. P., HOUTZAGER, V., RASPER, D. M., ROY, S., HAYDEN, M. R. & NICHOLSON, D. W. 2002. Recruitment and activation of caspase-8 by the Huntingtin-interacting protein Hip-1 and a novel partners Hippi. Nature Cell Biology, 4, 95-105.

GIL, J. M. & REGO, A. C. 2008. Mechanisms of neurodegeneration in Huntington's disease. European Journal of Neuroscience, 27, 2803-2820.

GINÉS, S., BOSCH, M., MARCO, S., GAVALDÀ, N., DÍAZ‐HERNÁNDEZ, M., LUCAS, J. J., CANALS, J. M. & ALBERCH, J. 2006. Reduced expression of the TrkB receptor in Huntington's disease mouse models and in human brain. European Journal of Neuroscience, 23, 649-658.

GINN, S. L., ALEXANDER, I. E., EDELSTEIN, M. L., ABEDI, M. R. & WIXON, J. 2013. Gene therapy clinical trials worldwide to 2012–an update. The journal of gene medicine, 15, 65-77.

GLATZEL, M. & AGUZZI, A. 2000. PrPC expression in the peripheral nervous system is a determinant of prion neuroinvasion. Journal of General Virology, 81, 2813-2821.

GONCALVES, V., MATOS, P. & JORDAN, P. 2008. The beta-catenin/TCF4 pathway modifies alternative splicing through modulation of SRp20 expression. Rna-a Publication of the Rna Society, 14, 2538-2549.

GRAFF, C. L. & POLLACK, G. M. 2003. P-glycoprotein attenuates brain uptake of substrates after nasal instillation. Pharmaceutical Research, 20, 1225-1230.

GRAHAM, F. L., SMILEY, J., RUSSELL, W. C. & NAIRN, R. 1977. CHARACTERISTICS OF A HUMAN CELL LINE TRANSFORMED BY DNA FROM HUMAN ADENOVIRUS TYPE-5. Journal of General Virology, 36, 59-72.

GRAMS, R. & KORGE, G. 1998. The mub gene encodes a protein containing three KH domains and is expressed in the mushroom bodies of Drosophila melanogaster. Gene, 215, 191-201.

GRAY, T. A., SMITHWICK, M. J., SCHALDACH, M. A., MARTONE, D. L., GRAVES, J. A. M., MCCARREY, J. R. & NICHOLLS, R. D. 1999. Concerted regulation and molecular evolution of the duplicated SNRPB '/B and SNRPN loci. Nucleic Acids Research, 27, 4577-4584.

GRISHOK, A. 2005. RNAi mechanisms in Caenorhabditis elegans. Febs Letters, 579, 5932-5939.

196

GROSS, E. A., SWENBERG, J. A., FIELDS, S. & POPP, J. A. 1982. COMPARATIVE MORPHOMETRY OF THE NASAL CAVITY IN RATS AND MICE. Journal of Anatomy, 135, 83-88.

GUSELLA, J. F. & MACDONALD, M. E. 2009. Huntington's disease: the case for genetic modifiers. Genome Med, 1, 80.

HAN, J. H., XIONG, J., WANG, D. & FU, X. D. 2011. Pre-mRNA splicing: where and when in the nucleus. Trends in Cell Biology, 21, 336-343.

HANSEN, M., HSU, A. L., DILLIN, A. & KENYON, C. 2005. New genes tied to endocrine, metabolic, and dietary regulation of lifespan from a Caenorhabditis elegans genomic RNAi screen. Plos Genetics, 1, 119-128.

HASEGAWA, Y., BROCKDORFF, N., KAWANO, S., TSUTUI, K., TSUTUI, K. & NAKAGAWA, S. 2010. The Matrix Protein hnRNP U Is Required for Chromosomal Localization of Xist RNA. Developmental Cell, 19, 469-476.

HASHIZUME, R., OZAWA, T., GRYAZNOV, S. M., BOLLEN, A. W., LAMBORN, K. R., FREY, W. H. & DEEN, D. F. 2008. New therapeutic approach for brain tumors: Intranasal delivery of telomerase inhibitor GRN163. Neuro-oncology, 10, 112-120.

HAY, D. G., SATHASIVAM, K., TOBABEN, S., STAHL, B., MARBER, M., MESTRIL, R., MAHAL, A., SMITH, D. L., WOODMAN, B. & BATES, G. P. 2004. Progressive decrease in chaperone protein levels in a mouse model of Huntington's disease and induction of stress proteins as a therapeutic approach. Human Molecular Genetics, 13, 1389-1405.

HAYCRAFT, C. J., SCHAFER, J. C., ZHANG, Q. H., TAULMAN, P. D. & YODER, B. K. 2003. Identification of CHE-13, a novel intraflagellar transport protein required for cilia formation. Experimental Cell Research, 284, 251-263.

HENG, M. Y., TALLAKSEN-GREENE, S. J., DETLOFF, P. J. & ALBIN, R. L. 2007. Longitudinal evaluation of the Hdh (CAG) 150 knock-in murine model of Huntington's disease. The Journal of neuroscience, 27, 8989-8998.

HICKEY, M. A. & CHESSELET, M.-F. 2011. Behavioral Assessment of Genetic Mouse Models of Huntington’s Disease. Animal Models of Movement Disorders. Springer.

HICKEY, M. A., KOSMALSKA, A., ENAYATI, J., COHEN, R., ZEITLIN, S., LEVINE, M. S. & CHESSELET, M.-F. 2008. Extensive early motor and non-motor behavioral deficits are followed by striatal neuronal loss in knock-in Huntington's disease mice. Neuroscience, 157, 280-295.

HILLIARD, M. A., APICELLA, A. J., KERR, R., SUZUKI, H., BAZZICALUPO, P. & SCHAFER, W. R. 2005. In vivo imaging of C-elegans ASH neurons: cellular response and adaptation to chemical repellents. Embo Journal, 24, 63-72.

HILLIARD, M. A., BARGMANN, C. I. & BAZZICALUPO, P. 2002. C-elegans responds to chemical repellents by integrating sensory inputs from the head and the tail. Current Biology, 12, 730-734.

HIPP, M. S., PATEL, C. N., BERSUKER, K., RILEY, B. E., KAISER, S. E., SHALER, T. A., BRANDEIS, M. & KOPITO, R. R. 2012. Indirect inhibition of 26S proteasome activity in a cellular model of Huntington's disease. Journal of Cell Biology, 196, 573-587.

HIROSE, Y. & MANLEY, J. L. 2000. RNA polymerase II and the integration of nuclear events. Genes & Development, 14, 1415-1429.

HIROSE, Y., TACKE, R. & MANLEY, J. L. 1999. Phosphorylated RNA polymerase II stimulates pre-mRNA splicing. Genes & Development, 13, 1234-1239.

HODGES, A., STRAND, A. D., ARAGAKI, A. K., KUHN, A., SENGSTAG, T., HUGHES, G., ELLISTON, L. A., HARTOG, C., GOLDSTEIN, D. R. & THU, D. 2006. Regional and cellular gene expression changes in human Huntington's disease brain. Human Molecular Genetics, 15, 965-977.

HOUSE, A. E. & LYNCH, K. W. 2006. An exonic splicing silencer represses spliceosome assembly after ATP-dependent exon recognition. Nature Structural & Molecular Biology, 13, 937-944.

HUANG, Y. Q. & STEITZ, J. A. 2005. SRprises along a messenger's journey. Molecular Cell, 17, 613-615.

197

HUGHES, R. E. 2002. Polyglutamine disease: Acetyltransferases awry. Current Biology, 12, R141-R143.

HUNEYCUTT, B., BI, Z., AOKI, C. & REISS, C. 1993. Central neuropathogenesis of vesicular stomatitis virus infection of immunodeficient mice. Journal of virology, 67, 6698-6706.

HUNTER, C. P. 1999. Genetics: A touch of elegance with RNAi. Current Biology, 9, R440-R442. ICHIMURA, T., FRASER, P. A. & CSERR, H. F. 1991. DISTRIBUTION OF EXTRACELLULAR TRACERS IN

PERIVASCULAR SPACES OF THE RAT-BRAIN. Brain Research, 545, 103-113. ILLUM, L. 2007. Nanoparticulate systems for nasal delivery of drugs: a real improvement over simple

systems? Journal of Pharmaceutical Sciences, 96, 473-483. IQBAL, M., LIN, W., JABBAL-GILL, I., DAVIS, S. S., STEWARD, M. W. & ILLUM, L. 2003. Nasal delivery of

chitosan-DNA plasmid expressing epitopes of respiratory syncytial virus (RSV) induces protective CTL responses in BALB/c mice. Vaccine, 21, 1478-1485.

JANA, N. R., TANAKA, M., WANG, G. H. & NUKINA, N. 2000. Polyglutamine length-dependent interaction of Hsp40 and Hsp70 family chaperones with truncated N-terminal huntingtin: their role in suppression of aggregation and cellular toxicity. Human Molecular Genetics, 9, 2009-2018.

JANSSON, B. & BJORK, E. 2002. Visualization of in vivo olfactory uptake and transfer using fluorescein dextran. Journal of Drug Targeting, 10, 379-386.

JAYARAMAN, M., MISHRA, R., KODALI, R., THAKUR, A. K., KOHARUDIN, L. M. I., GRONENBORN, A. M. & WETZEL, R. 2012. Kinetically Competing Huntingtin Aggregation Pathways Control Amyloid Polymorphism and Properties. Biochemistry, 51, 2706-2716.

JIA, Y., MU, J. C. & ACKERMAN, S. L. 2012. Mutation of a U2 snRNA gene causes global disruption of alternative splicing and neurodegeneration. Cell, 148, 296-308.

JIANG, H. B., POIRIER, M. A., LIANG, Y. D., PEI, Z., WEISKITTEL, C. E., SMITH, W. W., DEFRANCO, D. B. & ROSS, C. A. 2006. Depletion of CBP is directly caused by mutant huntingtin. Neurobiology of Disease, 23, 543-551.

JIANG, Y., WEI, N., ZHU, J., ZHAI, D., WU, L., CHEN, M., XU, G. & LIU, X. 2012. A new approach with less damage: intranasal delivery of tetracycline-inducible replication-defective herpes simplex virus type-1 vector to brain. Neuroscience, 201, 96-104.

JOHNSON, N. J., HANSON, L. R. & FREY, W. H. 2010. Trigeminal pathways deliver a low molecular weight drug from the nose to the brain and orofacial structures. Molecular pharmaceutics, 7, 884-893.

JONES, B. & ROBERTS, D. 1968. The quantitative measurement of motor inco‐ordination in naive mice using an accelerating rotarod. Journal of Pharmacy and Pharmacology, 20, 302-304.

JUMAA, H., WEI, G. & NIELSEN, P. J. 1999. Blastocyst formation is blocked in mouse embryos lacking the splicing factor SRp20. Current Biology, 9, 899-902.

KADENER, S., CRAMER, P., NOGUES, G., CAZALLA, D., DE LA MATA, M., FEDEDA, J. P., WERBAJH, S. E., SREBROW, A. & KORNBLIHTT, A. R. 2001. Antagonistic effects of T-Ag and VP16 reveal a role for RNA pol II elongation on alternative splicing. Embo Journal, 20, 5759-5768.

KALATHUR, R. K. R., HERNÁNDEZ-PRIETO, M. A. & FUTSCHIK, M. E. 2012. Huntington's Disease and its therapeutic target genes: a global functional profile based on the HD Research Crossroads database. BMC neurology, 12, 47.

KALETTA, T. & HENGARTNER, M. O. 2006. Finding function in novel targets: C-elegans as a model organism. Nature Reviews Drug Discovery, 5, 387-398.

KALSOTRA, A. & COOPER, T. A. 2011. Functional consequences of developmentally regulated alternative splicing. Nature Reviews Genetics, 12, 715-729.

KALTENBACH, L. S., ROMERO, E., BECKLIN, R. R., CHETTIER, R., BELL, R., PHANSALKAR, A., STRAND, A., TORCASSI, C., SAVAGE, J. & HURLBURT, A. 2007. Huntingtin interacting proteins are genetic modifiers of neurodegeneration. Plos Genetics, 3, e82.

198

KAMATH, R. S., MARTINEZ-CAMPOS, M., ZIPPERLEN, P., FRASER, A. G. & AHRINGER, J. 2001. Effectiveness of specific RNA-mediated interference through ingested double-stranded RNA in Caenorhabditis elegans. Genome biology, 2, RESEARCH0002.

KANAZAWA, T., AKIYAMA, F., KAKIZAKI, S., TAKASHIMA, Y. & SETA, Y. 2013. Delivery of siRNA to the brain using a combination of nose-to-brain delivery and cell-penetrating peptide-modified nano-micelles. Biomaterials, 34, 9220-9226.

KANAZAWA, T., MORISAKI, K., SUZUKI, S. & TAKASHIMA, Y. 2014. Prolongation of Life in Rats with Malignant Glioma by Intranasal siRNA/Drug Codelivery to the Brain with Cell-Penetrating Peptide-Modified Micelles. Molecular pharmaceutics, 11, 1471-1478.

KAR, K., JAYARAMAN, M., SAHOO, B., KODALI, R. & WETZEL, R. 2011. Critical nucleus size for disease-related polyglutamine aggregation is repeat-length dependent. Nature Structural & Molecular Biology, 18, 328-+.

KELLEY, N. W., HUANG, X. H., TAM, S., SPIESS, C., FRYDMAN, J. & PANDE, V. S. 2009. The Predicted Structure of the Headpiece of the Huntingtin Protein and Its Implications on Huntingtin Aggregation. Journal of Molecular Biology, 388, 919-927.

KENNEDY, L., EVANS, E., CHEN, C.-M., CRAVEN, L., DETLOFF, P. J., ENNIS, M. & SHELBOURNE, P. F. 2003. Dramatic tissue-specific mutation length increases are an early molecular event in Huntington disease pathogenesis. Human Molecular Genetics, 12, 3359-3367.

KENNEDY, S., WANG, D. & RUVKUN, G. 2004. A conserved siRNA-degrading RNase negatively regulates RNA interference in C-elegans. Nature, 427, 645-649.

KIM, D. Y., KIM, W., LEE, K. H., KIM, S. H., LEE, H. R., KIM, H. J., JUNG, Y., CHOI, J. H. & KIM, K. T. 2013. hnRNP Q regulates translation of p53 in normal and stress conditions. Cell Death and Differentiation, 20, 226-234.

KIM, I.-D., KIM, S.-W. & LEE, J.-K. 2009a. Gene knockdown in the olfactory bulb, amygdala, and

hypothalamus by intranasal siRNA administration. 대한해부학회지 제, 42. KIM, M. W., CHELLIAH, Y., KIM, S. W., OTWINOWSKI, Z. & BEZPROZVANNY, I. 2009b. Secondary

Structure of Huntingtin Amino-Terminal Region. Structure, 17, 1205-1212. KOBAYASHI, Y. & SOBUE, G. 2001. Protective effect of chaperones on polyglutamine diseases. Brain

Research Bulletin, 56, 165-168. KONDRASHOV, A. V., KIEFMANN, M., EBNET, K., KHANAM, T., MUDDASHETTY, R. S. & BROSIUS, J.

2005. Inhibitory effect of naked neural BC1 RNA or BC200 RNA on eukaryotic in vitro translation systems is reversed by poly(A)-binding protein (PABP). Journal of Molecular Biology, 353, 88-103.

KRISTENS.K & OLSSON, Y. 1971. UPTAKE OF EXOGENOUS PROTEINS IN MOUSE OLFACTORY CELLS. Acta Neuropathologica, 19, 145-&.

KUKALEV, A., NORD, Y., PALMBERG, C., BERGMAN, T. & PERCIPALLE, P. 2005. Actin and hnRNP U cooperate for productive transcription by RNA polymerase II. Nature Structural & Molecular Biology, 12, 238-244.

KULKARNI, K., BHAMBERE, T., CHAUDHARY, G., TALELE, S. & MOGHAL, R. 2013. Brain Targetting through Intranasal Route. Brain, 5, 1441-1450.

KWAN, W., TRAGER, U., DAVALOS, D., CHOU, A., BOUCHARD, J., ANDRE, R., MILLER, A., WEISS, A., GIORGINI, F., CHEAH, C., MOLLER, T., STELLA, N., AKASSOGLOU, K., TABRIZI, S. J. & MUCHOWSKI, P. J. 2012. Mutant huntingtin impairs immune cell migration in Huntington disease. Journal of Clinical Investigation, 122, 4737-4747.

LABBADIA, J. & MORIMOTO, R. I. 2013. Huntington's disease: underlying molecular mechanisms and emerging concepts. Trends in Biochemical Sciences, 38, 378-385.

LAKHANI, V. V., DING, F. & DOKHOLYAN, N. V. 2010. Polyglutamine Induced Misfolding of Huntingtin Exon1 is Modulated by the Flanking Sequences. Plos Computational Biology, 6.

LAKSO, M., VARTIAINEN, S., MOILANEN, A. M., SIRVIO, J., THOMAS, J. H., NASS, R., BLAKELY, R. D. & WONG, G. 2003. Dopaminergic neuronal loss and motor deficits in Caenorhabditis elegans overexpressing human alpha-synuclein. Journal of Neurochemistry, 86, 165-172.

199

LANDLES, C. & BATES, G. P. 2004. Huntingtin and the molecular pathogenesis of Huntington's disease. Fourth in molecular medicine review series. EMBO Reports, 5, 958-963.

LANE, M. A. & BAILEY, S. J. 2005. Role of retinoid signalling in the adult brain. Progress in Neurobiology, 75, 275-293.

LEAVITT, B. R., GUTTMAN, J. A., HODGSON, J. G., KIMEL, G. H., SINGARAJA, R., VOGL, A. W. & HAYDEN, M. R. 2001. Wild-type huntingtin reduces the cellular toxicity of mutant huntingtin in vivo. American Journal of Human Genetics, 68, 313-324.

LEE, E. B., LEE, V. M. Y. & TROJANOWSKI, J. Q. 2012. Gains or losses: molecular mechanisms of TDP43-mediated neurodegeneration. Nature Reviews Neuroscience, 13, 38-50.

LEE, M.-H., MORI, S. & RAYCHAUDHURI, P. 1996. trans-Activation by the hnRNP K protein involves an increase in RNA synthesis from the reporter genes. Journal of Biological Chemistry, 271, 3420-3427.

LEE, V. M. Y., GOEDERT, M. & TROJANOWSKI, J. Q. 2001. Neurodegenerative tauopathies. Annual Review of Neuroscience, 24, 1121-1159.

LEGLEITER, J., MITCHELL, E., LOTZ, G. P., SAPP, E., NG, C., DIFIGLIA, M., THOMPSON, L. M. & MUCHOWSKI, P. J. 2010. Mutant Huntingtin Fragments Form Oligomers in a Polyglutamine Length-dependent Manner in Vitro and in Vivo. Journal of Biological Chemistry, 285, 14777-14790.

LEJEUNE, F. X., MESROB, L., PARMENTIER, F., BICEP, C., VAZQUEZ-MANRIQUE, R. P., PARKER, J. A., VERT, J. P., TOURETTE, C. & NERI, C. 2012. Large-scale functional RNAi screen in C. elegans identifies genes that regulate the dysfunction of mutant polyglutamine neurons. Bmc Genomics, 13.

LEMIALE, F., KONG, W.-P., AKYÜREK, L. M., LING, X., HUANG, Y., CHAKRABARTI, B. K., ECKHAUS, M. & NABEL, G. J. 2003. Enhanced mucosal immunoglobulin A response of intranasal adenoviral vector human immunodeficiency virus vaccine and localization in the central nervous system. Journal of virology, 77, 10078-10087.

LEVINE, M. S., CEPEDA, C., HICKEY, M. A., FLEMING, S. M. & CHESSELET, M.-F. 2004. Genetic mouse models of Huntington's and Parkinson's diseases: illuminating but imperfect. Trends in neurosciences, 27, 691-697.

LI-KOROTKY, H. S., HEBDA, P. A., KELLY, L. A., LO, C. Y. & DOHAR, J. E. 2006. Identification of a pre-mRNA splicing factor, arginine/serine-rich 3 (Sfrs3), and its co-expression with fibronectin in fetal and postnatal rabbit airway mucosal and skin wounds. Biochimica Et Biophysica Acta-Molecular Basis of Disease, 1762, 34-45.

LI, C. 2001. Caenorhabditis elegans Nervous System. eLS. John Wiley & Sons, Ltd. LI, L.-B., YU, Z., TENG, X. & BONINI, N. M. 2008. RNA toxicity is a component of ataxin-3 degeneration

in Drosophila. Nature, 453, 1107-1111. LI, S. H., CHENG, A. L., ZHOU, H., LAM, S., RAO, M., LI, H. & LI, X. J. 2002. Interaction of Huntington

disease protein with transcriptional activator Sp1. Molecular and Cellular Biology, 22, 1277-1287.

LI, Y., FIELD, P. M. & RAISMAN, G. 2005. Olfactory ensheathing cells and olfactory nerve fibroblasts maintain continuous open channels for regrowth of olfactory nerve fibres. Glia, 52, 245-251.

LICATALOSI, D. D. & DARNELL, R. B. 2006. Splicing regulation in neurologic disease. Neuron, 52, 93-101.

LIN, C.-H., TALLAKSEN-GREENE, S., CHIEN, W.-M., CEARLEY, J. A., JACKSON, W. S., CROUSE, A. B., REN, S., LI, X.-J., ALBIN, R. L. & DETLOFF, P. J. 2001. Neurological abnormalities in a knock-in mouse model of Huntington’s disease. Human Molecular Genetics, 10, 137-144.

LIN, C. F., MOUNT, S. M., JARMOLOWSKI, A. & MAKALOWSKI, W. 2010. Evolutionary dynamics of U12-type spliceosomal introns. Bmc Evolutionary Biology, 10.

LIN, K., DORMAN, J. B., RODAN, A. & KENYON, C. 1997. daf-16: An HNF-3/forkhead family member that can function to double the life-span of Caenorhabditis elegans. Science, 278, 1319-1322.

200

LIN, K. T., LU, R. M. & TARN, W. Y. 2004. The WW domain-containing proteins interact with the early spliceosome and participate in pre-mRNA splicing in vivo. Molecular and Cellular Biology, 24, 9176-9185.

LINES, M. A., HUANG, L. J., SCHWARTZENTRUBER, J., DOUGLAS, S. L., LYNCH, D. C., BEAULIEU, C., GUION-ALMEIDA, M. L., ZECHI-CEIDE, R. M., GENER, B., GILLESSEN-KAESBACH, G., NAVA, C., BAUJAT, G., HORN, D., KINI, U., CALIEBE, A., ALANAY, Y., UTINE, G. E., LEV, D., KOHLHASE, J., GRIX, A. W., LOHMANN, D. R., HEHR, U., BOHM, D., MAJEWSKI, J., BULMAN, D. E., WIECZOREK, D., BOYCOTT, K. M. & CONSORTIUM, F. C. 2012. Haploinsufficiency of a Spliceosomal GTPase Encoded by EFTUD2 Causes Mandibulofacial Dysostosis with Microcephaly. American Journal of Human Genetics, 90, 369-377.

LINK, C. D. 1995. EXPRESSION OF HUMAN BETA-AMYLOID PEPTIDE IN TRANSGENIC CAENORHABDITIS-ELEGANS. Proceedings of the National Academy of Sciences of the United States of America, 92, 9368-9372.

LIONE, L. A., CARTER, R. J., HUNT, M. J., BATES, G. P., MORTON, A. J. & DUNNETT, S. B. 1999. Selective discrimination learning impairments in mice expressing the human Huntington's disease mutation. The Journal of neuroscience, 19, 10428-10437.

LIU, D., LI, C., CHEN, Y., BURNETT, C., LIU, X. Y., DOWNS, S., COLLINS, R. D. & HAWIGER, J. 2004. Nuclear import of proinflammatory transcription factors is required for massive liver apoptosis induced by bacterial lipopolysaccharide. Journal of Biological Chemistry, 279, 48434-48442.

LOCHHEAD, J. J. & THORNE, R. G. 2012. Intranasal delivery of biologics to the central nervous system. Advanced Drug Delivery Reviews, 64, 614-628.

LOPEZ, P. J. & SERAPHIN, B. 1999. Genomic-scale quantitative analysis of yeast pre-mRNA splicing: implications for splice-site recognition. Rna, 5, 1135-1137.

LUTHI-CARTER, R., STRAND, A., PETERS, N. L., SOLANO, S. M., HOLLINGSWORTH, Z. R., MENON, A. S., FREY, A. S., SPEKTOR, B. S., PENNEY, E. B. & SCHILLING, G. 2000. Decreased expression of striatal signaling genes in a mouse model of Huntington’s disease. Human Molecular Genetics, 9, 1259-1271.

MACDONALD, C. C., WILUSZ, J. & SHENK, T. 1994. The 64-kilodalton subunit of the CstF polyadenylation factor binds to pre-mRNAs downstream of the cleavage site and influences cleavage site location. Molecular and Cellular Biology, 14, 6647-6654.

MACLAUGHLIN, F. C., MUMPER, R. J., WANG, J. J., TAGLIAFERRI, J. M., GILL, I., HINCHCLIFFE, M. & ROLLAND, A. P. 1998. Chitosan and depolymerized chitosan oligomers as condensing carriers for in vivo plasmid delivery. Journal of Controlled Release, 56, 259-272.

MALLUCCI, G., DICKINSON, A., LINEHAN, J., KLÖHN, P.-C., BRANDNER, S. & COLLINGE, J. 2003. Depleting neuronal PrP in prion infection prevents disease and reverses spongiosis. Science, 302, 871-874.

MALLUCCI, G. R., WHITE, M. D., FARMER, M., DICKINSON, A., KHATUN, H., POWELL, A. D., BRANDNER, S., JEFFERYS, J. G. & COLLINGE, J. 2007. Targeting cellular prion protein reverses early cognitive deficits and neurophysiological dysfunction in prion-infected mice. Neuron, 53, 325-335.

MANDEL, M. & HIGA, A. 1970. CALCIUM-DEPENDENT BACTERIOPHAGE DNA INFECTION. Journal of Molecular Biology, 53, 159-&.

MANGIARINI, L., SATHASIVAM, K., SELLER, M., COZENS, B., HARPER, A., HETHERINGTON, C., LAWTON, M., TROTTIER, Y., LEHRACH, H. & DAVIES, S. W. 1996. Exon 1 of the< i> HD</i> Gene with an Expanded CAG Repeat Is Sufficient to Cause a Progressive Neurological Phenotype in Transgenic Mice. Cell, 87, 493-506.

MARTINEZ-CONTRERAS, R., FISETTE, J. F., NASIM, F. H., MADDEN, R., CORDEAU, M. & CHABOT, B. 2006. Intronic binding sites for hnRNP A/B and hnRNP F/H proteins stimulate pre-mRNA splicing. Plos Biology, 4, 172-185.

201

MASON, R. P., CASU, M., BUTLER, N., BREDA, C., CAMPESAN, S., CLAPP, J., GREEN, E. W., DHULKHED, D., KYRIACOU, C. P. & GIORGINI, F. 2013. Glutathione peroxidase activity is neuroprotective in models of Huntington's disease. Nature Genetics, 45, 1249-1254.

MATHISON, S., NAGILLA, R. & KOMPELLA, U. B. 1998. Nasal route for direct delivery of solutes to the central nervous system: Fact or fiction? Journal of Drug Targeting, 5, 415-441.

MATLIN, A. J., CLARK, F. & SMITH, C. W. J. 2005. Understanding alternative splicing: Towards a cellular code. Nature Reviews Molecular Cell Biology, 6, 386-398.

MAZAREI, G., NEAL, S. J., BECANOVIC, K., LUTHI-CARTER, R., SIMPSON, E. M. & LEAVITT, B. R. 2009. Expression analysis of novel striatal-enriched genes in Huntington disease. Human Molecular Genetics, ddp527.

MENALLED, L. B. 2005. Knock-in mouse models of Huntington's disease. NeuroRx, 2, 465-470. MENALLED, L. B. & CHESSELET, M.-F. 2002. Mouse models of Huntington's disease. Trends in

pharmacological sciences, 23, 32-39. MENALLED, L. B., SISON, J. D., WU, Y., OLIVIERI, M., LI, X.-J., LI, H., ZEITLIN, S. & CHESSELET, M.-F.

2002. Early motor dysfunction and striosomal distribution of huntingtin microaggregates in Huntington's disease knock-in mice. The Journal of neuroscience, 22, 8266-8276.

METZLER, M., CHEN, N. S., HELGASON, C. D., GRAHAM, R. K., NICHOL, K., MCCUTCHEON, K., NASIR, J., HUMPHRIES, R. K., RAYMOND, L. A. & HAYDEN, M. R. 1999. Life without Huntington: Normal differentiation into functional neurons. Journal of Neurochemistry, 72, 1009-1018.

MICHELOTTI, E. F., MICHELOTTI, G. A., ARONSOHN, A. I. & LEVENS, D. 1996. Heterogeneous nuclear ribonucleoprotein K is a transcription factor. Molecular and Cellular Biology, 16, 2350-2360.

MIKULA, M., DZWONEK, A., KARCZMARSKI, J., RUBEL, T., DADLEZ, M., WYRWICZ, L. S., BOMSZTYK, K. & OSTROWSKI, J. 2006. Landscape of the hnRNP K protein-protein interactome. Proteomics, 6, 2395-2406.

MILLER, B. R., DORNER, J. L., SHOU, M., SARI, Y., BARTON, S. J., SENGELAUB, D. R., KENNEDY, R. T. & REBEC, G. V. 2008. Up-regulation of GLT1 expression increases glutamate uptake and attenuates the Huntington's disease phenotype in the R6/2 mouse. Neuroscience, 153, 329-337.

MILLS, J. D. & JANITZ, M. 2012a. Alternative splicing of mRNA in the molecular pathology of neurodegenerative diseases. Neurobiology of Aging, 33, 1012. e11-1012. e24.

MILLS, J. D. & JANITZ, M. 2012b. Alternative splicing of mRNA in the molecular pathology of neurodegenerative diseases. Neurobiology of Aging, 33.

MIN, H. S., CHAN, R. C. & BLACK, D. L. 1995. THE GENERALLY EXPRESSED HNRNP-F IS INVOLVED IN A NEURAL-SPECIFIC PRE-MESSENGER-RNA SPLICING EVENT. Genes & Development, 9, 2659-2671.

MITRA, S. & FINKBEINER, S. 2008. The ubiquitin-proteasome pathway in Huntington's disease. Thescientificworldjournal, 8, 421-433.

MONTGOMERY, M. K. & FIRE, A. 1998. Double-stranded RNA as a mediator in sequence-specific genetic silencing and co-suppression. Trends in Genetics, 14, 255-258.

MORENO, J. A., RADFORD, H., PERETTI, D., STEINERT, J. R., VERITY, N., MARTIN, M. G., HALLIDAY, M., MORGAN, J., DINSDALE, D. & ORTORI, C. A. 2012. Sustained translational repression by eIF2 [agr]-P mediates prion neurodegeneration. Nature, 485, 507-511.

MORLEY, J. F., BRIGNULL, H. R., WEYERS, J. J. & MORIMOTO, R. I. 2002. The threshold for polyglutamine-expansion protein aggregation and cellular toxicity is dynamic and influenced by aging in Caenorhabditis elegans. Proceedings of the National Academy of Sciences of the United States of America, 99, 10417-10422.

MUCHOWSKI, P. J. & WACKER, J. L. 2005. Modulation of neurodegeneration by molecular chaperones. Nature Reviews Neuroscience, 6, 11-22.

MULLIS, K. B. & FALOONA, F. A. 1987. SPECIFIC SYNTHESIS OF DNA INVITRO VIA A POLYMERASE-CATALYZED CHAIN-REACTION. Methods in Enzymology, 155, 335-350.

202

MURPHY, K. P., CARTER, R. J., LIONE, L. A., MANGIARINI, L., MAHAL, A., BATES, G. P., DUNNETT, S. B. & MORTON, A. J. 2000. Abnormal synaptic plasticity and impaired spatial cognition in mice transgenic for exon 1 of the human Huntington's disease mutation. The Journal of neuroscience, 20, 5115-5123.

MUS, E., HOF, P. R. & TIEDGE, H. 2007. Dendritic BC200 RNA in aging and in Alzheimer's disease. Proceedings of the National Academy of Sciences of the United States of America, 104, 10679-10684.

MUSUNURU, K. & DARNELL, R. B. 2001. Paraneoplastic neurologic disease antigens: RNA-binding proteins and signaling proteins in neuronal degeneration. Annual Review of Neuroscience, 24, 239-262.

MYKOWSKA, A., SOBCZAK, K., WOJCIECHOWSKA, M., KOZLOWSKI, P. & KRZYZOSIAK, W. J. 2011. CAG repeats mimic CUG repeats in the misregulation of alternative splicing. Nucleic Acids Research, 39, 8938-8951.

NAARMANN, I. S., HARNISCH, C., FLACH, N., KREMMER, E., KUHN, H., OSTARECK, D. H. & OSTARECK-LEDERER, A. 2008. mRNA silencing in human erythroid cell maturation - Heterogeneous nuclear ribonucleoprotein K controls the expression of its regulator c-Src. Journal of Biological Chemistry, 283, 18461-18472.

NANOU, A. & AZZOUZ, M. 2009. Gene therapy for neurodegenerative diseases based on lentiviral vectors. Progress in brain research, 175, 187-200.

NASIM, M. T. & EPERON, I. C. 2006. A double-reporter splicing assay for determining splicing efficiency in mammalian cells. Nature Protocols, 1, 1022-1028.

NASIR, J., FLORESCO, S., O'KUSKY, J., DIEWERT, V., RICHMAN, J., ZEISLER, J., BOROWSKI, A., MCTZLER, M., GRAHAM, R., MARTH, J., PHILLIPS, A. & HAYDEN, M. 1995a. Targeted disruption of the Huntington's disease gene results in embryonic lethality in homozygotes and behavioral and neuropathological defects in heterozygotes. American Journal of Human Genetics, 57, A51.

NASIR, J., FLORESCO, S. B., O'KUSKY, J. R., DIEWERT, V. M., RICHMAN, J. M., ZEISLER, J., BOROWSKI, A., MARTH, J. D., PHILLIPS, A. G. & HAYDEN, M. R. 1995b. Targeted disruption of the Huntington's disease gene results in embryonic lethality and behavioral and morphological changes in heterozygotes. Cell, 81, 811-823.

NEEF, D. W., JAEGER, A. M. & THIELE, D. J. 2011. Heat shock transcription factor 1 as a therapeutic target in neurodegenerative diseases. Nature Reviews Drug Discovery, 10, 930-944.

NEEF, D. W., TURSKI, M. L. & THIELE, D. J. 2010. Modulation of Heat Shock Transcription Factor 1 as a Therapeutic Target for Small Molecule Intervention in Neurodegenerative Disease. Plos Biology, 8.

NELSON, R., SAWAYA, M. R., BALBIRNIE, M., MADSEN, A. O., RIEKEL, C., GROTHE, R. & EISENBERG, D. 2005. Structure of the cross-beta spine of amyloid-like fibrils. Nature, 435, 773-778.

NEWCOMER, M. E. 2001. Trading places. Nature Structural Biology, 8, 282-284. NI, J. Z., GRATE, L., DONOHUE, J. P., PRESTON, C., NOBIDA, N., O'BRIEN, G., SHIUE, L., CLARK, T. A.,

BLUME, J. E. & ARES, M. 2007. Ultraconserved elements are associated with homeostatic control of splicing regulators by alternative splicing and nonsense-mediated decay. Genes & Development, 21, 708-718.

NOLLEN, E. A. A., GARCIA, S. M., VAN HAAFTEN, G., KIM, S., CHAVEZ, A., MORIMOTO, R. I. & PLASTERK, R. H. A. 2004. Genome-wide RNA interference screen identifies previously undescribed regulators of polyglutamine aggregation. Proceedings of the National Academy of Sciences of the United States of America, 101, 6403-6408.

NONAKA, N., FARR, S. A., KAGEYAMA, H., SHIODA, S. & BANKS, W. A. 2008. Delivery of galanin-like peptide to the brain: targeting with intranasal delivery and cyclodextrins. Journal of Pharmacology and Experimental Therapeutics, 325, 513-519.

NYBAKKEN, K., VOKES, S. A., LIN, T. Y., MCMAHON, A. P. & PERRIMON, N. 2005. A genome-wide RNA interference screen in Drosophila melanogaster cells for new components of the Hh signaling pathway. Nature Genetics, 37, 1323-1332.

203

OKONIEWSKI, M. J., HEY, Y., PEPPER, S. D. & MILLER, C. J. 2007. High correspondence between Affymetrix exon and standard expression arrays. Biotechniques, 42, 181.

OSTROWSKI, J., KAWATA, Y., SCHULLERY, D. S., DENISENKO, O. N. & BOMSZTYK, K. 2003. Transient recruitment of the hnRNP K protein to inducibly transcribed gene loci. Nucleic Acids Research, 31, 3954-3962.

OU, S., WU, F., HARRICH, D., GARCÍA-MARTÍNEZ, L. F. & GAYNOR, R. B. 1995. Cloning and characterization of a novel cellular protein, TDP-43, that binds to human immunodeficiency virus type 1 TAR DNA sequence motifs. Journal of virology, 69, 3584-3596.

OWENS, D. R., ZINMAN, B. & BOLLI, G. 2003. Alternative routes of insulin delivery. Diabetic Medicine, 20, 886-898.

PADOWSKI, J. M. & POLLACK, G. M. 2010. Examination of the Ability of the Nasal Administration Route to Confer a Brain Exposure Advantage for Three Chemical Inhibitors of P-Glycoprotein. Journal of Pharmaceutical Sciences, 99, 3226-3233.

PAGANI, F. & BARALLE, F. E. 2004. Genomic variants in exons and introns: identifying the splicing spoilers. Nature Reviews Genetics, 5, 389-U2.

PALAZZOLO, I., NEDELSKY, N. B., ASKEW, C. E., HARMISON, G. G., KASANTSEV, A. G., TAYLOR, J. P., FISCHBECK, K. H. & PENNUTO, M. 2010. B2 Attenuates Polyglutamine-Expanded Androgen Receptor Toxicity in Cell and Fly Models of Spinal and Bulbar Muscular Atrophy. Journal of Neuroscience Research, 88, 2207-2216.

PALIDWOR, G. A., SHCHERBININ, S., HUSKA, M. R., RASKO, T., STELZL, U., ARUMUGHAN, A., FOUIIE, R., PORRAS, P., SANCHEZ-PULIDO, L., WANKER, E. E. & ANDRADE-NAVARRO, M. A. 2009. Detection of Alpha-Rod Protein Repeats Using a Neural Network and Application to Huntingtin. Plos Computational Biology, 5.

PARKER, R. & SHETH, U. 2007. P bodies and the control of mRNA translation and degradation. Molecular Cell, 25, 635-646.

PASSANI, L. A., BEDFORD, M. T., FABER, P. W., MCGINNIS, K. M., SHARP, A. H., GUSELLA, J. F., VONSATTEL, J.-P. & MACDONALD, M. E. 2000. Huntingtin’s WW domain partners in Huntington’s disease post-mortem brain fulfill genetic criteria for direct involvement in Huntington’s disease pathogenesis. Human Molecular Genetics, 9, 2175-2182.

PATHAN, S. A., IQBAL, Z., ZAIDI, S. M. A., TALEGAONKAR, S., VOHRA, D., JAIN, G. K., AZEEM, A., JAIN, N., LALANI, J. R., KHAR, R. K. & AHMAD, F. J. 2009. CNS drug delivery systems: novel approaches. Recent patents on drug delivery & formulation, 3, 71-89.

PAULSEN, R. D., SONI, D. V., WOLLMAN, R., HAHN, A. T., YEE, M. C., GUAN, A., HESLEY, J. A., MILLER, S. C., CROMWELL, E. F., SOLOW-CORDERO, D. E., MEYER, T. & CIMPRICH, K. A. 2009. A Genome-wide siRNA Screen Reveals Diverse Cellular Processes and Pathways that Mediate Genome Stability. Molecular Cell, 35, 228-239.

PAUSHKIN, S., GUBITZ, A. K., MASSENET, S. & DREYFUSS, G. 2002. The SMN complex, an assemblyosome of ribonucleoproteins. Current Opinion in Cell Biology, 14, 305-312.

PERUTZ, M. F., JOHNSON, T., SUZUKI, M. & FINCH, J. T. 1994. GLUTAMINE REPEATS AS POLAR ZIPPERS - THEIR POSSIBLE ROLE IN INHERITED NEURODEGENERATIVE DISEASES. Proceedings of the National Academy of Sciences of the United States of America, 91, 5355-5358.

PFAFFL, M. W. 2001. A new mathematical model for relative quantification in real-time RT-PCR. Nucleic Acids Research, 29.

PIRES, A., FORTUNA, A., ALVES, G. & FALCAO, A. 2009. Intranasal Drug Delivery: How, Why and What for? Journal of Pharmacy and Pharmaceutical Sciences, 12, 288-311.

POIRIER, M. A., JIANG, H. & ROSS, C. A. 2005. A structure-based analysis of huntingtin mutant polyglutamine aggregation and toxicity: evidence for a compact beta-sheet structure. Human Molecular Genetics, 14, 765-774.

POLLING, S., HILL, A. F. & HATTERS, D. M. 2012. Polyglutamine aggregation in Huntington and related diseases. Advances in experimental medicine and biology, 769, 125-40.

204

POLYMENIDOU, M., LAGIER-TOURENNE, C., HUTT, K. R., HUELGA, S. C., MORAN, J., LIANG, T. Y., LING, S. C., SUN, E., WANCEWICZ, E., MAZUR, C., KORDASIEWICZ, H., SEDAGHAT, Y., DONOHUE, J. P., SHIUE, L., BENNETT, C. F., YEO, G. W. & CLEVELAND, D. W. 2011. Long pre-mRNA depletion and RNA missplicing contribute to neuronal vulnerability from loss of TDP-43. Nature Neuroscience, 14, 459-U92.

PRUSINER, S. B. 1991. Molecular biology of prion diseases. Science, 252, 1515-1522. RAJAN, R. S., ILLING, M. E., BENCE, N. F. & KOPITO, R. R. 2001. Specificity in intracellular protein

aggregation and inclusion body formation. Proceedings of the National Academy of Sciences of the United States of America, 98, 13060-13065.

RALPH, G., BINLEY, K., WONG, L., AZZOUZ, M. & MAZARAKIS, N. 2006. Gene therapy for neurodegenerative and ocular diseases using lentiviral vectors. Clinical Science, 110, 37-46.

RALPH, G. S., RADCLIFFE, P. A., DAY, D. M., CARTHY, J. M., LEROUX, M. A., LEE, D. C., WONG, L.-F., BILSLAND, L. G., GREENSMITH, L. & KINGSMAN, S. M. 2005. Silencing mutant SOD1 using RNAi protects against neurodegeneration and extends survival in an ALS model. Nature Medicine, 11, 429-433.

RAO, D. D., VORHIES, J. S., SENZER, N. & NEMUNAITIS, J. 2009. siRNA vs. shRNA: similarities and differences. Advanced Drug Delivery Reviews, 61, 746-759.

RAOUL, C., ABBAS-TERKI, T., BENSADOUN, J.-C., GUILLOT, S., HAASE, G., SZULC, J., HENDERSON, C. E. & AEBISCHER, P. 2005. Lentiviral-mediated silencing of SOD1 through RNA interference retards disease onset and progression in a mouse model of ALS. Nature Medicine, 11, 423-428.

RAYMOND, L. A., ANDRE, V. M., CEPEDA, C., GLADDING, C. M., MILNERWOOD, A. J. & LEVINE, M. S. 2011. PATHOPHYSIOLOGY OF HUNTINGTON'S DISEASE: TIME-DEPENDENT ALTERATIONS IN SYNAPTIC AND RECEPTOR FUNCTION. Neuroscience, 198, 252-273.

REDDY, G., STRAUBB, J. E. & THIRUMALAI, D. 2009. Dynamics of locking of peptides onto growing amyloid fibrils. Proceedings of the National Academy of Sciences of the United States of America, 106, 11948-11953.

REGER, M. A., WATSON, G. S., GREEN, P. S., BAKER, L. D., CHOLERTON, B., FISHEL, M. A., PLYMATE, S. R., CHERRIER, M. M., SCHELLENBERG, G. D., FREY, W. H. & CRAFT, S. 2008. Intranasal insulin administration dose-dependently modulates verbal memory and plasma amyloid-beta in memory-impaired older adults. Journal of Alzheimers Disease, 13, 323-331.

REISS, C. S., PLAKHOV, I. V. & KOMATSU, T. 1998. Viral Replication in Olfactory Receptor Neurons and Entry into the Olfactory Bulb and Braina. Annals of the New York Academy of Sciences, 855, 751-761.

REN, Y.-G., WAGNER, K. W., KNEE, D. A., AZA-BLANC, P., NASOFF, M. & DEVERAUX, Q. L. 2004. Differential regulation of the TRAIL death receptors DR4 and DR5 by the signal recognition particle. Molecular biology of the cell, 15, 5064-5074.

RENNER, D. B., FREY II, W. H. & HANSON, L. R. 2012. Intranasal delivery of siRNA to the olfactory bulbs of mice< i> via</i> the olfactory nerve pathway. Neuroscience letters, 513, 193-197.

REVIL, T., PELLETIER, J., TOUTANT, J., CLOUTIER, A. & CHABOT, B. 2009. Heterogeneous Nuclear Ribonucleoprotein K Represses the Production of Pro-apoptotic Bcl-x(S) Splice Isoform. Journal of Biological Chemistry, 284, 21458-21467.

RIDLEY, R. M., FRITH, C. D., CROW, T. J. & CONNEALLY, P. M. 1988. ANTICIPATION IN HUNTINGTONS-DISEASE IS INHERITED THROUGH THE MALE LINE BUT MAY ORIGINATE IN THE FEMALE. Journal of Medical Genetics, 25, 589-595.

ROBERTS, W. K. & DARNELL, R. B. 2004. Neuroimmunology of the paraneoplastic neurological degenerations. Current Opinion in Immunology, 16, 616-622.

ROCKABRAND, E., SLEPKO, N., PANTALONE, A., NUKALA, V. N., KAZANTSEV, A., MARSH, J. L., SULLIVAN, P. G., STEFFAN, J. S., SENSI, S. L. & THOMPSON, L. M. 2007. The first 17 amino acids of Huntingtin modulate its sub-cellular localization, aggregation and effects on calcium homeostasis. Human Molecular Genetics, 16, 61-77.

205

ROSHON, M. J. & RULEY, H. E. 2005. Hypomorphic mutation in hnRNP U results in post-implantation lethality. Transgenic Research, 14, 179-192.

ROSICARELLI, B., SERAFINI, B., SBRICCOLI, M., LU, M., CARDONE, F., POCCHIARI, M. & ALOISI, F. 2005. Migration of dendritic cells into the brain in a mouse model of prion disease. Journal of neuroimmunology, 165, 114-120.

ROSKOSKI, R. 2004. Src protein-tyrosine kinase structure and regulation. Biochemical and Biophysical Research Communications, 324, 1155-1164.

ROSS, C. A. & POIRIER, M. A. 2004. Protein aggregation and neurodegenerative disease. ROYCHOUDHURY, P. & CHAUDHURI, K. 2007. Evidence for heterogeneous nuclear ribonucleoprotein

K overexpression in oral squamous cell carcinoma. British Journal of Cancer, 97, 574-575. RUBINSZTEIN, D. C. 2002. Lessons from animal models of Huntington's disease. Trends in Genetics,

18, 202-209. SANCHEZ, I., MAHLKE, C. & YUAN, J. Y. 2003. Pivotal role of oligomerization in expanded

polyglutamine neurodegenerative disorders. Nature, 421, 373-379. SANFORD, J. R., GRAY, N. K., BECKMANN, K. & CACERES, J. F. 2004. A novel role for shuttling SR

proteins in mRNA translation. Genes & Development, 18, 755-768. SATHASIVAM, K., LANE, A., LEGLEITER, J., WARLEY, A., WOODMAN, B., FINKBEINER, S., PAGANETTI,

P., MUCHOWSKI, P. J., WILSON, S. & BATES, G. P. 2010. Identical oligomeric and fibrillar structures captured from the brains of R6/2 and knock-in mouse models of Huntington's disease. Human Molecular Genetics, 19, 65-78.

SATHASIVAM, K., NEUEDER, A., GIPSON, T. A., LANDLES, C., BENJAMIN, A. C., BONDULICH, M. K., SMITH, D. L., FAULL, R. L. M., ROOS, R. A. C., HOWLAND, D., DETLOFF, P. J., HOUSMAN, D. E. & BATES, G. P. 2013. Aberrant splicing of HTT generates the pathogenic exon 1 protein in Huntington disease. Proceedings of the National Academy of Sciences of the United States of America, 110, 2366-2370.

SAYDOFF, J. A., GARCIA, R. A., BROWNE, S. E., LIU, L., SHENG, J., BRENNEMAN, D., HU, Z., CARDIN, S., GONZALEZ, A. & VON BORSTEL, R. W. 2006. Oral uridine pro-drug PN401 is neuroprotective in the R6/2 and N171-82Q mouse models of Huntington's disease. Neurobiology of Disease, 24, 455-465.

SCHILLING, G., BECHER, M. W., SHARP, A. H., JINNAH, H. A., DUAN, K., KOTZUK, J. A., SLUNT, H. H., RATOVITSKI, T., COOPER, J. K. & JENKINS, N. A. 1999. Intranuclear inclusions and neuritic aggregates in transgenic mice expressing a mutant N-terminal fragment of huntingtin. Human Molecular Genetics, 8, 397-407.

SCHWAB, C., ARAI, T., HASEGAWA, M., YU, S. & MCGEER, P. L. 2008. Colocalization of Transactivation-Responsive DNA-Binding Protein 43 and Huntingtin in Inclusions of Huntington Disease. Journal of Neuropathology and Experimental Neurology, 67, 1159-1165.

SCRANTON, R. A., FLETCHER, L., SPRAGUE, S., JIMENEZ, D. F. & DIGICAYLIOGLU, M. 2011. The Rostral Migratory Stream Plays a Key Role in Intranasal Delivery of Drugs into the CNS. Plos One, 6.

SEPHTON, C. F., CENIK, B., CENIK, B. K., HERZ, J. & YU, G. 2012. TDP-43 in central nervous system development and function: clues to TDP-43-associated neurodegeneration. Biological Chemistry, 393, 589-594.

SEPHTON, C. F., GOOD, S. K., ATKIN, S., DEWEY, C. M., MAYER, P., HERZ, J. & YU, G. 2010. TDP-43 Is a Developmentally Regulated Protein Essential for Early Embryonic Development. Journal of Biological Chemistry, 285, 6826-6834.

SHAO, Z., KRISHNAMOORTHY, R. & MITRA, A. K. 1992. Cyclodextrins as nasal absorption promoters of insulin: mechanistic evaluations. Pharmaceutical Research, 9, 1157-1163.

SHAW, G., MORSE, S., ARARAT, M. & GRAHAM, F. L. 2002. Preferential transformation of human neuronal cells by human adenoviruses and the origin of HEK 293 cells. Faseb Journal, 16, 869-+.

SHELBOURNE, P. F., KILLEEN, N., HEVNER, R. F., JOHNSTON, H. M., TECOTT, L., LEWANDOSKI, M., ENNIS, M., RAMIREZ, L., LI, Z. & IANNICOLA, C. 1999. A Huntington's disease CAG expansion

206

at the murine Hdh locus is unstable and associated with behavioural abnormalities in mice. Human Molecular Genetics, 8, 763-774.

SHEN, H. H. & GREEN, M. R. 2004. A pathway of sequential arginine-serine-rich domain-splicing signal interactions during mammalian spliceosome assembly. Molecular Cell, 16, 363-373.

SIJEN, T., FLEENOR, J., SIMMER, F., THIJSSEN, K. L., PARRISH, S., TIMMONS, L., PLASTERK, R. H. A. & FIRE, A. 2001. On the role of RNA amplification in dsRNA-triggered gene silencing. Cell, 107, 465-476.

SIMMER, F., MOORMAN, C., VAN DER LINDEN, A. M., KUIJK, E., VAN DEN BERGHE, P. V., KAMATH, R. S., FRASER, A. G., AHRINGER, J. & PLASTERK, R. H. 2003. Genome-wide RNAi of C. elegans using the hypersensitive rrf-3 strain reveals novel gene functions. Plos Biology, 1, e12.

SIMMER, F., TIJSTERMAN, M., PARRISH, S., KOUSHIKA, S. P., NONET, M. L., FIRE, A., AHRINGER, J. & PLASTERK, R. H. 2002. Loss of the Putative RNA-Directed RNA Polymerase RRF-3 Makes< i> C</i>.< i> elegans</i> Hypersensitive to RNAi. Current Biology, 12, 1317-1319.

SINGH, R. K. & COOPER, T. A. 2012. Pre-mRNA splicing in disease and therapeutics. Trends in Molecular Medicine, 18, 472-482.

SLEPKO, N., BHATTACHARYYA, A. M., JACKSON, G. R., STEFFAN, J. S., MARSH, J. L., THOMPSON, L. M. & WETZEL, R. 2006. Normal-repeat-length polyglutamine peptides accelerate aggregation nucleation and cytotoxicity of expanded polyglutamine proteins. Proceedings of the National Academy of Sciences of the United States of America, 103, 14367-14372.

SLOW, E. J., VAN RAAMSDONK, J., ROGERS, D., COLEMAN, S. H., GRAHAM, R. K., DENG, Y., OH, R., BISSADA, N., HOSSAIN, S. M. & YANG, Y.-Z. 2003. Selective striatal neuronal loss in a YAC128 mouse model of Huntington disease. Human Molecular Genetics, 12, 1555-1567.

SMALL, E. C., LEGGETT, S. R., WINANS, A. A. & STALEY, J. P. 2006. The EF-G-like GTPase Snu114p regulates spliceosome dynamics mediated by Brr2p, a DExD/H box ATPase. Molecular Cell, 23, 389-399.

SONKINA, S., TUKHFATULLINA, I. I., BENSENY‐CASES, N., IONOV, M., BRYSZEWSKA, M., SALAKHUTDINOV, B. A. & CLADERA, J. 2010. Interaction of the prion protein fragment PrP 185–206 with biological membranes: effect on membrane permeability. Journal of Peptide Science, 16, 342-348.

SONTAG, E. M., LOTZ, G. P., AGRAWAL, N., TRAN, A., ARON, R., YANG, G., NECULA, M., LAU, A., FINKBEINER, S., GLABE, C., MARSH, J. L., MUCHOWSKI, P. J. & THOMPSON, L. M. 2012. Methylene Blue Modulates Huntingtin Aggregation Intermediates and Is Protective in Huntington's Disease Models. Journal of Neuroscience, 32, 11109-11119.

SOTO, C. 2003. Unfolding the role of protein misfolding in neurodegenerative diseases. Nature Reviews Neuroscience, 4, 49-60.

STACK, E. C., KUBILUS, J. K., SMITH, K., CORMIER, K., DEL SIGNORE, S. J., GUELIN, E., RYU, H., HERSCH, S. M. & FERRANTE, R. J. 2005. Chronology of behavioral symptoms and neuropathological sequela in R6/2 Huntington's disease transgenic mice. Journal of Comparative Neurology, 490, 354-370.

STEELE, A. D., JACKSON, W. S., KING, O. D. & LINDQUIST, S. 2007. The power of automated high-resolution behavior analysis revealed by its application to mouse models of Huntington's and prion diseases. Proceedings of the National Academy of Sciences, 104, 1983-1988.

STEEN, E., TERRY, B. M., RIVERA, E. J., CANNON, J. L., NEELY, T. R., TAVARES, R., XU, X. J., WANDS, J. R. & DE LA MONTE, S. M. 2005. Impaired insulin and insulin-like growth factor expression and signaling mechanisms in Alzheimer's disease - is this type 3 diabetes? Journal of Alzheimers Disease, 7, 63-80.

STOILOV, P., CASTREN, E. & STAMM, S. 2002. Analysis of the human TrkB gene genomic organization reveals novel TrkB isoforms, unusual gene length, and splicing mechanism. Biochemical and Biophysical Research Communications, 290, 1054-1065.

207

STRONG, M. J., VOLKENING, K., HAMMOND, R., YANG, W., STRONG, W., LEYSTRA-LANTZ, C. & SHOESMITH, C. 2007. TDP43 is a human low molecular weight neurofilament (< i> h</i> NFL) mRNA-binding protein. Molecular and Cellular Neuroscience, 35, 320-327.

SUN, Y., SAVANENIN, A., REDDY, P. H. & LIU, Y. F. 2001. Polyglutamine-expanded Huntingtin promotes sensitization of N-methyl-D-aspartate receptors via post-synaptic density 95. Journal of Biological Chemistry, 276, 24713-24718.

SUTHERLAND, G. T., JANITZ, M. & KRIL, J. J. 2011. Understanding the pathogenesis of Alzheimer’s disease: will RNA‐Seq realize the promise of transcriptomics? Journal of Neurochemistry, 116, 937-946.

TABARA, H., YIGIT, E., SIOMI, H. & MELLO, C. C. 2002. The dsRNA binding protein RDE-4 interacts with RDE-1, DCR-1, and a DExX-box helicase to direct RNAi in C-elegans. Cell, 109, 861-871.

TAM, S., SPIESS, C., AUYEUNG, W., JOACHIMIAK, L., CHEN, B., POIRIER, M. A. & FRYDMAN, J. 2009. The chaperonin TRiC blocks a huntingtin sequence element that promotes the conformational switch to aggregation. Nature Structural & Molecular Biology, 16, 1279-U98.

TANAKA, K., KANAZAWA, T., SHIBATA, Y., SUDA, Y., FUKUDA, T., TAKASHIMA, Y. & OKADA, H. 2010. Development of cell-penetrating peptide-modified MPEG-PCL diblock copolymeric nanoparticles for systemic gene delivery. International Journal of Pharmaceutics, 396, 229-238.

TANAKA, M., MACHIDA, Y., NIU, S. Y., IKEDA, T., JANA, N. R., DOI, H., KUROSAWA, M., NEKOOKI, M. & NUKINA, N. 2004. Trehalose alleviates polyglutamine-mediated pathology in a mouse model of Huntington disease. Nature Medicine, 10, 148-154.

TANG, Y., HORIKAWA, I., AJIRO, M., ROBLES, A. I., FUJITA, K., MONDAL, A. M., STAUFFER, J. K., ZHENG, Z. M. & HARRIS, C. C. 2013. Downregulation of splicing factor SRSF3 induces p53 beta, an alternatively spliced isoform of p53 that promotes cellular senescence. Oncogene, 32, 2792-2798.

TAUFFENBERGER, A., CHITRAMUTHU, B. P., BATEMAN, A., BENNETT, H. P. J. & PARKER, J. A. 2013. Reduction of polyglutamine toxicity by TDP-43, FUS and progranulin in Huntington's disease models. Human Molecular Genetics, 22, 782-794.

THAKUR, A. K., JAYARAMAN, M., MISHRA, R., THAKUR, M., CHELLGREN, V. M., BYEON, I. J. L., ANJUM, D. H., KODALI, R., CREAMER, T. P., CONWAY, J. F., GRONENBORN, A. M. & WETZEL, R. 2009. Polyglutamine disruption of the huntingtin exon 1 N terminus triggers a complex aggregation mechanism. Nature Structural & Molecular Biology, 16, 380-389.

THORNE, R. G., EMORY, C. R., ALA, T. A. & FREY, W. H. 1995. QUANTITATIVE-ANALYSIS OF THE OLFACTORY PATHWAY FOR DRUG-DELIVERY TO THE BRAIN. Brain Research, 692, 278-282.

THORNE, R. G., PRONK, G. J., PADMANABHAN, V. & FREY, W. H. 2004. Delivery of insulin-like growth factor-I to the rat brain and spinal cord along olfactory and trigeminal pathways following intranasal administration. Neuroscience, 127, 481-496.

THORSEN, K., SØRENSEN, K. D., BREMS-ESKILDSEN, A. S., MODIN, C., GAUSTADNES, M., HEIN, A.-M. K., KRUHØFFER, M., LAURBERG, S., BORRE, M. & WANG, K. 2008. Alternative splicing in colon, bladder, and prostate cancer identified by exon array analysis. Molecular & Cellular Proteomics, 7, 1214-1224.

TIMMONS, L., COURT, D. L. & FIRE, A. 2001. Ingestion of bacterially expressed dsRNAs can produce specific and potent genetic interference in Caenorhabditis elegans. Gene, 263, 103-112.

TIMMONS, L. & FIRE, A. 1998. Specific interference by ingested dsRNA. Nature, 395, 854-854. TOLLERVEY, J. R., CURK, T., ROGELJ, B., BRIESE, M., CEREDA, M., KAYIKCI, M., KONIG, J.,

HORTOBAGYI, T., NISHIMURA, A. L., ZUPUNSKI, V., PATANI, R., CHANDRAN, S., ROT, G., ZUPAN, B., SHAW, C. E. & ULE, J. 2011. Characterizing the RNA targets and position-dependent splicing regulation by TDP-43. Nature Neuroscience, 14, 452-U180.

TOMÉ, S., MANLEY, K., SIMARD, J. P., CLARK, G. W., SLEAN, M. M., SWAMI, M., SHELBOURNE, P. F., TILLIER, E. R., MONCKTON, D. G. & MESSER, A. 2013. MSH3 polymorphisms and protein levels affect CAG repeat instability in Huntington's disease mice. Plos Genetics, 9, e1003280.

208

TOYAMA, H., ICHISE, M., LIOW, J.-S., VINES, D. C., SENECA, N. M., MODELL, K. J., SEIDEL, J., GREEN, M. V. & INNIS, R. B. 2004. Evaluation of anesthesia effects on [< sup> 18</sup> F] FDG uptake in mouse brain and heart using small animal PET. Nuclear medicine and biology, 31, 251-256.

TSAI, K.-J., YANG, C.-H., FANG, Y.-H., CHO, K.-H., CHIEN, W.-L., WANG, W.-T., WU, T.-W., LIN, C.-P., FU, W.-M. & SHEN, C.-K. J. 2010. Elevated expression of TDP-43 in the forebrain of mice is sufficient to cause neurological and pathological phenotypes mimicking FTLD-U. The Journal of experimental medicine, 207, 1661-1673.

ULE, J., STEFANI, G., MELE, A., RUGGIU, M., WANG, X. N., TANERI, B., GAASTERLAND, T., BLENCOWE, B. J. & DARNELL, R. B. 2006. An RNA map predicting Nova-dependent splicing regulation. Nature, 444, 580-586.

URLAUB, H., RAKER, V. A., KOSTKA, S. & LUHRMANN, R. 2001. Sm protein-Sm site RNA interactions within the inner ring of the spliceosomal snRNP core structure. Embo Journal, 20, 187-196.

VAKA, S. R. K., SAMMETA, S. M., DAY, L. B. & MURTHY, S. N. 2009. Delivery of Nerve Growth Factor to Brain Via Intranasal Administration and Enhancement of Brain Uptake. Journal of Pharmaceutical Sciences, 98, 3640-3646.

VAN DEN BERG, M. P., ROMEIJN, S. G., VERHOEF, J. C. & MERKUS, F. 2002. Serial cerebrospinal fluid sampling in a rat model to study drug uptake from the nasal cavity. Journal of Neuroscience Methods, 116, 99-107.

VAN DEN POL, A. N., DALTON, K. P. & ROSE, J. K. 2002. Relative neurotropism of a recombinant rhabdovirus expressing a green fluorescent envelope glycoprotein. Journal of virology, 76, 1309-1327.

VAN DUSEN, C. M., YEE, L., MCNALLY, L. M. & MCNALLY, M. T. 2010. A Glycine-Rich Domain of hnRNP H/F Promotes Nucleocytoplasmic Shuttling and Nuclear Import through an Interaction with Transportin 1. Molecular and Cellular Biology, 30, 2552-2562.

VAN RAAMSDONK, J. M., WARBY, S. C. & HAYDEN, M. R. 2007. Selective degeneration in YAC mouse models of Huntington disease. Brain Research Bulletin, 72, 124-131.

VANDAM, A., WINKEL, I., ZIJLSTRABAALBERGEN, J., SMEENK, R. & CUYPERS, H. T. 1989. CLONED HUMAN SNRNP PROTEIN-B AND PROTEIN-B' DIFFER ONLY IN THEIR CARBOXY-TERMINAL PART. Embo Journal, 8, 3853-3860.

VANDERWEYDE, T., YOUMANS, K., LIU-YESUCEVITZ, L. & WOLOZIN, B. 2013. Role of stress granules and RNA-binding proteins in neurodegeneration: a mini-review. Gerontology, 59, 524-533.

VENABLES, J. P., KOH, C. S., FROEHLICH, U., LAPOINTE, E., COUTURE, S., INKEL, L., BRAMARD, A., PAQUET, E. R., WATIER, V., DURAND, M., LUCIER, J. F., GERVAIS-BIRD, J., TREMBLAY, K., PRINOS, P., KLINCK, R., ABOU ELELA, S. & CHABOT, B. 2008. Multiple and specific mRNA processing targets for the major human hnRNP proteins. Molecular and Cellular Biology, 28, 6033-6043.

VENKATRAMAN, P., WETZEL, R., TANAKA, M., NUKINA, N. & GOLDBERG, A. L. 2004. Eukaryotic proteasomes cannot digest polyglutamine sequences and release them during degradation of polyglutamine-containing proteins. Molecular Cell, 14, 95-104.

VERALDI, K. L., ARHIN, G. K., MARTINCIC, K., CHUNG-GANSTER, L.-H., WILUSZ, J. & MILCAREK, C. 2001. hnRNP F influences binding of a 64-kilodalton subunit of cleavage stimulation factor to mRNA precursors in mouse B cells. Molecular and Cellular Biology, 21, 1228-1238.

VOISINE, C., VARMA, H., WALKER, N., BATES, E. A., STOCKWELL, B. R. & HART, A. C. 2007. Identification of Potential Therapeutic Drugs for Huntington's Disease using Caenorhabditis elegans. Plos One, 2.

VU, N. T., PARK, M. A., SHULTZ, J. C., GOEHE, R. W., HOEFERLIN, L. A., SHULTZ, M. D., SMITH, S. A., LYNCH, K. W. & CHALFANT, C. E. 2013. hnRNP U Enhances Caspase-9 Splicing and Is Modulated by AKT-dependent Phosphorylation of hnRNP L. Journal of Biological Chemistry, 288, 8575-8584.

209

VYAS, T., TIWARI, S. B. & AMIJI, M. M. 2006. Formulation and physiological factors influencing CNS delivery upon intranasal administration. Critical Reviews™ in Therapeutic Drug Carrier Systems, 23.

WADSWORTH, J. D. & COLLINGE, J. 2011. Molecular pathology of human prion disease. Acta Neuropathologica, 121, 69-77.

WALKER, F. O. 2007. Huntington's disease. Lancet, 369, 218-228. WANG, H., LIM, P. J., VOGEL, B. E. & MONTEIRO, M. J. 2006. Suppression of polyglutamine-induced

toxicity in cell and animal models of Huntington's disease by ubiquilin. Human Molecular Genetics, 15, 1025-1041.

WANG, Z., GERSTEIN, M. & SNYDER, M. 2009. RNA-Seq: a revolutionary tool for transcriptomics. Nature Reviews Genetics, 10, 57-63.

WANKER, E. E., SCHERZINGER, E., HEISER, V., SITTLER, A., EICKHOFF, H. & LEHRACH, H. 1998. Membrane filter assay for detection of amyloid-like polyglutamine-containing protein aggregates. Methods in Enzymology, 309, 375-386.

WATERS, C. W., VARUZHANYAN, G., TALMADGE, R. J. & VOSS, A. A. 2013. Huntington disease skeletal muscle is hyperexcitable owing to chloride and potassium channel dysfunction. Proceedings of the National Academy of Sciences, 110, 9160-9165.

WELLINGTON, C. L., ELLERBY, L. M., GUTEKUNST, C. A., ROGERS, D., WARBY, S., GRAHAM, R. K., LOUBSER, O., VAN RAAMSDONK, J., SINGARAJA, R., YANG, Y. Z., GAFNI, J., BREDESEN, D., HERSCH, S. M., LEAVITT, B. R., ROY, S., NICHOLSON, D. W. & HAYDEN, M. R. 2002. Caspase cleavage of mutant huntingtin precedes neurodegeneration in Huntington's disease. Journal of Neuroscience, 22, 7862-7872.

WHEELER, V. C., AUERBACH, W., WHITE, J. K., SRINIDHI, J., AUERBACH, A., RYAN, A., DUYAO, M. P., VRBANAC, V., WEAVER, M. & GUSELLA, J. F. 1999. Length-dependent gametic CAG repeat instability in the Huntington's disease knock-in mouse. Human Molecular Genetics, 8, 115-122.

WHEELER, V. C., GUTEKUNST, C.-A., VRBANAC, V., LEBEL, L.-A., SCHILLING, G., HERSCH, S., FRIEDLANDER, R. M., GUSELLA, J. F., VONSATTEL, J.-P. & BORCHELT, D. R. 2002. Early phenotypes that presage late-onset neurodegenerative disease allow testing of modifiers in Hdh CAG knock-in mice. Human Molecular Genetics, 11, 633-640.

WHEELER, V. C., WHITE, J. K., GUTEKUNST, C.-A., VRBANAC, V., WEAVER, M., LI, X.-J., LI, S.-H., YI, H., VONSATTEL, J.-P. & GUSELLA, J. F. 2000. Long glutamine tracts cause nuclear localization of a novel form of huntingtin in medium spiny striatal neurons in HdhQ92 and HdhQ111 knock-in mice. Human Molecular Genetics, 9, 503-513.

WHITE, J. G., SOUTHGATE, E., THOMSON, J. N. & BRENNER, S. 1986. THE STRUCTURE OF THE NERVOUS-SYSTEM OF THE NEMATODE CAENORHABDITIS-ELEGANS. Philosophical Transactions of the Royal Society of London Series B-Biological Sciences, 314, 1-340.

WHITE, J. K., AUERBACH, W., DUYAO, M. P., VONSATTEL, J.-P., GUSELLA, J. F., JOYNER, A. L. & MACDONALD, M. E. 1997. Huntingtin is required for neurogenesis and is not impaired by the Huntington's disease CAG expansion. Nature Genetics, 17, 404-410.

WHITE, M. D., FARMER, M., MIRABILE, I., BRANDNER, S., COLLINGE, J. & MALLUCCI, G. R. 2008a. Single treatment with RNAi against prion protein rescues early neuronal dysfunction and prolongs survival in mice with prion disease. Proceedings of the National Academy of Sciences, 105, 10238-10243.

WHITE, M. D., FARMER, M., MIRABILE, I., BRANDNER, S., COLLINGE, J. & MALLUCCI, G. R. 2008b. Single treatment with RNAi against prion protein rescues early neuronal dysfunction and prolongs survival in mice with prion disease. Proceedings of the National Academy of Sciences of the United States of America, 105, 10238-10243.

WILL, C. L. & LUHRMANN, R. 2005. Splicing of a rare class of introns by the U12-dependent spliceosome. Biological Chemistry, 386, 713-724.

210

WILLINGHAM, S., OUTEIRO, T. F., DEVIT, M. J., LINDQUIST, S. L. & MUCHOWSKI, P. J. 2003. Yeast genes that enhance the toxicity of a mutant huntingtin fragment or α-synuclein. Science, 302, 1769-1772.

WILS, H., KLEINBERGER, G., JANSSENS, J., PERESON, S., JORIS, G., CUIJT, I., SMITS, V., CEUTERICK-DE GROOTE, C., VAN BROECKHOVEN, C. & KUMAR-SINGH, S. 2010. TDP-43 transgenic mice develop spastic paralysis and neuronal inclusions characteristic of ALS and frontotemporal lobar degeneration. Proceedings of the National Academy of Sciences, 107, 3858-3863.

WONG, J., GARNER, B., HALLIDAY, G. M. & KWOK, J. B. J. 2012. Srp20 regulates TrkB pre-mRNA splicing to generate TrkB-Shc transcripts with implications for Alzheimer's disease. Journal of Neurochemistry, 123, 159-171.

WOOD, W. B. 1988. 1 Introduction to C. elegans Biology. Cold Spring Harbor Monograph Archive, 17, 1-16.

WYTTENBACH, A., SAUVAGEOT, O., CARMICHAEL, J., DIAZ-LATOUD, C., ARRIGO, A. P. & RUBINSZTEIN, D. C. 2002. Heat shock protein 27 prevents cellular polyglutamine toxicity and suppresses the increase of reactive oxygen species caused by huntingtin. Human Molecular Genetics, 11, 1137-1151.

XI, L., FEBER, A., GUPTA, V., WU, M., BERGEMANN, A. D., LANDRENEAU, R. J., LITLE, V. R., PENNATHUR, A., LUKETICH, J. D. & GODFREY, T. E. 2008. Whole genome exon arrays identify differential expression of alternatively spliced, cancer-related genes in lung cancer. Nucleic Acids Research, 36, 6535-6547.

XIAO, R., TANG, P., YANG, B., HUANG, J., ZHOU, Y., SHAO, C. W., LI, H. R., SUN, H., ZHANG, Y. & FU, X. D. 2012. Nuclear Matrix Factor hnRNP U/SAF-A Exerts a Global Control of Alternative Splicing by Regulating U2 snRNP Maturation. Molecular Cell, 45, 656-668.

YAMAMOTO, A., LUCAS, J. J. & HEN, R. 2000. Reversal of neuropathology and motor dysfunction in a conditional model of Huntington's disease. Cell, 101, 57-66.

YOSHIDA, T., KOKURA, K., MAKINO, Y., OSSIPOW, V. & TAMURA, T. A. 1999. Heterogeneous nuclear RNA‐ribonucleoprotein F binds to DNA via an oligo (dG)‐motif and is associated with RNA polymerase II. Genes to Cells, 4, 707-719.

YU, S. Y., ZHAO, Y., WU, F. L., ZHANG, X., LU, W. L., ZHANG, H. & ZHANG, Q. 2004. Nasal insulin delivery in the chitosan solution: in vitro and in vivo studies. International Journal of Pharmaceutics, 281, 11-23.

YU, Z.-X., LI, S.-H., EVANS, J., PILLARISETTI, A., LI, H. & LI, X.-J. 2003. Mutant huntingtin causes context-dependent neurodegeneration in mice with Huntington's disease. The Journal of neuroscience, 23, 2193-2202.

YUGAMI, M., KABE, Y., YAMAGUCHI, Y., WADA, T. & HANDA, H. 2007. hnRNP-U enhances the expression of specific genes by stabilizing mRNA. Febs Letters, 581, 1-7.

YUO, C. Y., LIN, H. H., CHANG, Y. S., YANG, W. K. & CHANG, J. G. 2008. 5-(N-ethyl-N-isopropyl)-amiloride enhances SMN2 Exon 7 inclusion and protein expression in spinal muscular atrophy cells. Annals of Neurology, 63, 26-34.

YURYEV, A., PATTURAJAN, M., LITINGTUNG, Y., JOSHI, R. V., GENTILE, C., GEBARA, M. & CORDEN, J. L. 1996. The C-terminal domain of the largest subunit of RNA polymerase II interacts with a novel set of serine/arginine-rich proteins. Proceedings of the National Academy of Sciences of the United States of America, 93, 6975-6980.

ZÁDORI, D., NYIRI, G., SZŐNYI, A., SZATMÁRI, I., FÜLÖP, F., TOLDI, J., FREUND, T. F., VÉCSEI, L. & KLIVÉNYI, P. 2011. Neuroprotective effects of a novel kynurenic acid analogue in a transgenic mouse model of Huntington’s disease. Journal of neural transmission, 118, 865-875.

ZALA, D., BENSADOUN, J.-C., PEREIRA DE ALMEIDA, L., LEAVITT, B. R., GUTEKUNST, C.-A., AEBISCHER, P., HAYDEN, M. R. & DÉGLON, N. 2004. Long-term lentiviral-mediated expression of ciliary neurotrophic factor in the striatum of Huntington's disease transgenic mice. Experimental neurology, 185, 26-35.

211

ZARRINGHALAM, K., KA, M., KOOK, Y.-H., TERRANOVA, J. I., SUH, Y., KING, O. D. & UM, M. 2012. An open system for automatic home-cage behavioral analysis and its application to male and female mouse models of Huntington's disease. Behavioural brain research, 229, 216-225.

ZHANG, K., SMOUSE, D. & PERRIMON, N. 1991. THE CROOKED NECK GENE OF DROSOPHILA CONTAINS A MOTIF FOUND IN A FAMILY OF YEAST-CELL CYCLE GENES. Genes & Development, 5, 1080-1091.

ZHANG, S., BINARI, R., ZHOU, R. & PERRIMON, N. 2010. A genomewide RNA interference screen for modifiers of aggregates formation by mutant Huntingtin in Drosophila. Genetics, 184, 1165-1179.

ZHANG, Y. U., LI, M. W., DROZDA, M., CHEN, M. H., REN, S. J., SANCHEZ, R. O. M., LEAVITT, B. R., CATTANEO, E., FERRANTE, R. J., HAYDEN, M. R. & FRIEDLANDER, R. M. 2003. Depletion of wild-type huntingtin in mouse models of neurologic diseases. Journal of Neurochemistry, 87, 101-106.

ZIELSKE, S. P., LINGAS, K. T., LI, Y. & GERSON, S. L. 2004. Limited lentiviral transgene expression with increasing copy number in an MGMT selection model: lack of copy number selection by drug treatment. Molecular Therapy, 9, 923-931.

ZUCCATO, C., VALENZA, M. & CATTANEO, E. 2010. Molecular mechanisms and potential therapeutical targets in Huntington's disease. Physiological Reviews, 90, 905-981.

ZWILLING, D., HUANG, S. Y., SATHYASAIKUMAR, K. V., NOTARANGELO, F. M., GUIDETTI, P., WU, H. Q., LEE, J., TRUONG, J., ANDREWS-ZWILLING, Y., HSIEH, E. W., LOUIE, J. Y., WU, T., SCEARCE-LEVIE, K., PATRICK, C., ADAME, A., GIORGINI, F., MOUSSAOUI, S., LAUE, G., RASSOULPOUR, A., FLIK, G., HUANG, Y. D., MUCHOWSKI, J. M., MASLIAH, E., SCHWARCZ, R. & MUCHOWSKI, P. J. 2011. Kynurenine 3-Monooxygenase Inhibition in Blood Ameliorates Neurodegeneration. Cell, 145, 863-874.