26
Huntington Disease and the Huntingtin Protein Zhiqiang Zheng and Marc I. Diamond Hope Center for Neurological Diseases, Knight-Alzheimer Disease Research Center, Department of Neurology, Washington University in St. Louis, St. Louis, Missouri, USA I. Introduction ............................................................................... 190 II. Genetic and Clinical Features......................................................... 190 III. Structural Features of the Huntingtin Protein .................................... 192 IV. Htt Function(s) ........................................................................... 193 V. Animal Models of Huntington Disease.............................................. 195 A. R6/1 and R6/2 ........................................................................ 195 B. N171-82Q ............................................................................. 196 C. YAC Transgenic HD Models ...................................................... 196 D. BACHD ............................................................................... 196 E. HD Knock-in ......................................................................... 196 F. Other Mammalian HD Models................................................... 197 VI. Pathogenesis of HD ..................................................................... 197 A. Proteolytic Cleavage ................................................................ 197 B. Misfolding and Aggregation of Htt .............................................. 197 C. Transcriptional Dysregulation .................................................... 199 D. Mitochondrial Dysfunction........................................................ 200 E. Defects in Htt Clearance: UPS and Autophagy (See Chapter ‘‘Protein Quality Control in Neurodegenerative Disease’’) ............................ 201 VII. Therapeutic Targets for HD ........................................................... 201 A. Transcription ......................................................................... 202 B. Mitochondrial Dysfunction........................................................ 202 C. Targeting Mutant Htt Protein .................................................... 203 VIII. Conclusion ................................................................................. 204 References ................................................................................. 205 Huntington disease (HD) is a devastating neurodegenerative disease that derives from CAG repeat expansion in the huntingtin gene. The clinical syndrome consists of progressive personality changes, movement disorder, and dementia and can develop in children and adults. The huntingtin protein is required for human development and normal brain function. It is subject to posttranslational modification, and some events, such as phosphorylation, can play an enormous role in regulating toxicity of the huntingtin protein. The function of huntingtin in the cell is unknown, and it may play a role as a scaffold. Multiple mouse models of HD have now been created with fragments and full-length protein. The models show variable fidelity to the disease in Progress in Molecular Biology Copyright 2012, Elsevier Inc. and Translational Science, Vol. 107 189 All rights reserved. DOI: 10.1016/B978-0-12-385883-2.00010-2 1877-1173/12 $35.00

Huntington Diseases and the Huntingtin Protein

Embed Size (px)

DESCRIPTION

aaa

Citation preview

Page 1: Huntington Diseases and the Huntingtin Protein

Huntington Disease and theHuntingtin Protein

Progress in Molecular Biologyand Translational Science, Vol. 107 189DOI: 10.1016/B978-0-12-385883-2.00010-2

Zhiqiang Zheng andMarc I. Diamond

Hope Center for Neurological Diseases,Knight-Alzheimer Disease Research Center,Department of Neurology, WashingtonUniversity in St. Louis, St. Louis, Missouri,USA

I.

I ntroduction ...... .. ... .. ... .. .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... ..

Copyright 2A

187

1

012, Ell righ7-117

90

lsevts re3/12

II.

G enetic and Clinical Features...... .. ... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 1 90 III. S tructural Features of the Huntingtin Protein .... ... .. ... .. .. ... .. ... .. ... .. ... .. 1 92 IV. H tt Function(s) .... ... .. ... .. .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 1 93 V. A nimal Models of Huntington Disease.... ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 1 95

A. R

6/1 and R6/2 .... .. ... .. .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 1 95 B. N 171-82Q ...... ... .. ... .. .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 1 96 C. Y AC Transgenic HD Models..... ... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 1 96 D. B ACHD ...... .. ... .. ... .. .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 1 96 E. H D Knock-in ..... .. ... .. .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 1 96 F. O ther Mammalian HD Models..... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 1 97

VI.

P athogenesis of HD ...... .. .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 1 97 A. P roteolytic Cleavage ..... ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 1 97 B. M isfolding and Aggregation of Htt .... ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 1 97 C. T ranscriptional Dysregulation ...... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 1 99 D. M itochondrial Dysfunction..... .. ... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 2 00 E. D efects in Htt Clearance: UPS and Autophagy (See Chapter ‘‘Protein

Quality Control in Neurodegenerative Disease’’) .... .. .. ... .. ... .. ... .. ... ..

201 VII. T herapeutic Targets for HD ........ .. ... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 2 01

A. T

ranscription ..... .. ... .. .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 2 02 B. M itochondrial Dysfunction..... .. ... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 2 02 C. T argeting Mutant Htt Protein ...... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 2 03

V

III. C onclusion..... ... .. ... .. ... .. .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 2 04 R eferences..... ... .. ... .. ... .. .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. .. ... .. ... .. ... .. ... .. 2 05

Huntington disease (HD) is a devastating neurodegenerative disease thatderives from CAG repeat expansion in the huntingtin gene. The clinicalsyndrome consists of progressive personality changes, movement disorder,and dementia and can develop in children and adults. The huntingtin proteinis required for human development and normal brain function. It is subject toposttranslational modification, and some events, such as phosphorylation, canplay an enormous role in regulating toxicity of the huntingtin protein. Thefunction of huntingtin in the cell is unknown, and it may play a role as ascaffold. Multiple mouse models of HD have now been created with fragmentsand full-length protein. The models show variable fidelity to the disease in

ier Inc.served.$35.00

Page 2: Huntington Diseases and the Huntingtin Protein

190 ZHENG AND DIAMOND

terms of genetics, pathology, and rates of progression. Pathogenesis of HDinvolves cleavage of the protein and is associated with neuronal accumulationof aggregated forms. The potential mechanisms of neurodegeneration aremyriad, including primary effects of protein homeostasis, gene expression,and mitochondrial dysfunction. Specific therapeutic approaches are similarlyvaried and include efforts to reduce huntingtin gene expression, protein accu-mulation, and protein aggregation.

I. Introduction

Huntington disease (HD) is an autosomal dominant neurodegenerativedisorder caused by the expansion of the CAG codon repeat within the hun-tingtin gene. It affects approximately 7 per 100,000.1 CAG encodes the aminoacid glutamine, so the expansion results in an abnormally long glutamine tractwithin the N-terminus of the huntingtin protein (Htt). In addition to HD, eightother dominantly inherited neurodegenerative diseases have been described,each caused by CAG expansion. These disorders are collectively termed ‘‘poly-glutamine’’ diseases. They are prototypical protein misfolding disorders—in allcases studied, the expanded polyglutamine tracts destabilize the native proteinconformation. Penetrance is virtually 100%, with the exception of spinobulbarmuscular atrophy, a motor neuron disease and sensory neuronopathy caused bypolyglutamine expansion in the androgen receptor.2–4 This disease afflicts onlymales because dihydrotestosterone initiates pathogenesis.5

II. Genetic and Clinical Features

Normal individuals have fewer than 36 CAG repeats in the huntingtin geneand most commonly about 15–25 repeats. Repeats between 27 and 35 areunstable during meiosis, which can lead to new expansion of CAG repeats intothe disease range in subsequent generations. When repeats are present in thepathogenic range, repeat expansion inherited by children leads to ‘‘geneticanticipation,’’ or earlier disease onset than was observed in the parent. This isusually seen when fathers pass the mutant gene to their children, which may bedue to the fact that repeat instability is observed during spermatogenesis.6

Virtually all individuals with 38 or more CAG repeats are predicted to developHD symptoms if they live long enough. However, individuals with fewer than40 repeats may only manifest symptoms at a late age.7 The length of the

Page 3: Huntington Diseases and the Huntingtin Protein

HUNTINGTON DISEASE AND THE HUNTINGTIN PROTEIN 191

expanded CAG repeats correlates inversely with the age at symptom onset.8,9

Extremely long CAG repeats (>60) are highly associated with early onset HD,also known as juvenile HD. It is estimated that CAG repeat length accounts forapproximately 70% of the variance in age of onset,8,10 which means that othergenetic and environmental factors also play an important role in determiningpathogenesis.

HD is particularly devastating because of its multiple cognitive and motorfeatures, which have a relentless course over 10–20 years. Juvenile HD is veryrapid, and the affected children seldom survive into adulthood. The generalfeatures of each stage of disease are detailed below, but the timing andconstellation of features are quite variable among patients. Early HD symp-toms relate to frontal/executive dysfunction and include progressive person-ality changes, for example, irritability, and difficulty with multitasking andprioritization of activities. Psychiatric disturbances, especially depression, aswell as slight involuntary movements and poor coordination may also bepresent.1 In the middle phase of the disease, in addition to progressivecognitive disturbances, a movement disorder becomes more prominent,with frequent involuntary twitching, and writhing movements of the face,trunk, and extremities termed ‘‘chorea.’’ Disturbances of gait manifest, withfrequent falls.11 A thought disorder may also appear, with bizarre ideationthat can merge into frank psychosis.12 In the later stages of the disease, choreacan become quite severe, and a hypermetabolic state produces progressiveweight loss.13 Difficulty in swallowing leads to choking and coughing epi-sodes. At this point, the cognitive disturbance is severe enough to producedisability. Patients are unable to maintain their usual employment, and oftenrequire significant caretaking, if not institutionalization.1 In the final stages,patients become bedbound and usually succumb to wasting and aspirationpneumonia.

The Htt protein is widely expressed and causes pathology throughout thebrain. This has been readily observed using serial MRI in presymptomatic andsymptomatic patients.14 Striking atrophy occurs in the striatum, especially thecaudate and putamen. Pathology here involves vulnerable GABAergicmedium-sized spiny neurons.15 Striatal atrophy develops long before symptomonset, with �50% reduction in striatal volume at the time of diagnosis.16,17 Yetit is misleading pathologically to look solely to the striatum. HD also affectsother brain regions, including cerebral cortex (layers III, V, and VI), hippocam-pus, thalamus, globus pallidus, subthalamic nucleus, substantia nigra, whitematter, and cerebellum.7,18 The involvement of so many brain structures helpsexplain the highly variable clinical presentations of the disease. The cloning ofthe Htt gene made it possible to recognize that Htt fragments accumulate asintracellular inclusions in multiple brain regions,19 highlighting the role ofprotein misfolding and aggregation in this disorder.

Page 4: Huntington Diseases and the Huntingtin Protein

192 ZHENG AND DIAMOND

III. Structural Features of the Huntingtin Protein

Htt and its homologs are evolutionarily conserved among many vertebratesand invertebrates. For example, the coding region of Htt from humans and theFugu fish is 69% identical at the nucleotide level.20 The gene-encoding humanHtt spans over 180kb, with 67 exons. Human Htt is ubiquitously expressed,with the highest levels in the central nervous system.21 The human Htt geneconsists of two alternatively polyadenylated forms, which differ in size by 3kb.The longer transcript is predominantly expressed in brain, whereas the shorterone is more widely expressed.22,23 Within the cell, Htt is associated withmultiple organelles, including the nucleus, plasma membrane, endocytic andautophagic vesicles, endoplasmic reticulum, endosome, Golgi, and mitochon-dria.24 This may reflect a role as a scaffold protein that links multiple factorsand is consistent with its large, multidomain structure. Htt is required fornormal embryonic development.25,26

Full-length Htt is composed of �3114 amino acids, producing a protein of�350kDa. Due to its size, it has not yet been possible to generate full-lengthHtt crystals. However, the structures of several domains have been described inrelative detail. To elucidate the role of the N-terminal polyglutamine tract innormal Htt function, an ablation of the seven CAG repeats in the mouse genewas performed.27 Development was normal (unlike knockout of the Htt gene,which is embryonic lethal), and adult mice exhibited only subtle defects inlearning and memory. This suggests that the polyglutamine stretch contributesto the subtle regulation of neural function by Htt but is not required for Htt’sessential functions.

Htt also contains HEAT (Htt, elongation factor 3, the PR65/A subunit ofprotein phosphatase 2A and the lipid kinase Tor) repeats. HEAT domainsconsist of �50 amino acids and comprise two antiparallel a-helices forming ahairpin that is normally involved in protein–protein interactions. A total of 16HEAT repeats have been identified in Htt, organized into four clusters.28,29

The biological role of HEAT domains in Htt function is unknown.Wild-type Htt is normally found in the cytoplasm, but mutant Htt accu-

mulates within the nucleus, especially as N-terminal truncation frag-ments.19,21,30–33 A conserved nuclear export signal (NES) in the C-terminushas been identified.34 Another potential NES exists in the N-terminal 17 aminoacid region that has been demonstrated to interact with the nuclear poreprotein (TPR), which is involved in the export of proteins from nucleus intocytoplasm.35

Multiple proteolytic cleavage sites have been identified in Htt, which likelygenerate a wide range of N-terminal fragments. Putative Htt proteases includecaspases, calpain, aspartic endopeptidases, and matrix metalloproteinases. Httis cleaved by caspase 3 in vitro at amino acids 513 and 552,36,37 caspase 2 at

Page 5: Huntington Diseases and the Huntingtin Protein

HUNTINGTON DISEASE AND THE HUNTINGTIN PROTEIN 193

amino acids 552,38 and caspase 6 at amino acids 586.37 Two putative calpainsites are located at amino acids 469 and 536.39,40 The matrix metalloproteinase,MMP-10, which is active in HD mice, cleaves Htt at amino acid 402.41 Httfragments with mutated cleavage sites, which are resistant to the cleavage bysome of these enzymes, exhibit less toxicity in vitro and in vivo, suggesting thatproteolytic processing of full-length Htt is probably an important event in theinitiation of HD pathogenesis.36,42,43

Htt also has many posttranslational modifications, including ubiquitination,phosphorylation, acetylation, sumoylation, and palmitoylation. These could reg-ulate its stability, localization, and function. Ubiquitination targets Htt for degra-dation.30,44 Three lysines in its N-terminus, K6, K9 and K15, have beenidentified as sumoylation sites.45 While ubiquitination of Htt may serve a protec-tive role by increasing its clearance, Htt sumoylation at these lysines exacerbatespolyglutamine-mediated neurodegeneration in a Drosophila model of HD.45

Many phosphorylation sites in the Htt N-terminus also have been identified.Phosphorylation of threonine 3 may affect Htt aggregation and toxicity.46 It hasalso been observed that the IkappaB kinase (IKK) complex phosphorylates Htt atserine 13 and 16, which can promote the modification of the adjacent lysines andenhance Htt clearance. Expansion of the polyglutamine tract may reduce Httphosphorylation, helping promote accumulation of mutant Htt.47 Additionally,the importance of these residues has been clearly shown in a full-length Httmouse model based on bacterial artificial chromosome (BAC) expression. In thiswork, phosphomimic aspartate substitutions at serines 13 and 16 completelyprevented polyglutamine-mediated toxicity, indicating that these amino acidsmay be very important determinants of pathogenesis.48 Several other phosphor-ylation sites in Htt have been identified, including S421 by Akt; S434, S1181, andS1201 by Cdk5t.49–52 Further studies, especially using phosphomimetic muta-tions in full-length mouse models, will be needed to clarify the role of phosphor-ylation at these sites in controlling Htt function and toxicity. Additionally, it hasbeen shown that Htt acetylation at lysine 444 is required for degradation by themacroautophagy pathway.53 Finally, palmitoylation at C214 by HIP14 may alterHtt aggregation. Polyglutamine tract expansion decreases the interaction be-tween HIP14 and Htt, resulting in decreased palmitoylation, increased inclu-sions, and enhanced neurotoxicity (Fig. 1).54

IV. Htt Function(s)

Htt probably plays multiple roles in development and physiology. It isclearly essential for embryonic development, as Htt gene knockout is lethalin mice by embryonic day 8.5.25,26 Increased cellular death in Htt knockoutanimals may indicate an anti-apoptotic role for Htt, which is supported by the

Page 6: Huntington Diseases and the Huntingtin Protein

N17 polyQ polyP NESHEATs

Proteolyic cleavage

HEATs HEATs

114–413 672–969 2667–2938400–600 2397

Phosphorylation:T3, S13, S16Ubiquitination/sumoyalation: K6, K9, K15NES: L4, L7, F11, L14

Phosphorylation421/434

Palmitoylation214

Phosphorylation1181/1201

Acetylation444

FIG. 1. Diagram of regions of the Htt protein. Full-length htt is composed of �3144 aminoacids, with a molecular weight of 350kDa. The N-terminal 17 amino acid (N17) region is highlyconserved and contains multiple posttranslational modifications, such as phosphorylation andubiquitination, which could strongly affect the pathogenesis. Following the N17 region are thepolyglutamine (polyQ) and polyproline (polyP) regions. The function of polyproline is still largelyunclear. Studies have suggested a hot spot of proteolyic cleavage exists between amino acids 400and 600. Putative proteases include caspases 2, 3, 6, calpain, and matrix metalloproteinase 10. Up to16 HEAT repeats with unknown functions have been identified in Htt. Multiple regions ofphosphorylation have been identified. The C-terminus contains a leucine-rich nuclear export signal(NES), which may target full-length Htt to cytoplasm. Another leucine-rich NES could exist in theN17 region.

194 ZHENG AND DIAMOND

findings that overexpression of full-length wild-type Htt in cultured striatalneurons protects them from apoptosis, whereas depletion of wild-type Httmakes neurons more sensitive to apoptotic insults.55–57 The molecular mecha-nism of the anti-apoptotic function of Htt is unclear. One possible mechanismis that Htt can bind and prevent the formation of the pro-apoptotic Hip1-HIPPI (Hip1 protein interactor) complex.58 Another possible mechanism isthat Htt can directly block the activation caspase 3 and 9.55,59

Many studies have implicated Htt in transcription. Htt interacts withvarious transcription factors and transcriptional regulatory proteins. For exam-ple, Htt has been proposed to regulate the production of the brain-derivedneurotrophic factor (BDNF).60,61 BDNF transcription is suppressed by thebinding of the repressor element-1 silencing transcription factor (REST)/neuro-restrictive silencer factor (NRSF) to its upstream DNA elementneuron-restrictive silencer element.62,63 Htt acts as positive regulator ofBDNF transcription by preventing the recruitment of REST/NRSF to itsDNA response element. Evidence suggests Htt may not interact directly withREST/NRSF, but rather forms a complex with HAP1 and RILP, which candirectly bind REST/NRSF.63 It is unclear which, if any, of the HD symptomsmight result from the effect on BDNF.

By interacting with a variety of endocytic/trafficking proteins, such as a-adaptin, HIP1, HIP14, HAP1, HAP40, PACSIN1 and SH3GL3, clathrin, anddynamin, Htt also may participate in long- and short-range axonal transport and

Page 7: Huntington Diseases and the Huntingtin Protein

HUNTINGTON DISEASE AND THE HUNTINGTIN PROTEIN 195

vesicle trafficking. For example, in addition to the role of transcriptionalregulation of BDNF, full-length wild-type Htt also stimulates BDNF vesiculartrafficking in neurons,64 while knockdown of Htt reduces such transport.65 Httis enriched at synaptic terminals and interacts with cytoskeletal and synapticvesicle proteins to regulate synaptic activity in neurons. By interacting with theSH3 domain of PSD95, a key protein in synaptic transmission, Htt mightregulate synaptic plasticity.66

V. Animal Models of Huntington Disease

With the cloning of the HD gene, multiple model organisms, such asCaenorhabditis elegans (worm), Drosophila (fruit fly), and Danio rerio (zebra-fish), have now been used to model polyglutamine toxicity.67–71 These areadvantageous in their rapid development of pathology and low cost. Theyalso exhibit many features of the human disease, such as progressive neuronalcell death, reduced lifespan, aggregation of Htt, and formation of inclusions.These models are especially useful in the identification of pathways involved inHD and the screening of potential suppressors of HD pathology. However, it isstill uncertain to what extent these will predict effects of chemical and geneticmodifiers of pathology in mammals. Thus they are probably best used toidentify candidates for further study in mammalian systems, rather than tovet candidates identified in mammalian systems.

Many mouse HD models now exist, both transgenic and knock-in. A listfollows of the commonly used models.

A. R6/1 and R6/2

The transgenic R6/1 and R6/2 HD mouse lines contain �1kb of human

HD promoter region followed by exon 1 of the Htt gene, carrying 115–156CAG repeats and 262bp of intron 1.19,72 R6/2 mice develop progressive neuro-logical phenotypes, such as reduced brain and striatal volume, extensive intra-neuronal inclusion formation, and relatively little cell death. Mice exhibit aresting tremor and shuddering movements as early as 5weeks of age. They haveprogressive motor deficits and seizures. The phenotype is rapidly progressive,and these animals die between 12 and 18 weeks, depending on the laboratorysetting and strain background. Their predictable onset and rapid course hasmade them a mainstay of preclinical drug trials. However, their distribution ofpathology is not anatomically faithful to human disease, and thus they mayserve best as a model of ‘‘polyglutaminopathy’’ (which is still very useful), ratherthan true HD. The R6/1 line has a more slowly progressive phenotype than theR6/2 line.

Page 8: Huntington Diseases and the Huntingtin Protein

196 ZHENG AND DIAMOND

B. N171-82Q

The N171-82Q transgenic HD mouse line was generated by expressing a

cDNA encoding an N-terminal fragment (171 amino acids) of huntingtin with82 glutamines under the control of the mouse prion promoter.73 Starting at theage of 10 weeks, these mice develop behavioral abnormalities including coor-dination loss, tremors, hypokinesis, and abnormal gait. Intranuclear inclusionbodies and neuritic aggregates are also observed in multiple types of neurons.

C. YAC Transgenic HD Models

Yeast artificial chromosome (YAC) transgenic HD mice were generated by

expressing the full-length human Htt gene with varying numbers of CAGrepeats.74 YAC46 and YAC72 were first created. YAC72 mice show earlyelectrophysiological abnormalities and selective degeneration of mediumspiny neurons by 12 months. Interestingly, neurodegeneration in these miceoccurs in the absence of obvious Htt inclusion formation. To create more rapiddisease onset and reduce inter-animal variability, YAC128 was generated (128CAG repeats).33,75 This line exhibits uniform phenotypes as early as 3 monthsof age. A progressive motor deficit on the rotarod occurs at the age of 6 months.Severe neurodegeneration occurs, with striatal atrophy at 9 months, and corti-cal atrophy at 12 months of age.

D. BACHD

Recently, new full-length transgenic HD mice have been generated via the

BAC technique. Like the YACmodels, the BAC transgenic allows expression ofthe human Htt gene within the context of the surrounding genomic regulatoryelements. The BAC readily enables modification of the Htt gene within thiscontext, which facilitates the study of various mutants. In BACDHDmice, full-length human Htt with 97 CAG repeats is expressed under the control ofendogenous Htt regulatory elements. These mice exhibit progressive motordeficits in rotarod performance starting as early as 2 months of age, neuronalsynaptic dysfunction, and late-onset selective neurodegeneration in both stria-tum and cortex.76 This model has also been used with great success to definethe effect of specific N-terminal mutations in the Htt gene.48

E. HD Knock-in

In transgenic mice, the gene of interest is randomly inserted into the

genome. This may interfere with the activity of other genes. Also, overexpres-sion of the transgene may produce artifactual results. To overcome thesedrawbacks, ‘‘knock-in’’ HDmice have been generated in which the endogenousmouse Htt gene is replaced by a mutated Htt that carries expanded CAGrepeats. The knock-in models are generally considered the most ‘‘faithful’’ tothe human disease in that mutated Htt is placed in the appropriate context

Page 9: Huntington Diseases and the Huntingtin Protein

HUNTINGTON DISEASE AND THE HUNTINGTIN PROTEIN 197

within the mouse genome and expression is controlled by the endogenous Httpromoter. However, knock-in HD mice generally exhibit subtle phenotypes.Neuropathology is observed at a very old age, even with the presence of a verylong polyglutamine tract.77–79 While these animals may be less useful forpreclinical studies, they may be very useful for studies of more subtle HDphenotypes, for example, network dysfunction or early metabolic changes.

F. Other Mammalian HD Models

A transgenic rat HD model has been generated by expressing a truncated

Htt fragment carrying 51 CAG repeats under the control of the endogenous ratHtt promoter.80 This model exhibits adult-onset neurological phenotypes,slowly progressive motor dysfunction, and formation of intracellular inclusionsin neurons. In vivo imaging has demonstrated significant striatal shrinkage,suggesting this model could be useful for preclinical imaging studies. A trans-genic monkey model has been developed by expressing Htt exon 1 with 84CAG repeats fused to the green fluorescent protein (GFP). This model showsextensive neuronal inclusions and mimics some important clinical features ofHD, such as dystonia and chorea (Table I).81

VI. Pathogenesis of HD

A. Proteolytic Cleavage

The proteolysis of full-length of Htt, which results in N-terminal fragments

containing the polyglutamine tract, appears to be a crucial step for progressionof HD. The truncated N-terminal fragments tend to display more aggregationand neuronal toxicity than full-length Htt. The presence of these truncatedN-terminal fragments in both HDmice and patients has been shown.31,40,82 Sofar, a variety of proteases that cleave Htt have been identified, includingcaspases 2, 3, 6, 7, calpains, and aspartic endopeptidases. Some proteaseinhibitors, or mutations that block Htt cleavage, have been demonstrated toslow HD progression in mice.37,42 Cleavage of Htt could produce a repertoireof fragments, each with a distinct propensity to cause disease. Further work willdetermine the specific role of these different truncated N-terminal fragmentsin HD progression, and to what extent a single protease can be targeted to slowpathogenesis.

B. Misfolding and Aggregation of Htt

The role of protein aggregation in the pathogenesis of HD is a confusing

and controversial area. Htt inclusions are formed within cells. They comprisemacromolecular aggregates of Htt, but the distribution of aggregate types andstructures is poorly understood. Aggregates of many types have been

Page 10: Huntington Diseases and the Huntingtin Protein

TABLE IMOUSE MODELS OF HUNTINGTON DISEASE

Model Promoter ConstructPolyglutaminelength Age of onset Neuropathology Behavior phenotype

R6/1 Human Htt �1.9kb fragment of the50-human Htt gene

115 for R6/1 5 months forR6/1

Intranuclear and neuropilaggregates throughout thebrain; global brainatrophy; minimal celldeath

Tremors and gaitabnormalities; rotaroddeficit; clasping behavior;learning deficit

R6/2 144 for R6/2 2 months forR6/2

N171-82Q Mouse prion N-terminal 171 aminoacids of human Htt

82 5 months Inclusions in striatum,cortex, hippocampus andamygdala; degeneration instriatum

Tremors and gaitabnormalities; rotaroddeficit; loss ofcoordination; hypokinesis

YAC128 Human Htt YAC expressing full-length human Htt

128 3 months Inclusions in striatum;neuron loss in striatum.

Rotarod deficit; clasping; gaitabnormalities; circlingbehavior

BACHD Human Htt BAC expressing full-length human Htt

97 3 months Synaptic dysfunction;cortical and striatalatrophy.

Rotarod deficit

HdhQ92–111 Mouse Htt Replace mouse Htt exon1 with a humanmutant exon 1

92–111 20 months No brain atrophy andneuronal loss; gliosis instriatum

No obvious behavior deficit

HdhQ140 Mouse Htt Replace the endogenouspolyglutamine withexpandedpolyglutamine

140 12 months Nuclear and neuropilinclusions in striatum,cortex, nucleusaccumbens, and olfactorytubercle

Increased locomotor activityand rearing at 1 month ofage; hypoactivity and gaitabnormalities later

HdhQ150 Mouse Httpromoter

Replace the endogenouspolyglutamine withexpandedpolyglutamine

150 4 months Nuclear inclusions instriatum; striatal gliosis

Clasping behavior; gaitabnormalities; rotaroddeficit; hypoactivity

Page 11: Huntington Diseases and the Huntingtin Protein

HUNTINGTON DISEASE AND THE HUNTINGTIN PROTEIN 199

described. Fibrils are long, ordered, para-crystalline polymers of Htt peptidesthat are readily sedimented by centrifugation. Oligomers are small, orderedstructures that may range in size from two to possibly hundreds of mono-mers. Other ordered structures also may exist (e.g., annuli). In addition,amorphous aggregates form that lack ordered structure. There is also likelyto be considerable conformational diversity even among these structures. Anemerging consensus holds that oligomers, rather than large fibrils or intracel-lular inclusions, are the toxic agents in HD. It is clear that Htt formsinclusions in patients, but their prevalence within affected regions is notcompletely certain.19,83–85

Expanded polyglutamine Htt is aggregation-prone in vitro and in vivo, anda similar polyglutamine length threshold exists for in vitro aggregation and forin vivo toxicity in patients.86 Aggregation of Htt fragments depends highly onthe length of polyglutamine, fragment size, and intracellular protein interac-tors.87–90 Polyglutamine expansion could promote aggregation by facilitating aconformational transition to a novel aggregation-prone form of Htt. Either amonomer or an oligomer then could serve to nucleate (seed) aggregate growth,producing structures rich in b-sheets.91

It is very difficult to dissociate Htt cellular toxicity from its ability toaggregate using site-directed mutagenesis—mutations in Htt that block aggre-gation also block its toxicity.92 Yet, it is clearly possible to dissociate inclusionformation from toxicity.93 Although in vivo studies have shown an association ofinclusions with HD progression,72,94 several lines of evidence also suggest littlecorrelation between inclusions and behavioral dysfunction or neuronal loss.74–76

In addition, in some HD animal models, little inclusion formation is observedduring HD progression, which has caused some to argue that aggregation maynot be required for HD pathogenesis. However, it is likely that ‘‘micro-aggregates,’’ or oligomers, are present before the appearance of large proteindeposits. Emerging in vitro and in vivo evidence has shown the existence ofsuch Htt oligomers, which appear to be more toxic than Htt fibrils, and are thuslikely to be a principal pathogenic agent.83,84,95,96

C. Transcriptional Dysregulation

A number of studies have documented alterations of gene expression that

are associated with HD pathogenesis. The challenge has been to sort outpathogenic changes, that is, those that, if corrected, would alter the course ofdisease, from epiphenomena which simply reflect metabolic changes in sickcells. Htt with expanded polyglutamine has been reported to bind and changethe activity of many transcription factors and cofactors, including CREB-binding protein (CBP), p300/CBP-associated factor (p/CAF), p53, sp1,TAFII130, and PQBO1.97,98 Htt binding may sequester these transcriptionalfactors/cofactors away from their normal DNA elements. For example, binding

Page 12: Huntington Diseases and the Huntingtin Protein

200 ZHENG AND DIAMOND

of mutated Htt to CBP leads to relocation of CBP from the nucleus into Httinclusions.99 Another well-characterized transcription factor, REST/NRSF,binds the NRSE within the BDNF promoter to suppress its transcription.62,63

Wild-type Htt binds cytosolic REST/NRSF and prevents its translocation tothe nucleus, whereas Htt with expanded polyglutamine interacts with REST/NRSF less effectively, resulting in the accumulation of REST/NRSF in thenucleus. Mutated Htt also occupies the promoter region of PGC-1a by asso-ciating with the CREB–TAF4 complex. This is reported to suppress expressionof PGC-1a. Reduced PGC-1a mRNA expression has been found in bothmouse and human HD brains and may partly account for the impaired energymetabolism observed in HD.100 So far, it has not been possible to demonstratea single gene or group of genes whose elimination or overexpression couldreplicate the HD phenotype, so that the question of whether HD truly derivesfrom a primary genetic disturbance remains uncertain.

D. Mitochondrial Dysfunction

As with transcriptional dysregulation, it is not yet clear whether the mitochon-

drial dysfunction found inHD is a cause, or simply an epiphenomenon, of neuronaltoxicity. Indeed, mitochondrial dysfunction and oxidative stress have been impli-cated in a variety of neurodegenerative diseases.101,102 Mitochondria are an impor-tant source of reactive oxygen species (ROS) and also are a key target for ROSdamage. ROS can cause peroxidation of lipids and oxidative damage of DNA byforming 8-hydroxy-20-deoxyguanosine (8-OHdG). Increased striatal lipid peroxi-dation is found to parallel the neurological phenotype in HD transgenic mice.103

Oxidative damage to DNA also is found in both HD patients and transgenicmice.104,105 Mutated Htt has been reported to damage mitochondria by multiplemechanisms, including reduction of energy production and membrane potential,alteration in calcium homeostasis, enhancement of oxidative stress, and triggeringof the caspase-mediated apoptotic pathways.106–109 These mitochondrial defectsmight be caused by an interaction betweenmutantHtt and the outermembrane ofmitochondria.109–111Htt could also indirectly impactmitochondrial biogenesis andfunction through effects on PGC-1a gene expression.

It has been demonstrated that mitochondrial energy metabolism is im-paired in both HD patients and HDmouse models.112,113 Decreased activity ofseveral important enzymes involved in energy metabolism has been reported,including respiratory chain complexes I, II, III, and IV.114–116 Reduction ATPproduction and energy failure could make striatal cells more vulnerable toother pathological insults. Mitochondria isolated from cells expressing expand-ed Htt also have decreased membrane potential. Neuronal cells derived fromknock-in HD mice HdhQ111 exhibit high sensitivity to the calcium-inducedpermeability transition pore opening. It has also been demonstrated that anexpanded N-terminal Htt fragment, but not wild-type Htt, significantly

Page 13: Huntington Diseases and the Huntingtin Protein

HUNTINGTON DISEASE AND THE HUNTINGTIN PROTEIN 201

decreases the calcium concentration threshold necessary to trigger the mito-chondrial permeability transition pore opening and increases the propensity forcalcium-induced cytochrome c release.110 Mitochondria from clonal striatalcells expressing mutated Htt also show a reduced buffering capacity for calci-um uptake.117

E. Defects in Htt Clearance: UPS and Autophagy(See Chapter ‘‘Protein Quality Control inNeurodegenerative Disease’’)

The ubiquitin–proteasome system (UPS) mediates attachment of ubiquitin

to target proteins and their subsequent degradation by the proteasome. Dys-function of the UPS in HD has been demonstrated.30,118,119 This may be linkedmore broadly to general defects in protein quality control that are induced by asingle misfolded species.120 Htt aggregates are modified with ubiquitin andcolocalize with proteasome subunits, suggesting that aggregated or misfoldedHtt may be targeted for degradation by the UPS. It has been proposed thatsequestration of UPS components into Htt inclusions alters their subcellularlocalization and thus the efficiency of the UPS.19,30 Alternatively, polyglutamine-containing proteins may not be degraded efficiently within the proteasome, andexpanded polyglutamine sequences trapped in the proteasome may block theentry of other substrates into the barrel of its catalytic core.121 Finally, it has beenproposed that inefficient degradation of expanded polyglutamine tracts may leadto the production of aggregation-prone fragments.119,122 The importance of Httdegradation in the aggregation process is also linked to its effect on protein levels.Proteasome inhibitors increase the formation of Htt inclusions, whereas en-hancement of proteasome degradation by overexpression of heat shock proteinsor chaperones decreases Htt aggregation.123–125

Multiple studies have now shown that autophagy also plays a role in theclearance of huntingtin aggregates.126,127 Activation of autophagy, either bychemical treatments (e.g., rapamycin) or by overexpression of autophagy com-ponents, increases the clearance of mutated Htt and reduces its aggrega-tion.126,128 Posttranslational modification of the mutant Htt by acetylation atlysine residue 444 (K444) also enhances its clearance by increasing Htt deliveryto the autophagosome.53

VII. Therapeutic Targets for HD

There are no effective therapies for HD, and treatments are limited tocontrolling psychiatric symptoms and chorea. Various therapeutic strategies arenow proposed but are still in their experimental stages.

Page 14: Huntington Diseases and the Huntingtin Protein

202 ZHENG AND DIAMOND

A. Transcription

Alterations in gene expression in animal models of HD have led some to

propose that correcting these deficits will rescue neuronal dysfunction anddegeneration. BDNF is perhaps the most compelling target.60,129,130 Likewise,disruption of PGC1-a transcription by Htt has also been proposed as a mecha-nism for defective mitochondrial function.100,131,132 For most transcriptionalchanges, however, it remains unclear whether they represent a cause or conse-quence of HD pathology. In certain cases, gene expression alterations might infact be adaptive. Mutant Htt has been proposed to disrupt the activity of manykey transcriptional factors or transcription-regulatory proteins, generally lead-ing to the downregulation of target gene expression. A number of compoundshave been proposed to restore normal transcription activity in HD animalmodels. For example, HDAC inhibitors such as suberoylanilide hydroxamicacid (SAHA), sodium butyrate, and phenylbutyrate have been administeredto experimental animals,133–137 but the enzymatic targets of these compoundsremain elusive, and may not be acetyl transferases that are directly related totranscription. Indeed, some studies suggest that HDAC inhibitors may beneuroprotective by increasing acetylation of other nonhistone proteins, suchas tubulin,138 or by increasing levels of heat shock proteins.139 Other effectsof HDAC inhibitors, such as prevention of Htt aggregation, anti-inflammation, and anti-apoptotic properties, may also mediate their neuro-protection in HD.140,141 Further studies are needed to clarify the contribu-tions of epigenetic modifications and nonhistone effects when applyingHDAC inhibitors as HD therapies. HDAC inhibitors are of limited potentialbecause of their high toxicity, and thus identification of more specific targetscould improve their utility.

B. Mitochondrial Dysfunction

Mutant Htt, directly or indirectly, is associated with mitochondrial dys-

function, resulting in impaired energy production, increased intracellularcalcium, oxidative stress, and apoptosis. Creatine, which stimulates mito-chondrial respiration and has antioxidant properties, has alleviated symptomsin mouse HD models.142,143 Coenzyme Q10, which promotes energy produc-tion, is also neuroprotective in R6/2 HD mice.144,145 However, clinical trialswith these compounds did not show significant benefits in HD patients.146,147

While PGC1-a, a transcriptional regulator of mitochondrial biogenesis andrespiration, has been implicated as a potential target of HD therapy,100,131,132

it is not clear yet how this pathway could be targeted specifically with smallmolecules.

Page 15: Huntington Diseases and the Huntingtin Protein

HUNTINGTON DISEASE AND THE HUNTINGTIN PROTEIN 203

C. Targeting Mutant Htt Protein1. PROTEOLYSIS

If Htt proteolysis initiates pathogenesis, then compounds that decrease itscleavage could have therapeutic value. Caspase 6 has been implicated as apotential therapeutic target, since mutation of a cleavage site for this proteasereduced toxicity of the full-length protein.42 No inhibitors of this enzyme areyet in clinical trials. Minocycline had small effects when administered to R6/2mice, improving lifespan.148,149 This compound may inhibit caspase activity,although it has been shown to produce benefits only in a mouse model thatexpresses an Htt fragment, and thus it is unlikely to function via preventing Httcleavage. Nonetheless, a clinical trial with minocycline is currently beingconducted (www.huntington-study-group.com) because this is an FDA-approved drug. If specific proteases are clearly implicated in Htt proteolysis,this may create new opportunities to design more targeted therapies.

2. PROTEIN CLEARANCE

Activation of autophagy may enhance clearance of aggregated Htt and offerprotection in HD.150 A principal challenge in developing therapies will be toactivate components of this pathway that are selective for protein aggregation,without inducing more general toxicity. The mTOR inhibitor rapamycin, whichinduces autophagy, has been shown to be beneficial in cell, Drosophila, andmouse HD models.126,151,152 However, strong side effects, especially immuno-suppression, prevent its chronic use in HD patients. Recently, new compoundswere described to selectively stimulate the clearance of mutant Htt; however,the underlying molecular mechanisms remain elusive.153–155 Given the wide-spread role of protein aggregation in neurodegenerative diseases, this mode oftherapy could be applicable more generally.

3. PROTEIN AGGREGATION

Much effort has been devoted to the discovery of compounds that inhibitHtt aggregation. Many compounds that block Htt aggregation directly orindirectly have been identified in cell-based assays and in Drosophila mod-els.156–159 However, effects of these compounds in HDmouse models are so farmodest, at best.156,157,160,161 This may be due to poor bioavailability, and in thecase of compounds that stoichiometrically inhibit Htt aggregation via directinteraction, it may be especially difficult to achieve sufficient brain concentra-tions without concurrent toxicity. Induction of molecular chaperones, such asHsp70, may promote protein refolding or degradation of mutant Htt.162–164

It remains unclear, however, whether global induction of chaperones can beaccomplished safely on a chronic basis. A better understanding of protein

Page 16: Huntington Diseases and the Huntingtin Protein

204 ZHENG AND DIAMOND

interactions and regulatory pathways that control Htt misfolding and aggrega-tion could help create more effective therapies using this strategy, as the targetsnow are somewhat limited.

4. GENE THERAPY

Gene therapy using RNA interference (RNAi) or antisense oligonucleo-tides is promising for a variety of genetic disorders. Indeed, a phase I clinicaltrial using antisense oligonucleotides to treat patients suffering from amyo-trophic lateral sclerosis due to mutation in SOD1 is now under way (http://www.alsa.org/patient/drug.cfm?id¼1592). The use of shRNA- or siRNA-basedknockdown of mutant Htt expression has been validated in HD mouse mod-els,165,166 as has the use of antisense oligonucleotides.167 Several issues must beaddressed before these approaches can be applied to patients. Off-targeteffects are probably more widespread than generally appreciated.168,169 Deliv-ery of siRNA or antisense oligonucleotides requires chronic administration intothe cerebrospinal fluid (CSF) via catheter and whole brain penetration is notfully established for large animals. Finally, the safety of virus-based geneknockdown in the brain is not yet readily controlled or proven safe. Althoughthere are still significant hurdles, the potential of genetic therapy in HD isnonetheless promising.

5. ARTIFICIAL PEPTIDES OR INTRABODIES

Intrabodies are recombinant antibody fragments that target intracellularHtt.170–173 The protein specificity created by a peptide has allowed targeting ofvarious domains of the Htt molecule, including the N-terminal region and thepolyglutamine tract, to reduce aggregation and promote protein clearance.174

This therapeutic strategy is still in a very early stage and will need more work tosolve technical issues such as stability and specificity of intrabodies and optimaldelivery methods, yet it could offer highly selective modulation of intracellularHtt clearance and aggregation.

VIII. Conclusion

Genetic and experimental evidence suggests that most polyglutamine pro-teins, including Htt, exert toxic effects as a result of an abnormal conformationinduced by the polyglutamine expansion. This is likely distinct from theirnormal function and linked to aggregation and abnormal protein interactions.This contrasts other types of disease-causing mutations that accentuate orinhibit normal enzymatic activities. It is still unknown what Htt’s crucial cellularfunctions are, and indeed, it is likely that the protein has several, given its largesize and multiple conserved regions. Despite the fact that Htt’s disease-causing

Page 17: Huntington Diseases and the Huntingtin Protein

HUNTINGTON DISEASE AND THE HUNTINGTIN PROTEIN 205

activities represent an apparent gain of toxic function, its normal interactionswithin the cell will hold important clues to how this pathogenic activity isreleased, for example, via proteolysis, posttranslational modifications, or pro-tein contacts that prevent or promote misfolding or regulate clearance. This hasobvious implications for Htt-specific therapy. Genetic therapies will becomemore feasible as delivery systems evolve, while nonspecific therapies aimed atbolstering general cellular resistance to protein misfolding and metabolicdefects may also play a role. The plethora of mouse models will facilitatepreclinical testing, especially as our understanding develops about how bestto use them. Thus, the treatment of HD is likely to depend on a detailed andevolving knowledge of Htt biology, coupled with interventions that targetcrucial events in the pathogenic cascade.

References

1. Bates G, Harper P, Jones L, editors.Huntington’s disease. New York: Oxford University Press;2002.

2. Harding AE. Molecular genetics and clinical aspects of inherited disorders of nerve andmuscle. Curr Opin Neurol Neurosurg 1992;5:600–4.

3. Kennedy WR, Alter M, Sung JH. Progressive proximal spinal and bulbar muscular atrophy oflate onset. A sex-linked recessive trait. Neurology 1968;18:671–80.

4. La Spada AR, Wilson EM, Lubahn DB, Harding AE, Fischbeck KH. Androgen receptor genemutations in X-linked spinal and bulbar muscular atrophy. Nature 1991;352:77–9.

5. Katsuno M, Adachi H, Kume A, Li M, Nakagomi Y, Niwa H, et al. Testosterone reductionprevents phenotypic expression in a transgenic mouse model of spinal and bulbar muscularatrophy. Neuron 2002;35:843–54.

6. Duyao M, Ambrose C, Myers R, Novelletto A, Persichetti F, Frontali M, et al. Trinucleotiderepeat length instability and age of onset in Huntington’s disease. Nat Genet 1993;4:387–92.

7. Walker FO. Huntington’s disease. Lancet 2007;369:218–28.8. Andrew SE, Goldberg YP, Kremer B, Telenius H, Theilmann J, Adam S, et al. The relation-

ship between trinucleotide (CAG) repeat length and clinical features of Huntington’s disease.Nat Genet 1993;4:398–403.

9. Rubinsztein DC, Barton DE, Davison BC, Ferguson-Smith MA. Analysis of the huntingtingene reveals a trinucleotide-length polymorphism in the region of the gene that contains twoCCG-rich stretches and a correlation between decreased age of onset of Huntington’s diseaseand CAG repeat number. Hum Mol Genet 1993;2:1713–5.

10. Brinkman RR, Mezei MM, Theilmann J, Almqvist E, Hayden MR. The likelihood of beingaffected with Huntington disease by a particular age, for a specific CAG size. Am J HumGenet1997;60:1202–10.

11. Grimbergen YA, Knol MJ, Bloem BR, Kremer BP, Roos RA, Munneke M. Falls and gaitdisturbances in Huntington’s disease. Mov Disord 2008;23:970–6.

12. van Duijn E, Kingma EM, van der Mast RC. Psychopathology in verified Huntington’s diseasegene carriers. J Neuropsychiatry Clin Neurosci 2007;19:441–8.

13. Politis M, Pavese N, Tai YF, Tabrizi SJ, Barker RA, Piccini P. Hypothalamic involvement inHuntington’s disease: an in vivo PET study. Brain 2008;131:2860–9.

Page 18: Huntington Diseases and the Huntingtin Protein

206 ZHENG AND DIAMOND

14. Kloppel S, Henley SM, Hobbs NZ,Wolf RC, Kassubek J, Tabrizi SJ, et al. Magnetic resonanceimaging of Huntington’s disease: preparing for clinical trials. Neuroscience 2009;164:205–19.

15. Mitchell IJ, Cooper AJ, Griffiths MR. The selective vulnerability of striatopallidal neurons.Prog Neurobiol 1999;59:691–719.

16. Aylward EH, Brandt J, Codori AM, Mangus RS, Barta PE, Harris GJ. Reduced basal gangliavolume associated with the gene for Huntington’s disease in asymptomatic at-risk persons.Neurology 1994;44:823–8.

17. Aylward EH, Sparks BF, Field KM, Yallapragada V, Shpritz BD, Rosenblatt A, et al. Onset andrate of striatal atrophy in preclinical Huntington disease. Neurology 2004;63:66–72.

18. Vonsattel JP. Huntington disease models and human neuropathology: similarities and differ-ences. Acta Neuropathol 2008;115:55–69.

19. Davies SW, Turmaine M, Cozens BA, DiFiglia M, Sharp AH, Ross CA, et al. Formation ofneuronal intranuclear inclusions underlies the neurological dysfunction in mice transgenic forthe HD mutation. Cell 1997;90:537–48.

20. Sathasivam K, Baxendale S, Mangiarini L, Bertaux F, Hetherington C, Kanazawa I, et al.Aberrant processing of the Fugu HD (FrHD) mRNA in mouse cells and in transgenic mice.Hum Mol Genet 1997;6:2141–9.

21. DiFiglia M, Sapp E, Chase K, Schwarz C, Meloni A, Young C, et al. Huntingtin is acytoplasmic protein associated with vesicles in human and rat brain neurons. Neuron1995;14:1075–81.

22. Lin B, Rommens JM, Graham RK, Kalchman M, MacDonald H, Nasir J, et al. Differential30 polyadenylation of the Huntington disease gene results in two mRNA species with variabletissue expression. Hum Mol Genet 1993;2:1541–5.

23. Lin B, Nasir J, MacDonald H, Hutchinson G, Graham RK, Rommens JM, et al. Sequence ofthe murine Huntington disease gene: evidence for conservation, alternate splicing andpolymorphism in a triplet (CCG) repeat [corrected]. Hum Mol Genet 1994;3:85–92.

24. Trottier Y, Devys D, Imbert G, Saudou F, An I, Lutz Y, et al. Cellular localization of theHuntington’s disease protein and discrimination of the normal and mutated form. Nat Genet1995;10:104–10.

25. Nasir J, Floresco SB, O’Kusky JR, Diewert VM, Richman JM, Zeisler J, et al. Targeteddisruption of the Huntington’s disease gene results in embryonic lethality and behavioraland morphological changes in heterozygotes. Cell 1995;81:811–23.

26. Duyao MP, Auerbach AB, Ryan A, Persichetti F, Barnes GT, McNeil SM, et al. Inactivation ofthe mouse Huntington’s disease gene homolog Hdh. Science 1995;269:407–10.

27. Clabough EB, Zeitlin SO. Deletion of the triplet repeat encoding polyglutamine within themouse Huntington’s disease gene results in subtle behavioral/motor phenotypes in vivo andelevated levels of ATP with cellular senescence in vitro. Hum Mol Genet 2006;15:607–23.

28. Andrade MA, Bork P. HEAT repeats in the Huntington’s disease protein. Nat Genet1995;11:115–6.

29. Tartari M, Gissi C, Lo Sardo V, Zuccato C, Picardi E, Pesole G, et al. Phylogenetic comparisonof huntingtin homologues reveals the appearance of a primitive polyQ in sea urchin.Mol BiolEvol 2008;25:330–8.

30. DiFiglia M, Sapp E, Chase KO, Davies SW, Bates GP, Vonsattel JP, et al. Aggregation ofhuntingtin in neuronal intranuclear inclusions and dystrophic neurites in brain. Science1997;277:1990–3.

31. Landles C, Sathasivam K, Weiss A, Woodman B, Moffitt H, Finkbeiner S, et al. Proteolysis ofmutant huntingtin produces an exon 1 fragment that accumulates as an aggregated protein inneuronal nuclei in Huntington disease. J Biol Chem 2010;285:8808–23.

32. Becher MW, Kotzuk JA, Sharp AH, Davies SW, Bates GP, Price DL, et al. Intranuclearneuronal inclusions in Huntington’s disease and dentatorubral and pallidoluysian atrophy:

Page 19: Huntington Diseases and the Huntingtin Protein

HUNTINGTON DISEASE AND THE HUNTINGTIN PROTEIN 207

correlation between the density of inclusions and IT15 CAG triplet repeat length. NeurobiolDis 1998;4:387–97.

33. Van Raamsdonk JM, Murphy Z, Slow EJ, Leavitt BR, Hayden MR. Selective degenerationand nuclear localization of mutant huntingtin in the YAC128 mouse model of Huntingtondisease. Hum Mol Genet 2005;14:3823–35.

34. Xia J, Lee DH, Taylor J, Vandelft M, Truant R. Huntingtin contains a highly conserved nuclearexport signal. Hum Mol Genet 2003;12:1393–403.

35. Cornett J, Cao F, Wang CE, Ross CA, Bates GP, Li SH, et al. Polyglutamine expansion ofhuntingtin impairs its nuclear export. Nat Genet 2005;37:198–204.

36. Wellington CL, Ellerby LM, Hackam AS, Margolis RL, TrifiroMA, Singaraja R, et al. Caspasecleavage of gene products associated with triplet expansion disorders generates truncatedfragments containing the polyglutamine tract. J Biol Chem 1998;273:9158–67.

37. Wellington CL, Singaraja R, Ellerby L, Savill J, Roy S, Leavitt B, et al. Inhibiting caspasecleavage of huntingtin reduces toxicity and aggregate formation in neuronal and nonneuronalcells. J Biol Chem 2000;275:19831–8.

38. Hermel E, Gafni J, Propp SS, Leavitt BR, Wellington CL, Young JE, et al. Specific caspaseinteractions and amplification are involved in selective neuronal vulnerability in Huntington’sdisease. Cell Death Differ 2004;11:424–38.

39. Gafni J, Ellerby LM. Calpain activation in Huntington’s disease. J Neurosci 2002;22:4842–9.40. Kim YJ, Yi Y, Sapp E, Wang Y, Cuiffo B, Kegel KB, et al. Caspase 3-cleaved N-terminal

fragments of wild-type and mutant huntingtin are present in normal and Huntington’s diseasebrains, associate with membranes, and undergo calpain-dependent proteolysis. Proc NatlAcad Sci USA 2001;98:12784–9.

41. Miller JP, Holcomb J, Al-Ramahi I, de Haro M, Gafni J, Zhang N, et al. Matrix metallopro-teinases are modifiers of huntingtin proteolysis and toxicity in Huntington’s disease. Neuron2010;67:199–212.

42. Graham RK, Deng Y, Slow EJ, Haigh B, Bissada N, Lu G, et al. Cleavage at the caspase-6 siteis required for neuronal dysfunction and degeneration due to mutant huntingtin. Cell2006;125:1179–91.

43. Gafni J, Hermel E, Young JE, Wellington CL, Hayden MR, Ellerby LM. Inhibition of calpaincleavage of huntingtin reduces toxicity: accumulation of calpain/caspase fragments in thenucleus. J Biol Chem 2004;279:20211–20.

44. Kalchman MA, Graham RK, Xia G, Koide HB, Hodgson JG, Graham KC, et al. Huntingtin isubiquitinated and interacts with a specific ubiquitin-conjugating enzyme. J Biol Chem1996;271:19385–94.

45. Steffan JS, Agrawal N, Pallos J, Rockabrand E, Trotman LC, Slepko N, et al. SUMOmodification of Huntingtin and Huntington’s disease pathology. Science 2004;304:100–4.

46. Aiken CT, Steffan JS, Guerrero CM, Khashwji H, Lukacsovich T, Simmons D, et al. Phos-phorylation of threonine 3: implications for huntingtin aggregation and neurotoxicity. J BiolChem 2009;284:29427–36.

47. Thompson LM, Aiken CT, Kaltenbach LS, Agrawal N, Illes K, Khoshnan A, et al. IKKphosphorylates Huntingtin and targets it for degradation by the proteasome and lysosome.J Cell Biol 2009;187:1083–99.

48. Gu X, Greiner ER, Mishra R, Kodali R, Osmand A, Finkbeiner S, et al. Serines 13 and 16 arecritical determinants of full-length human mutant huntingtin induced disease pathogenesis inHD mice. Neuron 2009;64:828–40.

49. Humbert S, Bryson EA, Cordelieres FP, Connors NC, Datta SR, Finkbeiner S, et al. TheIGF-1/Akt pathway is neuroprotective in Huntington’s disease and involves Huntingtinphosphorylation by Akt. Dev Cell 2002;2:831–7.

Page 20: Huntington Diseases and the Huntingtin Protein

208 ZHENG AND DIAMOND

50. Luo S, Vacher C, Davies JE, Rubinsztein DC. Cdk5 phosphorylation of huntingtin reduces itscleavage by caspases: implications for mutant huntingtin toxicity. J Cell Biol 2005;169:647–56.

51. Warby SC, Chan EY, Metzler M, Gan L, Singaraja RR, Crocker SF, et al. Huntingtinphosphorylation on serine 421 is significantly reduced in the striatum and by polyglutamineexpansion in vivo. Hum Mol Genet 2005;14:1569–77.

52. Anne SL, Saudou F, Humbert S. Phosphorylation of huntingtin by cyclin-dependent kinase 5is induced by DNA damage and regulates wild-type and mutant huntingtin toxicity inneurons. J Neurosci 2007;27:7318–28.

53. Jeong H, Then F, Melia Jr. TJ, Mazzulli JR, Cui L, Savas JN, et al. Acetylation targets mutanthuntingtin to autophagosomes for degradation. Cell 2009;137:60–72.

54. Yanai A, Huang K, Kang R, Singaraja RR, Arstikaitis P, Gan L, et al. Palmitoylationof huntingtin by HIP14 is essential for its trafficking and function. Nat Neurosci2006;9:824–31.

55. Rigamonti D, Bauer JH, De-Fraja C, Conti L, Sipione S, Sciorati C, et al. Wild-typehuntingtin protects from apoptosis upstream of caspase-3. J Neurosci 2000;20:3705–13.

56. Zhang Y, Li M, Drozda M, Chen M, Ren S, Mejia Sanchez RO, et al. Depletion of wild-typehuntingtin in mouse models of neurologic diseases. J Neurochem 2003;87:101–6.

57. Leavitt BR, van Raamsdonk JM, Shehadeh J, Fernandes H, Murphy Z, Graham RK, et al.Wild-type huntingtin protects neurons from excitotoxicity. J Neurochem 2006;96:1121–9.

58. Gervais FG, Singaraja R, Xanthoudakis S, Gutekunst CA, Leavitt BR, Metzler M, et al.Recruitment and activation of caspase-8 by the Huntingtin-interacting protein Hip-1 and anovel partner Hippi. Nat Cell Biol 2002;4:95–105.

59. Rigamonti D, Sipione S, Goffredo D, Zuccato C, Fossale E, Cattaneo E. Huntingtin’sneuroprotective activity occurs via inhibition of procaspase-9 processing. J Biol Chem2001;276:14545–8.

60. Zuccato C, Ciammola A, Rigamonti D, Leavitt BR, Goffredo D, Conti L, et al. Loss ofhuntingtin-mediated BDNF gene transcription in Huntington’s disease. Science2001;293:493–8.

61. Zuccato C, Cattaneo E. Brain-derived neurotrophic factor in neurodegenerative diseases.NatRev Neurol 2009;5:311–22.

62. Zuccato C, Belyaev N, Conforti P, Ooi L, Tartari M, Papadimou E, et al. Widespreaddisruption of repressor element-1 silencing transcription factor/neuron-restrictive silencerfactor occupancy at its target genes in Huntington’s disease. J Neurosci 2007;27:6972–83.

63. Zuccato C, Tartari M, Crotti A, Goffredo D, Valenza M, Conti L, et al. Huntingtin interactswith REST/NRSF to modulate the transcription of NRSE-controlled neuronal genes. NatGenet 2003;35:76–83.

64. Gauthier LR, Charrin BC, Borrell-Pages M, Dompierre JP, Rangone H, Cordelieres FP, et al.Huntingtin controls neurotrophic support and survival of neurons by enhancing BDNFvesicular transport along microtubules. Cell 2004;118:127–38.

65. Her LS, Goldstein LS. Enhanced sensitivity of striatal neurons to axonal transport defectsinduced by mutant huntingtin. J Neurosci 2008;28:13662–72.

66. Smith R, Brundin P, Li JY. Synaptic dysfunction in Huntington’s disease: a new perspective.Cell Mol Life Sci 2005;62:1901–12.

67. Faber PW, Alter JR, MacDonald ME, Hart AC. Polyglutamine-mediated dysfunction andapoptotic death of a Caenorhabditis elegans sensory neuron. Proc Natl Acad Sci USA1999;96:179–84.

68. Parker JA, Connolly JB, Wellington C, Hayden M, Dausset J, Neri C. Expandedpolyglutamines in Caenorhabditis elegans cause axonal abnormalities and severe dysfunctionof PLM mechanosensory neurons without cell death. Proc Natl Acad Sci USA2001;98:13318–23.

Page 21: Huntington Diseases and the Huntingtin Protein

HUNTINGTON DISEASE AND THE HUNTINGTIN PROTEIN 209

69. Gunawardena S, Her LS, Brusch RG, Laymon RA, Niesman IR, Gordesky-Gold B, et al.Disruption of axonal transport by loss of huntingtin or expression of pathogenic polyQproteins in Drosophila. Neuron 2003;40:25–40.

70. Marsh JL, Thompson LM. Drosophila in the study of neurodegenerative disease. Neuron2006;52:169–78.

71. Miller VM, Nelson RF, Gouvion CM, Williams A, Rodriguez-Lebron E, Harper SQ, et al.CHIP suppresses polyglutamine aggregation and toxicity in vitro and in vivo. J Neurosci2005;25:9152–61.

72. Mangiarini L, Sathasivam K, Seller M, Cozens B, Harper A, Hetherington C, et al. Exon 1 ofthe HD gene with an expanded CAG repeat is sufficient to cause a progressive neurologicalphenotype in transgenic mice. Cell 1996;87:493–506.

73. Schilling G, Becher MW, Sharp AH, Jinnah HA, Duan K, Kotzuk JA, et al. Intranuclearinclusions and neuritic aggregates in transgenic mice expressing a mutant N-terminal frag-ment of huntingtin. Hum Mol Genet 1999;8:397–407.

74. Hodgson JG, Agopyan N, Gutekunst CA, Leavitt BR, LePiane F, Singaraja R, et al. A YACmouse model for Huntington’s disease with full-length mutant huntingtin, cytoplasmic toxici-ty, and selective striatal neurodegeneration. Neuron 1999;23:181–92.

75. Slow EJ, van Raamsdonk J, Rogers D, Coleman SH, Graham RK, Deng Y, et al. Selectivestriatal neuronal loss in a YAC128 mouse model of Huntington disease. Hum Mol Genet2003;12:1555–67.

76. Gray M, Shirasaki DI, Cepeda C, Andre VM, Wilburn B, Lu XH, et al. Full-length humanmutant huntingtin with a stable polyglutamine repeat can elicit progressive and selectiveneuropathogenesis in BACHD mice. J Neurosci 2008;28:6182–95.

77. Wheeler VC, White JK, Gutekunst CA, Vrbanac V, Weaver M, Li XJ, et al. Long glutaminetracts cause nuclear localization of a novel form of huntingtin in medium spiny striatal neuronsin HdhQ92 and HdhQ111 knock-in mice. Hum Mol Genet 2000;9:503–13.

78. Lin CH, Tallaksen-Greene S, Chien WM, Cearley JA, Jackson WS, Crouse AB, et al. Neuro-logical abnormalities in a knock-in mouse model of Huntington’s disease. Hum Mol Genet2001;10:137–44.

79. Menalled LB, Sison JD, Dragatsis I, Zeitlin S, Chesselet MF. Time course of early motor andneuropathological anomalies in a knock-in mouse model of Huntington’s disease with 140CAG repeats. J Comp Neurol 2003;465:11–26.

80. von Horsten S, Schmitt I, Nguyen HP, Holzmann C, Schmidt T, Walther T, et al. Transgenicrat model of Huntington’s disease. Hum Mol Genet 2003;12:617–24.

81. Yang SH, Cheng PH, Banta H, Piotrowska-Nitsche K, Yang JJ, Cheng EC, et al.Towards a transgenic model of Huntington’s disease in a non-human primate. Nature2008;453:921–4.

82. Wellington CL, Ellerby LM, Gutekunst CA, Rogers D, Warby S, Graham RK, et al. Caspasecleavage of mutant huntingtin precedes neurodegeneration in Huntington’s disease. J Neu-rosci 2002;22:7862–72.

83. Sathasivam K, Lane A, Legleiter J, Warley A, Woodman B, Finkbeiner S, et al. Identicaloligomeric and fibrillar structures captured from the brains of R6/2 and knock-in mousemodels of Huntington’s disease. Hum Mol Genet 2010;19:65–78.

84. Legleiter J, Mitchell E, Lotz GP, Sapp E, Ng C, DiFiglia M, et al. Mutant huntingtinfragments form oligomers in a polyglutamine length-dependent manner in vitro and in vivo.J Biol Chem 2010;285:14777–90.

85. Sapp E, Schwarz C, Chase K, Bhide PG, Young AB, Penney J, et al. Huntingtin localization inbrains of normal and Huntington’s disease patients. Ann Neurol 1997;42:604–12.

86. Wanker EE. Protein aggregation and pathogenesis of Huntington’s disease: mechanisms andcorrelations. Biol Chem 2000;381:937–42.

Page 22: Huntington Diseases and the Huntingtin Protein

210 ZHENG AND DIAMOND

87. Martindale D, Hackam A, Wieczorek A, Ellerby L, Wellington C, McCutcheon K, et al.Length of huntingtin and its polyglutamine tract influences localization and frequency ofintracellular aggregates. Nat Genet 1998;18:150–4.

88. Hackam AS, Singaraja R, Wellington CL, Metzler M, McCutcheon K, Zhang TQ, et al. Theinfluence of Huntingtin protein size on nuclear localization and cellular toxicity. J Cell Biol1998;141:1097–105.

89. Li SH, Li XJ. Aggregation of N-terminal huntingtin is dependent on the length of itsglutamine repeats. Hum Mol Genet 1998;7:777–82.

90. Chen SM, Ferrone FA, Wetzel R. Huntington’s disease age-of-onset linked to polyglutamineaggregation nucleation. Proc Natl Acad Sci USA 2002;99:11884–9.

91. Zuccato C, Valenza M, Cattaneo E. Molecular mechanisms and potential therapeutical targetsin Huntington’s disease. Physiol Rev 2010;90:905–81.

92. Ross CA, Poirier MA. Opinion: what is the role of protein aggregation in neurodegeneration?Nat Rev Mol Cell Biol 2005;6:891–8.

93. Arrasate M, Mitra S, Schweitzer ES, Segal MR, Finkbeiner S. Inclusion body formationreduces levels of mutant huntingtin and the risk of neuronal death. Nature 2004;431:805–10.

94. Yamamoto A, Lucas JJ, Hen R. Reversal of neuropathology and motor dysfunction in aconditional model of Huntington’s disease. Cell 2000;101:57–66.

95. Poirier MA, Li H, Macosko J, Cai S, Amzel M, Ross CA. Huntingtin spheroids and protofibrilsas precursors in polyglutamine fibrilization. J Biol Chem 2002;277:41032–7.

96. Wacker JL, Zareie MH, Fong H, Sarikaya M, Muchowski PJ. Hsp70 and Hsp40 attenuateformation of spherical and annular polyglutamine oligomers by partitioning monomer. NatStruct Mol Biol 2004;11:1215–22.

97. Dunah AW, Jeong H, Griffin A, Kim YM, Standaert DG, Hersch SM, et al. Sp1 and TAFII130transcriptional activity disrupted in early Huntington’s disease. Science 2002;296:2238–43.

98. Okazawa H. Polyglutamine diseases: a transcription disorder? Cell Mol Life Sci2003;60:1427–39.

99. Nucifora Jr. FC, Sasaki M, Peters MF, Huang H, Cooper JK, YamadaM, et al. Interference byhuntingtin and atrophin-1 with cbp-mediated transcription leading to cellular toxicity. Science2001;291:2423–8.

100. Cui L, Jeong H, Borovecki F, Parkhurst CN, Tanese N, Krainc D. Transcriptional repressionof PGC-1alpha by mutant huntingtin leads to mitochondrial dysfunction and neurodegenera-tion. Cell 2006;127:59–69.

101. Trushina E, McMurray CT. Oxidative stress and mitochondrial dysfunction in neurodegener-ative diseases. Neuroscience 2007;145:1233–48.

102. Moreira PI, Zhu X, Wang X, Lee HG, Nunomura A, Petersen RB, et al. Mitochondria: atherapeutic target in neurodegeneration. Biochim Biophys Acta 2010;1802:212–20.

103. Perez-Severiano F, Rios C, Segovia J. Striatal oxidative damage parallels the expression of aneurological phenotype in mice transgenic for the mutation of Huntington’s disease. BrainRes 2000;862:234–7.

104. Polidori MC, Mecocci P, Browne SE, Senin U, Beal MF. Oxidative damage to mitochondrialDNA in Huntington’s disease parietal cortex. Neurosci Lett 1999;272:53–6.

105. Bogdanov MB, Andreassen OA, Dedeoglu A, Ferrante RJ, Beal MF. Increased oxidativedamage to DNA in a transgenic mouse model of Huntington’s disease. J Neurochem2001;79:1246–9.

106. Browne SE, Bowling AC, MacGarvey U, Baik MJ, Berger SC, Muqit MM, et al. Oxidativedamage and metabolic dysfunction in Huntington’s disease: selective vulnerability of the basalganglia. Ann Neurol 1997;41:646–53.

107. Browne SE, Ferrante RJ, Beal MF. Oxidative stress in Huntington’s disease. Brain Pathol1999;9:147–63.

Page 23: Huntington Diseases and the Huntingtin Protein

HUNTINGTON DISEASE AND THE HUNTINGTIN PROTEIN 211

108. Brustovetsky N, LaFrance R, Purl KJ, Brustovetsky T, Keene CD, Low WC, et al. Age-dependent changes in the calcium sensitivity of striatal mitochondria in mouse models ofHuntington’s disease. J Neurochem 2005;93:1361–70.

109. Panov AV, Gutekunst CA, Leavitt BR, Hayden MR, Burke JR, Strittmatter WJ, et al. Earlymitochondrial calcium defects in Huntington’s disease are a direct effect of polyglutamines.Nat Neurosci 2002;5:731–6.

110. Choo YS, Johnson GV, MacDonald M, Detloff PJ, Lesort M. Mutant huntingtin directlyincreases susceptibility of mitochondria to the calcium-induced permeability transition andcytochrome c release. Hum Mol Genet 2004;13:1407–20.

111. Orr AL, Li S, Wang CE, Li H, Wang J, Rong J, et al. N-terminal mutant huntingtin associateswith mitochondria and impairs mitochondrial trafficking. J Neurosci 2008;28:2783–92.

112. Seong IS, Ivanova E, Lee JM, Choo YS, Fossale E, Anderson M, et al. HD CAG repeatimplicates a dominant property of huntingtin in mitochondrial energy metabolism. Hum MolGenet 2005;14:2871–80.

113. Jenkins BG, Koroshetz WJ, Beal MF, Rosen BR. Evidence for impairment of energy metabo-lism in vivo in Huntington’s disease using localized 1H NMR spectroscopy. Neurology1993;43:2689–95.

114. Arenas J, Campos Y, Ribacoba R,Martin MA, Rubio JC, Ablanedo P, et al. Complex I defect inmuscle from patients with Huntington’s disease. Ann Neurol 1998;43:397–400.

115. Browne SE. Mitochondria and Huntington’s disease pathogenesis: insight from genetic andchemical models. Ann N Y Acad Sci 2008;1147:358–82.

116. Gu M, Gash MT, Mann VM, Javoy-Agid F, Cooper JM, Schapira AH. Mitochondrial defect inHuntington’s disease caudate nucleus. Ann Neurol 1996;39:385–9.

117. Milakovic T, Quintanilla RA, Johnson GV. Mutant huntingtin expression induces mitochon-drial calcium handling defects in clonal striatal cells: functional consequences. J Biol Chem2006;281:34785–95.

118. Bence NF, Sampat RM, Kopito RR. Impairment of the ubiquitin-proteasome system byprotein aggregation. Science 2001;292:1552–5.

119. Holmberg CI, Staniszewski KE, Mensah KN, Matouschek A, Morimoto RI. Inefficientdegradation of truncated polyglutamine proteins by the proteasome. EMBO J2004;23:4307–18.

120. Gidalevitz T, Ben-Zvi A, Ho KH, Brignull HR, Morimoto RI. Progressive disruption ofcellular protein folding in models of polyglutamine diseases. Science 2006;311:1471–4.

121. Imarisio S, Carmichael J, Korolchuk V, Chen CW, Saiki S, Rose C, et al. Huntington’s disease:from pathology and genetics to potential therapies. Biochem J 2008;412:191–209.

122. Venkatraman P, Wetzel R, Tanaka M, Nukina N, Goldberg AL. Eukaryotic proteasomescannot digest polyglutamine sequences and release them during degradation ofpolyglutamine-containing proteins. Mol Cell 2004;14:95–104.

123. Carmichael J, Chatellier J, Woolfson A, Milstein C, Fersht AR, Rubinsztein DC. Bacterial andyeast chaperones reduce both aggregate formation and cell death in mammalian cell modelsof Huntington’s disease. Proc Natl Acad Sci USA 2000;97:9701–5.

124. Warrick JM, Chan HY, Gray-Board GL, Chai Y, Paulson HL, Bonini NM. Suppression ofpolyglutamine-mediated neurodegeneration in Drosophila by the molecular chaperoneHSP70. Nat Genet 1999;23:425–8.

125. Wyttenbach A, Carmichael J, Swartz J, Furlong RA, Narain Y, Rankin J, et al. Effects of heatshock, heat shock protein 40 (HDJ-2), and proteasome inhibition on protein aggregation incellular models of Huntington’s disease. Proc Natl Acad Sci USA 2000;97:2898–903.

126. Ravikumar B, Vacher C, Berger Z, Davies JE, Luo S, Oroz LG, et al. Inhibition of mTORinduces autophagy and reduces toxicity of polyglutamine expansions in fly and mouse modelsof Huntington disease. Nat Genet 2004;36:585–95.

Page 24: Huntington Diseases and the Huntingtin Protein

212 ZHENG AND DIAMOND

127. Sarkar S, Rubinsztein DC. Huntington’s disease: degradation of mutant huntingtin by autop-hagy. FEBS J 2008;275:4263–70.

128. Jia K, Hart AC, Levine B. Autophagy genes protect against disease caused by polyglutamineexpansion proteins in Caenorhabditis elegans. Autophagy 2007;3:21–5.

129. Xie Y, HaydenMR, Xu B. BDNF overexpression in the forebrain rescues Huntington’s diseasephenotypes in YAC128 mice. J Neurosci 2010;30:14708–18.

130. Lynch G, Kramar EA, Rex CS, Jia Y, Chappas D, Gall CM, et al. Brain-derived neurotrophicfactor restores synaptic plasticity in a knock-in mouse model of Huntington’s disease.J Neurosci 2007;27:4424–34.

131. St-Pierre J, Drori S, Uldry M, Silvaggi JM, Rhee J, Jager S, et al. Suppression of reactiveoxygen species and neurodegeneration by the PGC-1 transcriptional coactivators. Cell2006;127:397–408.

132. Weydt P, Pineda VV, Torrence AE, Libby RT, Satterfield TF, Lazarowski ER, et al.Thermoregulatory and metabolic defects in Huntington’s disease transgenic miceimplicate PGC-1alpha in Huntington’s disease neurodegeneration. Cell Metab2006;4:349–62.

133. Minamiyama M, Katsuno M, Adachi H, Waza M, Sang C, Kobayashi Y, et al. Sodium butyrateameliorates phenotypic expression in a transgenic mouse model of spinal and bulbar muscularatrophy. Hum Mol Genet 2004;13:1183–92.

134. Hockly E, Richon VM, Woodman B, Smith DL, Zhou X, Rosa E, et al. Suberoylanilidehydroxamic acid, a histone deacetylase inhibitor, ameliorates motor deficits in a mousemodel of Huntington’s disease. Proc Natl Acad Sci USA 2003;100:2041–6.

135. Ferrante RJ, Kubilus JK, Lee J, Ryu H, Beesen A, Zucker B, et al. Histone deacetylaseinhibition by sodium butyrate chemotherapy ameliorates the neurodegenerative phenotype inHuntington’s disease mice. J Neurosci 2003;23:9418–27.

136. Gardian G, Browne SE, Choi DK, Klivenyi P, Gregorio J, Kubilus JK, et al. Neuroprotectiveeffects of phenylbutyrate in the N171-82Q transgenic mouse model of Huntington’s disease.J Biol Chem 2005;280:556–63.

137. Steffan JS, Bodai L, Pallos J, Poelman M, McCampbell A, Apostol BL, et al. Histonedeacetylase inhibitors arrest polyglutamine-dependent neurodegeneration in Drosophila.Nature 2001;413:739–43.

138. Hubbert C, Guardiola A, Shao R, Kawaguchi Y, Ito A, Nixon A, et al. HDAC6 is amicrotubule-associated deacetylase. Nature 2002;417:455–8.

139. Kovacs JJ, Murphy PJ, Gaillard S, Zhao X, Wu JT, Nicchitta CV, et al. HDAC6 regulatesHsp90 acetylation and chaperone-dependent activation of glucocorticoid receptor. Mol Cell2005;18:601–7.

140. Iwata A, Riley BE, Johnston JA, Kopito RR. HDAC6 and microtubules are required forautophagic degradation of aggregated huntingtin. J Biol Chem 2005;280:40282–92.

141. Giorgini F, Moller T, Kwan W, Zwilling D, Wacker JL, Hong S, et al. Histone deacetylaseinhibition modulates kynurenine pathway activation in yeast, microglia, and mice expressing amutant huntingtin fragment. J Biol Chem 2008;283:7390–400.

142. Andreassen OA, Dedeoglu A, Ferrante RJ, Jenkins BG, Ferrante KL, Thomas M, et al.Creatine increase survival and delays motor symptoms in a transgenic animal model ofHuntington’s disease. Neurobiol Dis 2001;8:479–91.

143. Ferrante RJ, Andreassen OA, Jenkins BG, Dedeoglu A, Kuemmerle S, Kubilus JK, et al.Neuroprotective effects of creatine in a transgenic mouse model of Huntington’s disease.J Neurosci 2000;20:4389–97.

144. Ferrante RJ, Andreassen OA, Dedeoglu A, Ferrante KL, Jenkins BG, Hersch SM, et al.Therapeutic effects of coenzyme Q10 and remacemide in transgenic mouse models ofHuntington’s disease. J Neurosci 2002;22:1592–9.

Page 25: Huntington Diseases and the Huntingtin Protein

HUNTINGTON DISEASE AND THE HUNTINGTIN PROTEIN 213

145. Schilling G, Coonfield ML, Ross CA, Borchelt DR. Coenzyme Q10 and remacemide hydro-chloride ameliorate motor deficits in a Huntington’s disease transgenic mouse model. Neu-rosci Lett 2001;315:149–53.

146. Huntington Study Group. A randomized, placebo-controlled trial of coenzyme Q10 andremacemide in Huntington’s disease. Neurology 2001;57:397–404.

147. Verbessem P, Lemiere J, Eijnde BO, Swinnen S, Vanhees L, Van LeemputteM, et al. Creatinesupplementation in Huntington’s disease: a placebo-controlled pilot trial. Neurology2003;61:925–30.

148. Chen M, Ona VO, Li M, Ferrante RJ, Fink KB, Zhu S, et al. Minocycline inhibits caspase-1and caspase-3 expression and delays mortality in a transgenic mouse model of Huntingtondisease. Nat Med 2000;6:797–801.

149. Wang X, Zhu S, Drozda M, ZhangW, Stavrovskaya IG, Cattaneo E, et al. Minocycline inhibitscaspase-independent and -dependent mitochondrial cell death pathways in models of Hun-tington’s disease. Proc Natl Acad Sci USA 2003;100:10483–7.

150. Rubinsztein DC, Gestwicki JE, Murphy LO, Klionsky DJ. Potential therapeutic applicationsof autophagy. Nat Rev Drug Discov 2007;6:304–12.

151. Ravikumar B, Duden R, Rubinsztein DC. Aggregate-prone proteins with polyglutamine andpolyalanine expansions are degraded by autophagy. Hum Mol Genet 2002;11:1107–17.

152. Berger Z, Ravikumar B, Menzies FM, Oroz LG, Underwood BR, Pangalos MN, et al.Rapamycin alleviates toxicity of different aggregate-prone proteins. Hum Mol Genet2006;15:433–42.

153. Coufal M, Maxwell MM, Russel DE, Amore AM, Altmann SM, Hollingsworth ZR, et al.Discovery of a novel small-molecule targeting selective clearance of mutant huntingtinfragments. J Biomol Screen 2007;12:351–60.

154. Sarkar S, Perlstein EO, Imarisio S, Pineau S, Cordenier A, Maglathlin RL, et al. Smallmolecules enhance autophagy and reduce toxicity in Huntington’s disease models. NatChem Biol 2007;3:331–8.

155. Sarkar S, Krishna G, Imarisio S, Saiki S, O’Kane CJ, Rubinsztein DC. A rational mechanismfor combination treatment of Huntington’s disease using lithium and rapamycin. Hum MolGenet 2008;17:170–8.

156. Heiser V, Scherzinger E, Boeddrich A, Nordhoff E, Lurz R, Schugardt N, et al. Inhibition ofhuntingtin fibrillogenesis by specific antibodies and small molecules: implications for Hun-tington’s disease therapy. Proc Natl Acad Sci USA 2000;97:6739–44.

157. Zhang X, Smith DL, Meriin AB, Engemann S, Russel DE, Roark M, et al. A potent smallmolecule inhibits polyglutamine aggregation in Huntington’s disease neurons and suppressesneurodegeneration in vivo. Proc Natl Acad Sci USA 2005;102:892–7.

158. Desai UA, Pallos J, Ma AA, Stockwell BR, Thompson LM, Marsh JL, et al. Biologically activemolecules that reduce polyglutamine aggregation and toxicity. Hum Mol Genet2006;15:2114–24.

159. Pollitt SK, Pallos J, Shao J, Desai UA, Ma AA, Thompson LM, et al. A rapid cellularFRET assay of polyglutamine aggregation identifies a novel inhibitor. Neuron2003;40:685–94.

160. Li M, Huang Y, Ma AA, Lin E, Diamond MI. Y-27632 improves rotarod performance andreduces huntingtin levels in R6/2 mice. Neurobiol Dis 2009;36:413–20.

161. Tanaka M, Machida Y, Niu S, Ikeda T, Jana NR, Doi H, et al. Trehalose alleviatespolyglutamine-mediated pathology in a mouse model of Huntington disease. Nat Med2004;10:148–54.

162. Sittler A, Lurz R, Lueder G, Priller J, Lehrach H, Hayer-Hartl MK, et al. Geldanamycinactivates a heat shock response and inhibits huntingtin aggregation in a cell culture model ofHuntington’s disease. Hum Mol Genet 2001;10:1307–15.

Page 26: Huntington Diseases and the Huntingtin Protein

214 ZHENG AND DIAMOND

163. Hay DG, Sathasivam K, Tobaben S, Stahl B, Marber M, Mestril R, et al. Progressive decreasein chaperone protein levels in a mouse model of Huntington’s disease and induction of stressproteins as a therapeutic approach. Hum Mol Genet 2004;13:1389–405.

164. Katsuno M, Sang C, Adachi H, Minamiyama M, Waza M, Tanaka F, et al. Pharmacologicalinduction of heat-shock proteins alleviates polyglutamine-mediated motor neuron disease.Proc Natl Acad Sci USA 2005;102:16801–6.

165. Xia H, Mao Q, Eliason SL, Harper SQ, Martins IH, Orr HT, et al. RNAi suppressespolyglutamine-induced neurodegeneration in a model of spinocerebellar ataxia. Nat Med2004;10:816–20.

166. Harper SQ, Staber PD, He X, Eliason SL, Martins IH, Mao Q, et al. RNA interferenceimproves motor and neuropathological abnormalities in a Huntington’s disease mouse model.Proc Natl Acad Sci USA 2005;102:5820–5.

167. Hu J, Matsui M, Gagnon KT, Schwartz JC, Gabillet S, Arar K, et al. Allele-specific silencing ofmutant huntingtin and ataxin-3 genes by targeting expanded CAG repeats in mRNAs. NatBiotechnol 2009;27:478–84.

168. Denovan-Wright EM, Rodriguez-Lebron E, Lewin AS, Mandel RJ. Unexpected off-targetingeffects of anti-huntingtin ribozymes and siRNA in vivo. Neurobiol Dis 2008;29:446–55.

169. Jackson AL, Bartz SR, Schelter J, Kobayashi SV, Burchard J, Mao M, et al. Expressionprofiling reveals off-target gene regulation by RNAi. Nat Biotechnol 2003;21:635–7.

170. Colby DW, Chu Y, Cassady JP, Duennwald M, Zazulak H, Webster JM, et al. Potent inhibitionof huntingtin aggregation and cytotoxicity by a disulfide bond-free single-domain intracellularantibody. Proc Natl Acad Sci USA 2004;101:17616–21.

171. Khoshnan A, Ko J, Patterson PH. Effects of intracellular expression of anti-huntingtin anti-bodies of various specificities on mutant huntingtin aggregation and toxicity. Proc Natl AcadSci USA 2002;99:1002–7.

172. McLear JA, Lebrecht D, Messer A, Wolfgang WJ. Combinational approach of intrabody withenhanced Hsp70 expression addresses multiple pathologies in a fly model of Huntington’sdisease. FASEB J 2008;22:2003–11.

173. Wang CE, Zhou H, McGuire JR, Cerullo V, Lee B, Li SH, et al. Suppression of neuropilaggregates and neurological symptoms by an intracellular antibody implicates the cytoplasmictoxicity of mutant huntingtin. J Cell Biol 2008;181:803–16.

174. Kazantsev A, Walker HA, Slepko N, Bear JE, Preisinger E, Steffan JS, et al. A bivalentHuntingtin binding peptide suppresses polyglutamine aggregation and pathogenesis in Dro-sophila. Nat Genet 2002;30:367–76.