12
NATURE REVIEWS | NEUROLOGY VOLUME 7 | OCTOBER 2011 | 561 Centre for Molecular Medicine and Therapeutics, University of British Columbia, 950 West 28 th Avenue, Vancouver, BC V5Z 4H4, Canada (J. M. Karasinska, M. R. Hayden). Correspondence to: M. R. Hayden [email protected] Cholesterol metabolism in Huntington disease Joanna M. Karasinska and Michael R. Hayden Abstract | The CNS is rich in cholesterol, which is essential for neuronal development and survival, synapse maturation, and optimal synaptic activity. Alterations in brain cholesterol homeostasis are linked to neurodegeneration. Studies have demonstrated that Huntington disease (HD), a progressive and fatal neurodegenerative disorder resulting from polyglutamine expansion in the huntingtin protein, is associated with changes in cellular cholesterol metabolism. Emerging evidence from human and animal studies indicates that attenuated brain sterol synthesis and accumulation of cholesterol in neuronal membranes represent two distinct mechanisms occurring in the presence of mutant huntingtin that influence neuronal survival. Increased knowledge of how changes in intraneuronal cholesterol metabolism influence the pathogenesis of HD will provide insights into the potential application of brain cholesterol regulation as a therapeutic strategy for this devastating disease. Karasinska, J. M. & Hayden, M. R. Nat. Rev. Neurol. 7, 561–572 (2011); published online 6 September 2011; doi:10.1038/nrneurol.2011.132 Introduction Huntington disease (HD) is an autosomal dominant dis- ease that leads to progressive neurodegeneration. The molecular cause of HD is the expansion of a CAG tri- nucleotide repeat in the gene encoding huntingtin (HTT), resulting in a polyglutamine stretch in the N-terminus of the protein. The neuropathology of HD is characterized by the selective loss of medium spiny neurons from the striatum, leading to neuropsychiatric changes and move- ment disorder, including chorea. 1 HD pathogenesis has been under intense investigation since the discovery of HTT mutations around 20 years ago. Various cellu- lar mechanisms underlying neuronal death have been identified, including N-methyl-d-aspartate receptor (NMDAR)-mediated excitotoxicity, 2,3 impaired axonal transport, 4,5 and caspase activation, 6 and these path- ways are the focus of research into potential therapies. Recently, disturbances in cholesterol homeostasis have been described in patients with HD and animal models of the disease, raising the question of how changes in cholesterol metabolism might influence the development of HD. Cholesterol is critical to brain development, myeli- nation, and neuronal signaling and survival, and the maintenance of balanced cholesterol homeostasis is an important aspect of CNS function. 7,8 The significance of brain cholesterol is emphasized by the detrimental impact of genetic defects affecting cholesterol synthesis or cellular transport on brain development and neurodegeneration, as manifest in Smith–Lemli–Opitz Syndrome (SLOS) and Niemann–Pick type C (NPC) disease. In SLOS, loss of enzymatic activity of 7-dehydrocholesterol (7-DHC) reductase leads to failure of cholesterol synthesis, resulting in brain malformations and impaired cognitive function, among other symptoms. 9 NPC disease is characterized by loss-of-function mutations in the gene encoding NPC1, a protein involved in lysosomal cholesterol trans- port. This defect leads to progressive neurodegenera- tion. 10 Disturbances in cellular cholesterol metabolism, plasma membrane cholesterol composition and inter- cellular transport have also been linked to the etiology of Alzheimer disease (AD; Table 1). 11,12 Cellular cholesterol is derived from de novo synthesis and the diet. Organ and blood cholesterol levels are maintained by a tightly regulated balance of synthesis, catabolism, excretion and uptake. 13 However, tissue and—in particular—plasma cholesterol levels can be dramatically disturbed by mutations in genes encoding proteins involved in one of the cholesterol metabolic pathways, or by high dietary cholesterol intake. For example, loss-of-function mutations in the gene encod- ing ATP-binding cassette transporter A1 (ABCA1), which facilitates cellular cholesterol efflux to apolipo- protein acceptors, reduce plasma HDL levels by up to 95%. 14 Loss-of-function mutations in the PCSK9 gene, which encodes proprotein convertase subtilisin/kexin type 9 serine protease, lead to a decrease of up to 40% in plasma levels of LDL. 15 A Western-type diet high in fat raises plasma LDL. Fluctuations in blood and tissue cholesterol levels are evident in various human diseases, including atherosclerosis and diabetes. 16,17 By understanding how changes in cellular cholesterol homeostasis affect CNS diseases, including HD, we will substantially enhance our knowledge of the role of cholesterol in influencing the delicate balance between life and death in neuronal cells, in addition to identify- ing potential new avenues for treatment. This Review will summarize the roles of cholesterol in the CNS and describe the current state of knowledge of cholesterol dysregulation in HD. Competing interests The authors declare no competing interests. REVIEWS © 2011 Macmillan Publishers Limited. All rights reserved

Cholesterol Metabolism in Huntington Disease

Embed Size (px)

DESCRIPTION

article

Citation preview

Page 1: Cholesterol Metabolism in Huntington Disease

NATURE REVIEWS | NEUROLOGY VOLUME 7 | OCTOBER 2011 | 561

Centre for Molecular Medicine and Therapeutics, University of British Columbia, 950 West 28th Avenue, Vancouver, BC V5Z 4H4, Canada (J. M. Karasinska, M. R. Hayden).

Correspondence to: M. R. Hayden [email protected]

Cholesterol metabolism in Huntington diseaseJoanna M. Karasinska and Michael R. Hayden

Abstract | The CNS is rich in cholesterol, which is essential for neuronal development and survival, synapse maturation, and optimal synaptic activity. Alterations in brain cholesterol homeostasis are linked to neurodegeneration. Studies have demonstrated that Huntington disease (HD), a progressive and fatal neurodegenerative disorder resulting from polyglutamine expansion in the huntingtin protein, is associated with changes in cellular cholesterol metabolism. Emerging evidence from human and animal studies indicates that attenuated brain sterol synthesis and accumulation of cholesterol in neuronal membranes represent two distinct mechanisms occurring in the presence of mutant huntingtin that influence neuronal survival. Increased knowledge of how changes in intraneuronal cholesterol metabolism influence the pathogenesis of HD will provide insights into the potential application of brain cholesterol regulation as a therapeutic strategy for this devastating disease.

Karasinska, J. M. & Hayden, M. R. Nat. Rev. Neurol. 7, 561–572 (2011); published online 6 September 2011; doi:10.1038/nrneurol.2011.132

IntroductionHuntington disease (HD) is an autosomal dominant dis­ease that leads to progressive neurodegeneration. The molecular cause of HD is the expansion of a CAG tri­nucleo tide repeat in the gene encoding huntingtin (HTT), resulting in a polyglutamine stretch in the N­terminus of the protein. The neuropathology of HD is charac terized by the selective loss of medium spiny neurons from the striatum, leading to neuropsychiatric changes and move­ment disorder, including chorea.1 HD pathogenesis has been under intense investigation since the discovery of HTT mutations around 20 years ago. Various cellu­lar mechanisms underlying neuronal death have been identified, including N­methyl­d­aspartate receptor (NMDAR)­mediated excitotoxicity,2,3 impaired axonal transport,4,5 and caspase activation,6 and these path­ways are the focus of research into potential therapies. Recently, disturbances in cholesterol homeostasis have been described in patients with HD and animal models of the disease, raising the question of how changes in cholesterol metabolism might influence the development of HD.

Cholesterol is critical to brain development, myeli­nation, and neuronal signaling and survival, and the maintenance of balanced cholesterol homeostasis is an important aspect of CNS function.7,8 The significance of brain cholesterol is emphasized by the detrimental impact of genetic defects affecting cholesterol synthesis or cellular transport on brain development and neuro degeneration, as manifest in Smith–Lemli–Opitz Syndrome (SLOS) and Niemann–Pick type C (NPC) disease. In SLOS, loss of enzymatic activity of 7­ dehydrocholesterol (7­DHC) reductase leads to failure of cholesterol synthesis, resulting in brain malformations and impaired cognitive function,

among other symptoms.9 NPC disease is characterized by loss­of­function mutations in the gene encoding NPC1, a protein involved in lysosomal cholesterol trans­port. This defect leads to progressive neurodegenera­tion.10 Disturbances in cellular cholesterol metabolism, plasma membrane cholesterol composition and inter­cellular transport have also been linked to the etiology of Alzheimer disease (AD; Table 1).11,12

Cellular cholesterol is derived from de novo syn thesis and the diet. Organ and blood cholesterol levels are maintained by a tightly regulated balance of syn thesis, catabolism, excretion and uptake.13 However, tissue and—in particular—plasma cholesterol levels can be dramatically disturbed by mutations in genes encoding proteins involved in one of the cholesterol metabolic pathways, or by high dietary cholesterol intake. For example, loss­of­function mutations in the gene encod­ing ATP­binding cassette transporter A1 (ABCA1), which facilitates cellular cholesterol efflux to apolipo­protein acceptors, reduce plasma HDL levels by up to 95%.14 Loss­of­function mutations in the PCSK9 gene, which encodes proprotein convertase subtilisin/kexin type 9 serine protease, lead to a decrease of up to 40% in plasma levels of LDL.15 A Western­type diet high in fat raises plasma LDL. Fluctuations in blood and tissue cholesterol levels are evident in various human diseases, including atherosclerosis and diabetes.16,17

By understanding how changes in cellular cholesterol homeostasis affect CNS diseases, including HD, we will substantially enhance our knowledge of the role of cholesterol in influencing the delicate balance between life and death in neuronal cells, in addition to identify­ing potential new avenues for treatment. This Review will summarize the roles of cholesterol in the CNS and describe the current state of knowledge of cholesterol dysregulation in HD.

Competing interestsThe authors declare no competing interests.

REVIEWS

© 2011 Macmillan Publishers Limited. All rights reserved

Page 2: Cholesterol Metabolism in Huntington Disease

562 | OCTOBER 2011 | VOLUME 7 www.nature.com/nrneurol

Brain cholesterol homeostasisUnesterified cholesterol accounts for most of the brain’s sterol content, two­thirds of which is present in myelin and the rest of which resides in cellular plasma mem­branes.7,8 The brain has higher levels of cholesterol (15–20 mg/g) than any other organ, and in humans con­tains one­quarter of the total body cholesterol, despite accounting for less than 5% of the total body weight.8

The protective blood–brain barrier (BBB) consists of tight junctions between the endothelial cells of brain capillaries that severely restrict intercellular and trans­cellular transport. The BBB separates the CNS from the circulation and is impermeable to plasma components including cholesterol (Figure 1),8,18 thereby preventing fluctuations in brain cholesterol in response to chang­ing levels in the circulation. No significant cholesterol and lipoprotein transport occurs between the plasma and the brain under normal physiological conditions.8,19 How ever, the brain capillary endothelial cells express lipid transport proteins, including the cholesterol efflux transporters ABCA1 and ABCG1, the HDL receptor scavenger receptor class B member 1, and members of the LDL receptor (LDLR) family (Figure 1),20–23 and evi­dence is emerging from genetically manipulated mice that under some conditions cholesterol and lipoprotein

Key points

■ The brain is a cholesterol-rich organ, and cholesterol has a critical role in brain development, synaptogenesis, and neuronal activity and survival

■ Disturbances in brain cholesterol homeostasis are implicated in neurodegeneration

■ Attenuated plasma cholesterol levels, reduced brain cholesterol synthesis and cholesterol accumulation in neuronal plasma membranes are observed in patients with Huntington disease (HD) and in animal models of the disease

■ Multiple mechanisms are proposed to underlie cholesterol dysregulation in HD, including impaired activation of sterol regulatory element-binding protein, which regulates cholesterogenic gene expression, and reduced brain-derived neurotrophic factor signaling

■ Disturbances in cholesterol homeostasis may influence neuronal survival and susceptibility to excitotoxicity

■ Targeting of cholesterol metabolism represents a potential avenue for treatment to alleviate some neuronal impairments in HD

can be transported by the BBB.24–26 Although such transport may not markedly affect whole­brain choles­terol content, when one considers the extensive nature of the human brain vasculature (encompassing a 20 m2 surface area)18 the possibility arises that even small local­ized changes in brain cholesterol and lipoprotein uptake could have serious implications for neuronal health in affected brain areas.

Once the BBB is formed during embryonic develop­ment, the brain sterol pool becomes separated from the periphery, and brain cholesterol is synthesized in situ.7,8 All types of brain cells including neurons and glia (astrocytes, microglia and oligodendrocytes) synthe­size cholesterol during development. It has been pro­posed, however, that cholesterol synthesis is attenuated in adult neurons, which then rely mostly on astrocyte­derived cholesterol.7,27 If cholesterol synthesis is ablated in adulthood, neurons can still survive, presumably by importing cholesterol from neighbouring glia.28 By supplying cholesterol, glial cells make a vital contribu­tion to neuronal growth and activity. Astrocyte­derived cholesterol is required for optimal synaptogenesis and synaptic activity.29,30 Astrocytes synthesize and secrete cholesterol and apolipoprotein E (Apo­E) to form dis­coidal lipoprotein particles in a process regulated by the activity of ABC transporters (Figure 1). These Apo­E­containing particles are lipidated and become similar in size and density to peripheral HDL,31 and can be taken up by neurons via members of the LDLR family, which recognize Apo­E as a ligand.32 LDLR­mediated uptake of Apo­E­containing particles facilitates axonal and den­dritic growth33,34 and is implicated in the recovery from peripheral nerve injury, during which excess choles­terol from membrane debris is cleared or redistributed to regenerating nerves.35,36 In animal models of brain injury, Apo­E and ABCA1 expression37,38 is increased, presumably to facilitate this transport.

Cholesterol synthesis slows down significantly in the adult brain, and cholesterol turnover in the human brain is estimated at approximately 5 years, compared with a few months in the rodent brain.7,8 CNS choles­terol levels are strictly maintained by de novo synthesis and excretion. Excess brain cholesterol is metabolized

Table 1 | Cholesterol abnormalities in neurodegeneration

Disorder Cholesterol abnormalities Mechanisms

Smith–Lemli–Opitz syndrome

Cholesterol synthesis failure, accumulation of 7-dehydrocholesterol9

Loss-of-function mutations in 7-dehydrocholesterol reductase, an enzyme that converts 7-dehydrocholesterol to cholesterol9

Niemann–Pick type C disease

Endosomal and lysosomal cholesterol accumulation10

Loss-of-function mutations in NPC1, a protein required for transport of cholesterol from endosomes and lysosomes10

Alzheimer disease

Risk associated with plasma cholesterol profile; cholesterol accumulation in plaques; increased production of 24-hydroxycholesterol11,48

APOE ε4 allele risk factor; increased amyloid generation due to membrane cholesterol accumulation; reduced amyloid degradation due to decreased Apo-E lipidation11,12,42,43

Huntington disease

Attenuated brain sterol biosynthesis; reduced plasma cholesterol; reduced production of 24-hydroxycholesterol;51–53,59 cholesterol accumulation54

Reduced activation of SREBP leading to reduced HMG-CoAR transcription; reduced BDNF signaling;5,53 caveolin sequestration at plasma membrane; increased lipid rafts; neuronal cholesterol accumulation54,66,67

Abbreviations: Apo-E, apolipoprotein E; BDNF, brain-derived neurotrophic factor, HMG-CoAR, 3-hydroxy-3-methylglutaryl-CoA reductase, NPC1, Niemann–Pick C1 protein; SREBP, sterol regulatory element-binding protein.

REVIEWS

© 2011 Macmillan Publishers Limited. All rights reserved

Page 3: Cholesterol Metabolism in Huntington Disease

NATURE REVIEWS | NEUROLOGY VOLUME 7 | OCTOBER 2011 | 563

by the brain­specific enzyme cholesterol 24­hydroxylase (CYP46, encoded by the CYP46A1 gene) to 24­hydroxy­cholesterol (24­OHC),39 which crosses the BBB.40 CYP46 is highly expressed in neurons, suggesting that 24­OHC production is the main pathway for cholesterol excre­tion by neurons.7,8,39 The precise control of brain choles­terol content is evident in mice in which Cyp46a1 is disrupted: although CYP46­deficient mice are unable to convert cholesterol to 24­OHC, their brain cholesterol levels remain constant owing to compensatory suppres­sion of synthesis.41 The rigorous control of brain choles­terol levels demonstrates the need to maintain optimal cholesterol content and indicates that any imbalance of cholesterol homeostasis in the HD­affected brain could influence neuronal function and survival.

Cholesterol and neurodegenerationAbundant evidence links changes in cholesterol metabo­lism to neuropathological features of AD.11,12 The APOE ε4 allele is a strong risk factor for late­onset AD,42 indicat ing that pathways regulating brain lipoprotein and cholesterol transport influence the pathogenesis of AD. Degradation of amyloid­β (Aβ) peptide, the main constituent of plaques found in the brains of indivi duals with AD, is facilitated by lipidated Apo­E,43 suggesting that efficient transport of cholesterol to Apo­E helps

prevent plaque deposition. Cholesterol also influences the processing of amyloid precursor protein (APP). Low plasma membrane cholesterol is associated with reduced Aβ generation through inhibition of the activity of β­secretase, the enzyme responsible for the generation of Aβ from APP.12 ABCA1 polymorphisms are associated with an altered risk of AD.44,45 Recently, genome­wide association studies reported that variants at ABCA7, which encodes another ABC transporter expressed in the brain and involved in lipid efflux, are associated with AD,46,47 providing further evidence for lipid abnormali­ties in patients with AD. Interestingly, higher plasma HDL levels are associated with a reduced risk of develop­ing this disease.48 The mechanisms underlying the hypo­thetical link between altered plasma lipid profiles and neurodegeneration are unclear, but dyslipidemia may coexist with other risk factors for dementia.

The evidence that cholesterol abnormalities markedly influence AD underscores the importance of investigat­ing cholesterol metabolism in other neurodegenerative diseases, including HD.

Cholesterol metabolism in HDEvidence from human studiesInterest in the role of cholesterol homeostasis in HD is a relatively recent phenomenon. Early investigations in

Figure 1 | Cholesterol metabolism and transport in the CNS. Cellular cholesterol is synthesized from acetyl-CoA in a multistep pathway, with HMG-CoAR representing the rate-limiting step. In the adult brain, cholesterol synthesis is attenuated in neurons that rely on astrocyte-derived cholesterol. Cholesterol and Apo-E synthesized in astrocytes are secreted in an ABCA1-dependent process, forming discoidal lipoprotein particles, which can be further lipidated. Apo-E is a ligand for LDLR family members, which mediate neuronal lipoprotein uptake, thereby providing a supply of cholesterol to neurons. Excess cholesterol is metabolized to 24-OHC, which passes into the circulation. 24-OHC is a ligand for LXRs, which activate Apo-E and ABCA1 expression. No appreciable transfer of cholesterol from plasma HDL or LDL occurs under normal conditions, but the BCECs express cholesterol uptake proteins and transporters, including ABCA1, SRB1 and LDLR. mHtt inhibits HMGCoAR expression and activity, resulting in reduced cholesterol synthesis. Consequently, 24-OHC production might be reduced and expression of Apo-E and ABCA1 might be attenuated in an attempt to maintain cellular cholesterol levels. Abbreviations: 24-OHC, 24-hydroxycholesterol; ABCA1, ATP-binding cassette transporter A1; Apo-E, apolipoprotein E; BCEC, brain capillary endothelial cells; CYP46, cholesterol 24-hydroxylase; HMG-CoAR, 3-hydroxy-3-methylglutaryl-CoA reductase; LDLR, LDL receptor; mHtt, mutant huntingtin; LXR, liver X receptor; SRB1, scavenger receptor class B member 1.

BCEC

Brain

Blood

Astrocyte Neuron

ABCA1

ABCA1Apo-E

CYP46

Apo-Eparticle

Cerebrospinal �uid

LXR

Cholesterol

Acetyl CoA Acetyl CoA

HMG-CoAR HMG-CoAR

Cholesterol

ABCA1

HDL LDL

SRB1

SRB1LDLR

LDLR

mHtt

mHtt 24-OHC

mHtt

Indirect effect of mHttDirect effect of mHtt

Basolateral

Luminal

Synaptogenesis

Axonal and dendritic growth

Neuronal repair

REVIEWS

© 2011 Macmillan Publishers Limited. All rights reserved

Page 4: Cholesterol Metabolism in Huntington Disease

564 | OCTOBER 2011 | VOLUME 7 www.nature.com/nrneurol

small groups of patients with HD did not indicate any lipid abnormalities. Specifically, no significant changes were observed in plasma concentrations of lipids, includ­ing cholesterol.49 In addition, fibroblasts from patients with HD showed no evidence of alterations in cholesterol content and activity of 3­hydroxy­3­methylglutaryl­CoA reductase (HMG­CoAR), the rate­limiting step enzyme in the sterol synthesis pathway (Figure 2).50

By contrast, more­recent studies, which demonstrated altered cellular cholesterol synthesis and metabolism in HD patients and cell and animal models of HD, suggest that cholesterol homeostasis is impaired in HD. Blood

cholesterol levels are reduced in patients with late­stage HD,51 although whether this a direct effect of the HTT mutation or a result of late­stage metabolic changes remains to be established. Since direct measurements of brain cholesterol content in living individuals are not feasible, data are mostly generated from cell lines and postmortem tissue. Expression of mRNA for choles­terol biosynthetic genes, including those encoding HMG­CoAR and 7­DHC reductase, was reported to be reduced in a mutant­huntingtin­expressing cell line.52 A subsequent study reported reduced mRNA levels of HMG­CoAR, lanosterol 14 α­demethylase (CYP51) and 7­DHC reductase in fibroblasts and postmortem striatal and cortical tissue from patients with HD, and found that HD fibroblasts have a reduced ability to stimulate de novo cholesterol synthesis (Figure 2).53 Although sterol bio­synthesis is reduced in the brains of patients with HD, one group reported elevated cholesterol content in the caudate in this disorder.54

Two mechanisms could account for increased choles­terol content in the presence of reduced synthesis: first, a reduction in cholesterol conversion to 24­OHC, and second, an increase in brain cholesterol uptake from plasma to compensate for decreased cholesterol avail­ability from de novo synthesis. Whether sterol uptake from plasma is altered in patients with HD is not known. Brain cholesterol turnover to 24­OHC is estimated by measuring plasma 24­OHC levels, in view of the brain­specific distribution of CYP46, the enzyme responsible for 24­OHC generation.39,41 Plasma and cerebrospinal fluid (CSF) 24­OHC levels are reported to be altered in patients with AD or multiple sclerosis, raising the possibility that circulating 24­OHC levels may act as a biomarker for disease severity and progression in these conditions.55–58 CYP46 is expressed predominantly in neurons, and increased plasma 24­OHC in AD and multiple sclerosis signifies induction of neuronal cholesterol catabolism, possibly owing to increased cholesterol excretion by degenerating neurons. By contrast, plasma 24­OHC levels are reduced in patients with HD who exhibit motor symp­toms and normal in premanifest patients.51,59 One expla­nation for reduced 24­OHC in HD is a decrease in total neuronal cholesterol turnover as a result of reduced neu­ronal volume. However, the reduction in 24­OHC does not change across different HD stages, despite a progres­sive decrease in caudate volume.59 Hence, the more likely explanation is that attenuated production of 24­OHC in HD­affected brains is a result of inhibition of cholesterol excretion to compensate for reduced synthesis, in order to maintain normal brain cholesterol homeostasis.

Overall, human patient and tissue data demonstrate changes in body cholesterol metabolism and attenua­tion of brain cholesterol biosynthesis in HD, as well as a reduction in cholesterol excretion in the late stages of the disease.

Evidence from HD modelsUndoubtedly, the largest volume of data supporting a role for altered cholesterol metabolism in HD has been generated using cellular and animal models. However,

Acetyl-CoA+Acetoacetyl-CoA

HMG-CoA

Mevalonate

Lanosterol

HMG-CoA synthase

HMG-CoA reductase

Mevalonate-PP

Isopentenyl-PP

Geranyl-PP

Farnesyl-PP

Squalene

Zymosterol

Mestenol

24,25-Dihydrolanosterol

Lathosterol

Cholesterol

7-Dehydrocholesterol

24-Hydroxycholesterol

Cholesterol 24-hydroxylase

7-Dehydrocholesterol reductase

Desmosterol

7-Dehydrodesmosterol

Lanosterol 14-αdemethylase

Figure 2 | Influence of mutant huntingtin on the cholesterol biosynthesis pathway. Endogenous cholesterol is synthesized from acetyl-CoA. The rate-limiting step involves the synthesis of mevalonate by HMG-CoA reductase. After the synthesis of the first sterol precursor lanosterol, the pathway diverges and cholesterol can be synthesized from the precursors desmosterol or lathosterol. In the brain, cholesterol is mainly metabolized by cholesterol 24-hydroxylase to 24-hydroxycholesterol. Mutant huntingtin leads to a reduction in the expression of some enzymes and in the levels of some sterols, all of which are indicated in blue. By contrast, both a reduction and an accumulation in cholesterol have been reported in cellular and mouse models of Huntington disease. Abbreviations: HMG-CoA, 3-hydroxy-3-methylglutaryl-CoA; PP, pyrophosphate.

REVIEWS

© 2011 Macmillan Publishers Limited. All rights reserved

Page 5: Cholesterol Metabolism in Huntington Disease

NATURE REVIEWS | NEUROLOGY VOLUME 7 | OCTOBER 2011 | 565

the interpretations offered by numerous studies create a complicated picture, with different groups reporting seemingly conflicting results.

One line of evidence implicates reduced sterol bio­synthesis leading to reduced brain cholesterol content in HD animal models. The R6/2 transgenic mouse, which expresses exon 1 of the human HD gene includ­ing the polyglutamine­containing domain, and the yeast artificial chromosome 128 (YAC128) transgenic mouse, which expresses the full mutant Htt gene with 128 CAG repeats, both exhibit decreased brain cholesterol syn­thesis.53,60–62 These two models differ substantially in disease onset and lifespan, with R6/2 mice developing progressive motor deficits from about 9 weeks of age and dying before 16 weeks of age,63 and YAC128 mice showing motor abnormalities starting at 2 months, neuro pathology at 9 months and, in male mice, a reduced life span of 12 months.64,65 Hence, decreased cholesterol syn thesis can be attributed to the presence of mutant huntingtin rather than a secondary effect of long­term neuronal or metabolic changes.

HMG­CoAR, 7­DHC reductase and CYP51 expression levels, and HMG­CoAR activity were found to be dimin­ished in the brains of symptomatic R6/2 mice, leading to reduced levels of cholesterol precursors including lanosterol and lathosterol (Table 2).53,61 However, striatal cholesterol content analysis yielded inconsistent results, with no change in levels until the late symptomatic stage in this mouse strain.53,61 The turnover of total cholesterol content in the brain takes several months in the mouse,7,8 and the absence of robust cholesterol changes in R6/2 mice is probably explained by their short lifespan. In the longer­living YAC128 mice, brain HMG­CoAR activity, cholesterol precursors, cholesterol and 24­OHC levels are reduced during the symptomatic stage,60,62 indicat­ing that cholesterol synthesis and turnover are attenuated at a time when striatal atrophy is observed. Lathosterol, choles terol and 24­OHC levels are also decreased in symptomatic HD knock­in mice carrying the CAG

expansion within the endogenous Htt gene and a rat HD model carrying a truncated Htt fragment with 51 CAG repeats.62 However, cholesterol accumulation has been reported in YAC72 mice, which express the full mutant Htt gene with 72 CAG repeats and have milder motor behavioral and neuropathological abnormalities than YAC128 mice,65 and in HD knock­in mice.54,66

The detailed analysis of several animal models mirrors data from human samples demonstrating reduced sterol biosynthesis in the presence of mutant huntingtin. How­ever, another line of evidence implies a functional role for neuronal cholesterol accumulation in HD. When mutant­huntingtin­expressing neurons are isolated from the brain and grown in culture, they accumulate free cholesterol at the plasma membrane, and this process has been shown to contribute to excitotoxicity.54,66,67 Although the idea that reduced cholesterol synthesis and increased cellular cholesterol levels occur concomitantly and contribute to HD pathogenesis might seem counter­intuitive, several mechanisms could explain these obser­vations. As most cholesterol synthesis in the adult brain is thought to occur in glia, reduced sterol synthesis in HD brains may reflect cholesterol status in this particu­lar cell type. Indeed, levels of HMG­CoAR and 7­DHC reductase are attenuated in HD astrocytes.62 In addition, reduced brain sterol synthesis does not preclude accu­mulation of cholesterol in single cells, especially if cel­lular membrane properties are altered (as discussed in the next section), or if cholesterol trafficking is impaired. For example, ABCA1 and Apo­E, which are critical in cellular cholesterol efflux in the brain, are reduced in mutant­huntingtin­expressing astrocytes.62 Failure to efficiently remove cholesterol from the plasma mem­brane owing to reduced efflux could lead to neuronal cholesterol accumulation.

Several techniques can be used to assess sterol content in tissues and cells. Differences in methodologies affect the sensitivity of the measurements, and could underlie some of the variations in cholesterol levels reported by

Table 2 | Changes in plasma and brain cholesterol metabolism in HD

Study population or model Plasma Brain

Patients with HD Reduced levels of cholesterol, precursors lanosterol and lathosterol, and cholesterol metabolite 24-OHC51,59

Postmortem brain tissue:Reduced expression of enzymes involved in cholesterol synthesis, including HMG-CoAR, CYP51 and 7-DHC reductase53

Increased caudate cholesterol content54

R6/2 mouse model NA Reduced expression of HMG-CoAR, CYP51 and 7-DHC reductase53

Reduced HMG-CoAR activity61

Reduced levels of lanosterol and lathosterol61

Reduced cholesterol in late neuropathological stages53

YAC128 mouse model Reduced 24-OHC, reduced total cholesterol and precursors60

Reduced HMG-CoAR activity60

Reduced levels of lanosterol, lathosterol, cholesterol and 24-OHC60,62

YAC72 mouse model NA Cholesterol accumulation in neuronsIncreased striatal cholesterol content66

Hdh knock-in mouse model NA Reduced lathosterol, cholesterol and 24-OHC62

Increased cholesterol54

Rat model NA Reduced lathosterol, cholesterol and 24-OHC62

Abbreviations: 7-DHC, 7-dehydrocholesterol; 24-OHC, 24-hydroxycholesterol; CYP51, lanosterol 14-α demethylase; HD, Huntington disease; Hdh, murine HD homolog; HMG-CoAR, 3-hydroxy-3-methylglutaryl-CoA reductase; NA, not available.

REVIEWS

© 2011 Macmillan Publishers Limited. All rights reserved

Page 6: Cholesterol Metabolism in Huntington Disease

566 | OCTOBER 2011 | VOLUME 7 www.nature.com/nrneurol

separate groups. Striatal cholesterol was decreased in YAC72 mice as analyzed by gas chromatography–mass spectrometry,62 but increased cholesterol levels in the striatum were reported in YAC72 mice when measured by filipin staining and thin­layer chromatography (TLC),66 as well as in HD knock­in mice when measured by an enzymatic assay.54 In general, filipin staining is difficult to quantify and, owing to rapid photobleaching, is the least reliable method to measure changes in cholesterol content.68 Gas chromatography, TLC and enzymatic assays can also differ in their measurements of free and esteri­fied cholesterol and cholesterol oxidation products.69–71 To gain a better understanding of specific changes in cho­lesterol metabolism in HD, therefore, it is important not only to assess brain and neuronal cholesterol content in HD models, but also to study other markers of cholesterol turnover such as sterol pre cursors and 24­OHC, as well as intracellular cholesterol distribution.

Taken together, human and animal data demon­strate reduced cholesterol biosynthesis and turnover in HD­affected brains, and cholesterol accumulation in mutant­huntingtin­expressing neurons. Most of the changes in cholesterol synthesis are described in sympto­matic animals, suggesting that they occur as a result of mutant­huntingtin­induced cellular damage rather than acting as a contributing factor to neuronal death. How­ever, support for involvement of huntingtin in choles terol synthesis is provided by the finding that in contrast to YAC128 mice, YAC18 mice, which over express wild­type human HTT with 18 CAG repeats, show increased HMG­CoAR expression and activity.60,62 These obser­vations suggest that the cholesterol synthesis pathway is facilitated by wild­type huntingtin but impaired by mutant huntingtin.

Mechanisms of cholesterol change in HDHuntingtin is a ubiquitously expressed, 348 kDa intra­cellular protein, the highest levels of which are detected in neurons and the testes. Huntingtin is found in several intracellular compartments, including the nucleus, endo­plasmic reticulum, Golgi complex and mitochondria, and it also associates with clathrin­coated vesicles, cave­olae and microtubules.72,73 This extensive intra cellular distribution suggests involvement in many cellular pro­cesses. Huntingtin also has many interacting partners,73 which further broadens its potential influence on cell function. Huntingtin has been implicated in embryo­nic development, neurogenesis and neuronal survival, and at the molecular level it influences gene transcrip­tion, neuronal vesicle transport and endocytosis, and post synaptic activity.72,74

Huntingtin may influence the expression of choles­tero genic genes through its interaction with specific transcription factors. For example, huntingtin inter­acts with Sp1,74 a transcription factor also known to co operate with sterol regulatory element­binding protein (SREBP),75 which activates gene transcription when cellul ar ch olesterol is depleted.

Reduced cholesterol synthesisLow levels of cellular cholesterol lead to induction of sterol synthesis by activating SREBPs, which control the express ion of genes encoding cholesterogenic enzymes, includ ing HMG­CoAR.76 The SREBP family has three mem bers, SREBP­1a, SREBP­1c and SREBP2, which form a complex with the SREBP cleavage­activating pro­tein (SCAP) in the endoplasmic reticulum mem brane (Figure 3). Under low­cholesterol conditions, the SREBP–SCAP complex translocates to the Golgi apparatus

SREBPSCAP

Cytoplasm

Golgi

High cholesterol

ER lumen

SRE

Nucleus

SREBP

SCAP

Cytoplasm

Golgi

Low cholesterol

mHtt

INSIG

ER lumen

SRE

NucleusSREBPactive

INSIG

Increasedcholesterolsynthesis

Figure 3 | Influence of mHtt on SREBP-dependent cellular cholesterol synthesis. When cellular cholesterol is high, the SREBP–SCAP complex is retained in the ER membrane by the INSIG proteins, leading to inhibition of sterol synthesis. Under low-cholesterol conditions, INSIG dissociates from the SREBP–SCAP complex, which is then transported to the Golgi membrane. The active form of SREBP is cleaved and translocates into the nucleus, where it activates the expression of genes containing the SRE in their promoter region, including genes involved in cholesterol synthesis. In the presence of mHtt, inhibition of SREBP activation is observed when cellular cholesterol levels are low, resulting in reduced cholesterol synthesis. The details of the mechanism underlying reduced SREBP activation are unknown. Abbreviations: ER, endoplasmic reticulum; INSIG, insulin-induced gene; mHtt, mutant huntingtin; SCAP, SREBP cleavage-activating protein; SRE, sterol regulatory element; SREBP, sterol regulatory element-binding protein.

REVIEWS

© 2011 Macmillan Publishers Limited. All rights reserved

Page 7: Cholesterol Metabolism in Huntington Disease

NATURE REVIEWS | NEUROLOGY VOLUME 7 | OCTOBER 2011 | 567

membrane, where SREBP is activated by proteo lysis, releasing the active N­terminal domain into the cyto­sol.76,77 This protein fragment enters the nucleus and induces transcription of genes containing the sterol regu­latory element (SRE) in their promoter region. Levels of the active SREBP fragment are reduced in the nucleus and SRE induction is impaired in mutant­huntingtin­ expressing cells cultured in delipidated serum, suggesting that mutant huntingtin interferes with SREBP activation.53 Reduced SREBP activation leads to decreased expression of HMG­CoAR and, consequently, attenuated cholesterol synthesis, as observed in HD. INSIG­1 and INSIG­2 are SCAP­interacting proteins that retain the SREBP–SCAP complex in the endoplasmic reticulum in the presence of high cellular sterol concentrations. Excess cellular cholesterol inhibits HMG­CoAR activity in two ways: by promoting the interaction of INSIG­1 and INSIG­2 with SCAP, thereby retaining the SREBP–SCAP complex in the endoplasmic reticulum and inhibiting HMG­CoAR transcription, and through INSIG­1­mediated degrada­tion of HMG­CoAR protein.77,78 Hence, cholesterol accu­mulation in mutant­huntingtin­expressing cells may lead to reduced sterol synthesis through both impairment of SREBP activation and INSIG­mediated HMG­CoAR inhibition (Figure 4).

Cholesterol synthesis is also influenced by brain­derived neurotrophic factor (BDNF), which is secreted by cortical neurons projecting to the striatum and has an important role in neuronal survival (Figure 5). BDNF

stimulates neuronal de novo cholesterol synthesis.79 Huntingtin facilitates vesicular transport of BDNF along microtubules,5 and BDNF levels are reduced in post­mortem striatal tissue from patients with HD and brains of HD mice.80 Hence, attenuated BDNF levels may con­tribute to decreased striatal cholesterol synthesis—a model that is consistent with the preferential reduction in sterol levels in the striatum of HD mice.60,62

Cellular cholesterol accumulationNeurons expressing mutant huntingtin accumulate choles terol despite downregulation of cholesterol syn­thesis, indicating that mechanisms other than increased choles terol production are in operation. As already discussed, reduced efflux and decreased conversion to 24­OHC may partially contribute to cellular cholesterol accumulation. Data from in vitro experiments demon­strate that specific changes at the plasma membrane also lead to elevated cholesterol content (Figure 4). Lipid rafts are cholesterol­rich membrane domains that cluster the distribution of proteins, including neuro transmitter receptors, thereby forming signaling platforms that are essential for intracellular signal transduction.81 Caveolae, which are also rich in cholesterol, are membrane invagi­nations involved in cell signal ing and endocytosis.82 Mutant­huntingtin­expressing neurons show enhanced levels of the lipid raft marker ganglioside GM1, suggest­ing that membrane cholesterol accumulation is associated with an increased prevalence of lipid rafts.54 Caveolin­1

Indirect effect of mHttDirect effect of mHtt

mHtt

Endosome

Caveola

Increased membranecaveolin-1Endocytosis

Increased lipid rafts

ReducedHMG-CoARexpression

Cholesterolaccumulation

SREBPSCAP

Endoplasmic reticulum

INSIG

Nucleus

Caveolin-1

Cholesterol

Lipid rafts

Figure 4 | Neuronal cholesterol accumulation in the presence of mHtt. Caveolin-1 is expressed in caveolae, cholesterol-rich plasma membrane invaginations involved in endocytosis. Caveolin-1 binds cholesterol, and endocytosis of caveolin-1 is inhibited by mHtt, possibly leading to membrane cholesterol accumulation. In addition, mHtt increases the presence of lipid rafts—membrane signaling platforms enriched in cholesterol. Membrane cholesterol accumulation in mHtt-expressing neurons may lead to increased cholesterol in the endoplasmic reticulum, resulting in retention of the SREBP–SCAP complex and inhibition of cholesterol synthesis. The details of the mechanism, however, remain to be established. Abbreviations: HMG-CoAR, 3-hydroxy-3-methylglutaryl-CoA reductase; INSIG, insulin-induced gene; mHtt, mutant huntingtin; SCAP, SREBP cleavage-activating protein; SREBP, sterol regulatory element-binding protein.

REVIEWS

© 2011 Macmillan Publishers Limited. All rights reserved

Page 8: Cholesterol Metabolism in Huntington Disease

568 | OCTOBER 2011 | VOLUME 7 www.nature.com/nrneurol

is a protein found in caveolae that binds cholesterol.83,84 Mutant huntingtin impairs the internalization of caveo­lin­1, increasing its distribution at the plasma membrane and thereby leading to accumulation of membrane choles­terol.66 However, whether concentration of caveolin­1

results in increased membrane cholesterol in vivo is ques­tionable, as increased cholesterol distribution to mem­brane compart ments in response to elevated membrane caveolin­1 should stimulate cholesterol efflux in the pres­ence of extra cellular acceptors.85 Efflux may relieve mem­brane choles terol excess, as indicated by reduced sterol content in synaptosomes—plasma membrane fragments containing presynaptic and postsynaptic components —isolated from symptomatic R6/2 mouse brains.62 The presence of smaller, less­lipidated lipoprotein particles in the CSF of HD mice,62 however, suggests that cho les­terol efflux to Apo­E is reduced in HD brains, which may contribute to membrane cholesterol accumulation.

The evidence to date demonstrates that reduced choles terol synthesis and cellular cholesterol accumu­lation are both occurring in HD, and are connected through pathways that link cellular cholesterol levels to the regulation of its synthesis, catabolism and transport. One of the key challenges is to determine which aspect of cho les terol dysregulation predominantly influences neuronal function in HD.

Cholesterol and neuronal function in HDAn important question is whether changes in cho les­terol metabolism contribute to the pathogenesis in HD by influencing neuronal survival or neuroinflamma tion, or whether they exacerbate the consequences of known pathological mechanisms, including glutamate recep­tor­mediated excitotoxicity and disruption of axonal transport (Figure 6).

Neuronal function and survivalChanges in cellular and plasma membrane cholesterol content can have a marked impact on neuronal activ­ity and survival under basal and excitotoxic conditions. Reduced cholesterol synthesis in HD could impair synapse maturation, neurotransmitter vesicle generation and synaptic activity, all of which strongly depend on the presence of cholesterol.27,30 ABCA1 and Apo­E expression is decreased in YAC128 astrocytes, possibly to preserve intracellular cholesterol stores.62 However, the decrease in expression of these two proteins, which are critical for cholesterol transport, may compromise the astrocyte­to­neuron cholesterol shuttle, thereby limiting the amount of cholesterol delivered to neurons. Apo­E­containing lipoprotein particles are abnormally small in size in the CSF of YAC128 mice, indicating reduced cholesterol content.62 In addition to influencing basal neuronal activ­ity, a decrease in Apo­E and ABCA1 levels could impair cholesterol redistribution to injured neurons, which may render these cells more vulnerable to death.

The cholesterol content of synaptic vesicles is higher than that of other intracellular organelles.27 A reduc­tion in cholesterol availability for optimal synapto­genesis could result in decreases in synaptic vesicle numbers and synaptic activity. Presynaptic activity, as measured by paired­pulse facilitation, is reduced in HD mice, indicating reduced neurotransmitter release.86,87 Dopa mine release is attenuated in brain slices from sympto matic R6/2 mice owing to a depletion of reserve

P

PostsynapticPresynaptic

mHtt TrkB receptor

BDNF

Huntingtin

Microtubule

Motorcomplex

BDNF vesicle

P

Indirect effect of mHttDirect effect of mHtt

Cholesterol synthesis

Increased HMG-CoARexpression

Neuronal survival

Synaptic plasticity

Figure 5 | Huntingtin, BDNF trafficking and neuronal cholesterol synthesis. BDNF is released by cortical neurons projecting into the striatum. In addition to its recognized role in neuronal survival and synaptic plasticity, BDNF stimulates neuronal cholesterol synthesis via activation of Trk-B. Huntingtin promotes vesicular transport of BDNF vesicles along microtubules. This process is inhibited by mHtt, resulting in decreased BDNF levels in the striatum, which may be one of the pathways that contributes to reduced neuronal cholesterol synthesis in Huntington disease. Abbreviations: BDNF, brain-derived neurotrophic factor; HMG-CoAR, 3-hydroxy-3-methylglutaryl-CoA reductase; mHtt, mutant huntingtin; P, phosphate; Trk-B, BDNF/NT-3 growth factors receptor.

mHtt-inducedcellularcholesterolmetabolismchanges

In�uence oncellularfunction

In�uence onneuropathology

Glia

■ Reduced cholesterol synthesis■ Reduced cholesterol ef�ux

Neurons

■ Reduced cholesterol synthesis■ Reduced cholesterol ef�ux■ Reduced BDNF release■ Membrane cholesterol accumulation

■ Impaired cholesterol supply to neurons■ Impaired cholesterol- dependent response during neuronal injury■ Altered in�ammatory response

Neuroin�ammation

ExcitotoxicityReduced neuronalsurvival

■ Impaired generation of synaptic vesicles■ Increased NMDA receptors in lipid rafts■ Impaired caveolin-dependent endocytosis■ Impaired BDNF-dependent cholesterol synthesis

Figure 6 | Potential cholesterol-dependent pathways influencing neuropathology in Huntington disease. Several functions in glia and neurons are known to be dependent on cholesterol homeostasis. Changes in cellular cholesterol in mHtt-expressing glia and neurons can lead to discrete impairments in cell-type-specific functions (yellow-shaded boxes). Consequently, these impairments may influence the neuropathological features of Huntington disease, including neuronal survival, susceptibility to excitotoxicity, and neuroinflammation. Abbreviations: BDNF, brain-derived neurotrophic factor; mHtt, mutant huntingtin; NMDA, N-methyl-d-aspartate.

REVIEWS

© 2011 Macmillan Publishers Limited. All rights reserved

Page 9: Cholesterol Metabolism in Huntington Disease

NATURE REVIEWS | NEUROLOGY VOLUME 7 | OCTOBER 2011 | 569

vesicles available for mobilization.88,89 One should note that although vesicle generation and fusion depend on choles terol,27,84,90 vesicle content and release is also likely to be affected by the direct role of huntingtin in vesicular trafficking. Huntingtin interacts with proteins involved in endocytosis and vesicle trafficking, including HIP­1, HAP­1, kinesin heavy chain, and the p150Glued subunit of dynactin,5 and mutant huntingtin results in defective axonal vesicle transport.4

NeuroinflammationCellular cholesterol metabolism is controlled by liver X receptors (LXRs)—ligand­activated transcription factors that increase the expression of sterol transport genes in response to cellular cholesterol accumulation.91 LXR activation represses inflammation via inhibition of the nuclear factor­κB pathway.92,93 LXR agonists inhibit microglial activation, reduce neuro inflammation and improve outcomes in mouse models of AD, NPC disease, experimental autoimmune encephalo myelitis, and focal cerebral ischemia.94–99 APOE and ABCA1 are LXR target genes, and reduced APOE and ABCA1 mRNA in astro­cytes expressing mutant huntingtin, as well as reduced brain production of 24­OHC, an endogenous LXR agonist, suggest that LXR activation is attenuated in HD. The presence of mutant huntingtin has also been associ­ated with reduced LXR activation.100 Reduced LXR signal­ing leads to increased susceptibility to neuro inflammation and exacerbates neuropathology in AD mice,99 and may also play a part in neuronal death in HD, particularly since neuroinflammation is present in HD brains.101,102

ExcitotoxicityPlasma membrane cholesterol accumulation in mutant­huntingtin­expressing neurons is likely to disrupt sig­naling of neurotransmitter receptors distributed in lipid rafts. One proposed mechanism underlying neuronal death in HD is excitotoxicity mediated by NMDARs —ionotropic glutamate receptors that associate with lipid rafts.103 Increased distribution of NMDAR subunits to lipid rafts is observed in HD neurons,54 and post synaptic NMDAR activity is augmented in various HD mouse models.3,104–106 Mutant huntingtin also accumulates in lipid rafts with glycogen synthase kinase­3β (GSK­3β),107 which has been linked to tau phosphorylation, one of the neuropathological hallmarks of AD.108 GSK­3β activity promotes neuronal death, which can be attenuated by membrane cholesterol depletion.109–111 GSK­3β inhibitors reduce death of mutant­huntingtin­expressing neurons, suggesting that increased GSK­3β activity contributes to neuronal demise in HD.107 Taken together, these findings indicate that increased presence of membrane cho lesterol and lipid rafts may directly contribute to glutamate recep­tor­mediated excitotoxicity as well as GSK­3β activation in HD neurons.

Cholesterol as a therapeutic targetAnticonvulsants, neuroleptics and antidepressants are used to alleviate motor and behavioral symptoms in patients with HD,112 but no effective treatment that delays

onset or progression of the disease is currently available. After two decades of intense research into endogenous huntingtin function and mechanisms underlying the pathology of HD, however, specific targets for potential therapy are being identified. These targets include inhibi­tion of NMDAR and caspase­6 activity, as well as regu­lation of vesicular transport.6,112 NMDAR antagonism improves outcome in HD mice,113 and caspase­6 inhibitors for testing in mouse models are under development.

Animal studies using drugs aimed at altering brain cholesterol synthesis or reducing membrane cholesterol accumulation should establish whether targeting of choles terol metabolism represents a potential avenue for treating some of the neuronal impairments in HD. If the main culprit influencing the disease is reduced choles­terol synthesis, compounds that are capable of stimulat­ing the sterol synthesis pathway or increasing uptake of cellular cholesterol may prove beneficial to neurons. However, given that increased membrane cholesterol is found in HD neurons, stimulation of synthesis or uptake of the sterol would potentially result in further membrane cholesterol accumulation.

The use of cholesterol­reducing agents also has potential caveats. Statins, which inhibit the activity of HMG­CoAR, are widely used in the treatment of hyper­cholesterolemia, and have shown some neuro protective effects in neuro degenerative conditions, including AD, although the evidence from prospective studies is inconsistent.12,114 Simvastatin protects neurons from NMDA­induced excitotoxicity115 and reduces NMDA excitotoxicity in mutant­huntingtin­expressing cells by decreasing the prevalence of lipid rafts.54 Although statins have beneficial effects in neuronal culture experiments, however, a further decrease in HMG­CoAR activity resulting in attenuated sterol synthesis may exacerbate neuro pathology in HD. Another major caveat is that although drugs such as statins show effects in vitro, their direct effects on cholesterol synthesis in the CNS have not been proven. In fact, the neuroprotective function of statins is sometimes attributed to their pleiotropic effects other than inhibition of cholesterol synthesis.116

The effectiveness of any drug designed to target brain cholesterol will depend on its ability to cross the BBB and influence brain cholesterol metabolism. The different manifestations of cholesterol dysregulation in HD suggest that one approach to re­establishing brain cholesterol homeostasis is to restore efficient cholesterol transport out of cells as well as between cells in the CNS. Activation of LXRs, which include the LXRα and LXRβ isoforms, increases expression of genes encoding proteins involved in cholesterol transport, including ABCA1, ABCG1 and Apo­E.91 In cells that express mutant huntingtin, LXR activation should stimulate cholesterol efflux thereby normalizing membrane cholesterol content, facilitate cholesterol transport on Apo­E between astrocytes and neurons, and stimulate cholesterol synthesis if low sterol levels are reached. These potential effects of LXR activa­tion, as well as its recognized anti­inflammatory func­tion in neuro degenerative disease, indicate that the use of selective LXR agonists may be of interest in HD therapy.

REVIEWS

© 2011 Macmillan Publishers Limited. All rights reserved

Page 10: Cholesterol Metabolism in Huntington Disease

570 | OCTOBER 2011 | VOLUME 7 www.nature.com/nrneurol

However, the benefits of long­term LXR activation with currently available agents are hindered by their induc­tion of liver steatosis—a serious adverse effect—owing to increased lipogenesis via SREBP­1c.117 One way to over­come this problem is to develop agonists that selectively target the LXRβ isoform, which is widely expressed in the brain. Mice lacking LXRα, which is expressed pre­dominantly in the liver, do not exhibit LXR­agonist­induced hepatic steatosis.118 The therapeutic potential of selective LXRβ activation is recognized in the treat­ment of athero sclerosis as well as neurodegeneration, but LXRβ agonists are not yet available for testing.

Recently, a novel approach to modification of cholesterol metabolism has been reported. MicroRNAs (miRNAs) are endogenous, small, noncoding RNA molecules that bind to target mRNAs, leading to translational repression, and are abundantly expressed in the brain.119 miR­33 has recently been described as a regulator of cellular choles­terol metabolism because its activity reduces levels of the transporters ABCA1 and ABCG1.120,121 miR­33 inhibi­tors increase HDL formation—a finding that has raised substantial interest in miR­33 as a potential target in the treatment of atherosclerosis. Although most interest so far has been in the action of miR­33 in the liver, miR­33 is also highly expressed in the brain,120 suggesting that it could regulate cholesterol metabolism in the CNS. The widespread role of miRNAs in physiological processes makes them promising targets for drug develop ment, but whether miRNA targeting in the brain represents a valid therapeutic approach in neuro degeneration remains to be established.

ConclusionsData obtained over the past few years from humans and animal models indicate that cholesterol metabolism is altered in HD. Emerging evidence suggests that cholesterol changes in the presence of mutant huntingtin influence intercellular cholesterol transport, neuronal excitotoxicity and apoptosis. Changes in brain cholesterol homeostasis in HD underscore the importance of cholesterol in neu­ronal function and survival, and provide another example of the involvement of cholesterol dysregulation in neuro­degeneration. As the detailed role of cholesterol metabo­lism in HD emerges, it becomes clear that future studies should aim to resolve the question of whether reduced sterol biosynthesis or neuronal cholesterol accumulation is the predominant mechanism contributing to HD patho­genesis in vivo. Studies in whole animals, investigating the effects of cholesterol regulation on neuronal survival, are needed to provide support for research into the potential benefits of cholesterol­ targeting therapy in HD.

Review criteria

PubMed and Google Scholar databases were searched using the terms “Huntington disease cholesterol”, “huntingtin cholesterol”, “brain cholesterol” and “cholesterol neurodegeneration”. Publications from all years were searched and were limited to full-text papers written in English. Additionally, the bibliographies of selected papers were searched for further references on specific subjects for background information where necessary.

1. [No authors listed] A novel gene containing a trinucleotide repeat that is expanded and unstable on Huntington’s disease chromosomes. The Huntington’s Disease Collaborative Research Group. Cell 72, 971–983 (1993).

2. Zeron, M. M. et al. Increased sensitivity to N-methyl-d-aspartate receptor-mediated excitotoxicity in a mouse model of Huntington’s disease. Neuron 33, 849–860 (2002).

3. Levine, M. S. et al. Enhanced sensitivity to N-methyl-d-aspartate receptor activation in transgenic and knockin mouse models of Huntington’s disease. J. Neurosci. Res. 58, 515–532 (1999).

4. Trushina, E. et al. Mutant huntingtin impairs axonal trafficking in mammalian neurons in vivo and in vitro. Mol. Cell Biol. 24, 8195–8209 (2004).

5. Gauthier, L. R. et al. Huntingtin controls neurotrophic support and survival of neurons by enhancing BDNF vesicular transport along microtubules. Cell 118, 127–138 (2004).

6. Graham, R. K. et al. Cleavage at the caspase-6 site is required for neuronal dysfunction and degeneration due to mutant huntingtin. Cell 125, 1179–1191 (2006).

7. Bjorkhem, I. & Meaney, S. Brain cholesterol: long secret life behind a barrier. Arterioscler. Thromb. Vasc. Biol. 24, 806–815 (2004).

8. Dietschy, J. M. & Turley, S. D. Thematic review series: brain lipids. Cholesterol metabolism in the central nervous system during early development and in the mature animal. J. Lipid Res. 45, 1375–1397 (2004).

9. Porter, F. D. & Herman, G. E. Malformation syndromes caused by disorders of cholesterol synthesis. J. Lipid Res. 52, 6–34 (2011).

10. Vance, J. E. Lipid imbalance in the neurological disorder, Niemann–Pick C disease. FEBS Lett. 580, 5518–5524 (2006).

11. Puglielli, L., Tanzi, R. E. & Kovacs, D. M. Alzheimer’s disease: the cholesterol connection. Nat. Neurosci. 6, 345–351 (2003).

12. Wolozin, B. Cholesterol and the biology of Alzheimer’s disease. Neuron 41, 7–10 (2004).

13. Ikonen, E. Mechanisms for cellular cholesterol transport: defects and human disease. Physiol. Rev. 86, 1237–1261 (2006).

14. Clee, S. M. et al. Age and residual cholesterol efflux affect HDL cholesterol levels and coronary artery disease in ABCA1 heterozygotes. J. Clin. Invest. 106, 1263–1270 (2000).

15. Cohen, J. et al. Low LDL cholesterol in individuals of African descent resulting from frequent nonsense mutations in PCSK9. Nat. Genet. 37, 161–165 (2005).

16. Brunham, L. R. et al. β-cell ABCA1 influences insulin secretion, glucose homeostasis and response to thiazolidinedione treatment. Nat. Med. 13, 340–347 (2007).

17. Glass, C. K. & Witztum, J. L. Atherosclerosis: the road ahead. Cell 104, 503–516 (2001).

18. Zlokovic, B. V. The blood–brain barrier in health and chronic neurodegenerative disorders. Neuron 57, 178–201 (2008).

19. Wilson, J. D. The measurement of the exchangeable pools of cholesterol in the baboon. J. Clin. Invest. 49, 655–665 (1970).

20. Dehouck, B. et al. A new function for the LDL receptor: transcytosis of LDL across the blood–brain barrier. J. Cell Biol. 138, 877–889 (1997).

21. Goti, D. et al. Scavenger receptor class B, type I is expressed in porcine brain capillary endothelial cells and contributes to selective uptake of HDL-associated vitamin E. J. Neurochem. 76, 498–508 (2001).

22. Panzenboeck, U. et al. ABCA1 and scavenger receptor class B, type I, are modulators of reverse sterol transport at an in vitro blood–brain barrier constituted of porcine brain capillary endothelial cells. J. Biol. Chem. 277, 42781–42789 (2002).

23. Panzenboeck, U. et al. Regulatory effects of synthetic liver X receptor- and peroxisome-proliferator activated receptor agonists on sterol transport pathways in polarized cerebrovascular endothelial cells. Int. J. Biochem. Cell Biol. 38, 1314–1329 (2006).

24. Do, T. M. et al. Direct evidence of abca1-mediated efflux of cholesterol at the mouse blood–brain barrier. Mol. Cell Biochem. doi:10.1007/s11010-011-0910–6.

25. Karasinska, J. M. et al. Specific loss of brain ABCA1 increases brain cholesterol uptake and influences neuronal structure and function. J. Neurosci. 29, 3579–3589 (2009).

26. Saito, K. et al. Ablation of cholesterol biosynthesis in neural stem cells increases their VEGF expression and angiogenesis but causes neuron apoptosis. Proc. Natl Acad. Sci. USA 106, 8350–8355 (2009).

REVIEWS

© 2011 Macmillan Publishers Limited. All rights reserved

Page 11: Cholesterol Metabolism in Huntington Disease

NATURE REVIEWS | NEUROLOGY VOLUME 7 | OCTOBER 2011 | 571

27. Pfrieger, F. W. Role of cholesterol in synapse formation and function. Biochim. Biophys. Acta 1610, 271–280 (2003).

28. Fünfschilling, U., Saher, G., Xiao, L., Mobius, W. & Nave, K. A. Survival of adult neurons lacking cholesterol synthesis in vivo. BMC Neurosci. 8, 1 (2007).

29. Barres, B. A. & Smith, S. J. Neurobiology. Cholesterol—making or breaking the synapse. Science 294, 1296–1297 (2001).

30. Mauch, D. H. et al. CNS synaptogenesis promoted by glia-derived cholesterol. Science 294, 1354–1357 (2001).

31. LaDu, M. J. et al. Nascent astrocyte particles differ from lipoproteins in CSF. J. Neurochem. 70, 2070–2081 (1998).

32. Vance, J. E., Hayashi, H. & Karten, B. Cholesterol homeostasis in neurons and glial cells. Semin. Cell Dev. Biol. 16, 193–212 (2005).

33. Handelmann, G. E., Boyles, J. K., Weisgraber, K. H., Mahley, R. W. & Pitas, R. E. Effects of apolipoprotein E, β-very low density lipoproteins, and cholesterol on the extension of neurites by rabbit dorsal root ganglion neurons in vitro. J. Lipid Res. 33, 1677–1688 (1992).

34. de Chaves, E. I., Rusinol, A. E., Vance, D. E., Campenot, R. B. & Vance, J. E. Role of lipoproteins in the delivery of lipids to axons during axonal regeneration. J. Biol. Chem. 272, 30766–30773 (1997).

35. Ignatius, M. J., Shooter, E. M., Pitas, R. E. & Mahley, R. W. Lipoprotein uptake by neuronal growth cones in vitro. Science 236, 959–962 (1987).

36. Snipes, G. J., McGuire, C. B., Norden, J. J. & Freeman, J. A. Nerve injury stimulates the secretion of apolipoprotein E by nonneuronal cells. Proc. Natl Acad. Sci. USA 83, 1130–1134 (1986).

37. Cartagena, C. M. et al. Cortical injury increases cholesterol 24S hydroxylase (Cyp46) levels in the rat brain. J. Neurotrauma 25, 1087–1098 (2008).

38. Fukumoto, H., Deng, A., Irizarry, M. C., Fitzgerald, M. L. & Rebeck, G. W. Induction of the cholesterol transporter ABCA1 in central nervous system cells by liver X receptor agonists increases secreted Aβ levels. J. Biol. Chem. 277, 48508–48513 (2002).

39. Lund, E. G., Guileyardo, J. M. & Russell, D. W. cDNA cloning of cholesterol 24-hydroxylase, a mediator of cholesterol homeostasis in the brain. Proc. Natl Acad. Sci. USA 96, 7238–7243 (1999).

40. Bjorkhem, I., Lutjohann, D., Breuer, O., Sakinis, A. & Wennmalm, A. Importance of a novel oxidative mechanism for elimination of brain cholesterol. Turnover of cholesterol and 24(S)-hydroxycholesterol in rat brain as measured with 18O2 techniques in vivo and in vitro. J. Biol. Chem. 272, 30178–30184 (1997).

41. Lund, E. G. et al. Knockout of the cholesterol 24-hydroxylase gene in mice reveals a brain-specific mechanism of cholesterol turnover. J. Biol. Chem. 278, 22980–22988 (2003).

42. Corder, E. H. et al. Gene dose of apolipoprotein E type 4 allele and the risk of Alzheimer’s disease in late onset families. Science 261, 921–923 (1993).

43. Jiang, Q. et al. ApoE promotes the proteolytic degradation of Aβ. Neuron 58, 681–693 (2008).

44. Chu, L. W. et al. A novel intronic polymorphism of ABCA1 gene reveals risk for sporadic Alzheimer’s disease in Chinese. Am. J. Med. Genet. B Neuropsychiatr. Genet. 144B, 1007–1013 (2007).

45. Rodriguez-Rodriguez, E. et al. Association of genetic variants of ABCA1 with Alzheimer’s

disease risk. Am. J. Med. Genet. B Neuropsychiatr. Genet. 144B, 964–968 (2007).

46. Hollingworth, P. et al. Common variants at ABCA7, MS4A6A/MS4A4E, EPHA1, CD33 and CD2AP are associated with Alzheimer’s disease. Nat. Genet. 43, 429–435 (2011).

47. Naj, A. C. et al. Common variants at MS4A4/MS4A6E, CD2AP, CD33 and EPHA1 are associated with late-onset Alzheimer’s disease. Nat. Genet. 43, 436–444 (2011).

48. Reitz, C. et al. Association of higher levels of high-density lipoprotein cholesterol in elderly individuals and lower risk of late-onset Alzheimer disease. Arch. Neurol. 67, 1491–1497 (2010).

49. Hooghwinkel, G. J., Borri, P. F. & Bruyn, G. W. Biochemical studies in Huntington’s chorea. II. Composition of blood lipids. Acta Neurol. Scand. 42, 213–220 (1966).

50. Maltese, W. A. Cholesterol synthesis in cultured skin fibroblasts from patients with Huntington’s disease. Biochem. Med. 32, 144–150 (1984).

51. Leoni, V. et al. Whole body cholesterol metabolism is impaired in Huntington’s disease. Neurosci. Lett. 494, 245–249 (2011).

52. Sipione, S. et al. Early transcriptional profiles in huntingtin-inducible striatal cells by microarray analyses. Hum. Mol. Genet. 11, 1953–1965 (2002).

53. Valenza, M. et al. Dysfunction of the cholesterol biosynthetic pathway in Huntington’s disease. J. Neurosci. 25, 9932–9939 (2005).

54. del Toro, D. et al. Altered cholesterol homeostasis contributes to enhanced excitotoxicity in Huntington’s disease. J. Neurochem. 115, 153–167 (2010).

55. Leoni, V. et al. Changes in human plasma levels of the brain specific oxysterol 24S- hydroxycholesterol during progression of multiple sclerosis. Neurosci. Lett. 331, 163–166 (2002).

56. Lutjohann, D. et al. Plasma 24S- hydroxycholesterol (cerebrosterol) is increased in Alzheimer and vascular demented patients. J. Lipid Res. 41, 195–198 (2000).

57. Papassotiropoulos, A. et al. 24S- hydroxycholesterol in cerebrospinal fluid is elevated in early stages of dementia. J. Psychiatr. Res. 36, 27–32 (2002).

58. Schonknecht, P. et al. Cerebrospinal fluid 24S- hydroxycholesterol is increased in patients with Alzheimer’s disease compared to healthy controls. Neurosci. Lett. 324, 83–85 (2002).

59. Leoni, V. et al. Plasma 24S-hydroxycholesterol and caudate MRI in pre-manifest and early Huntington’s disease. Brain 131, 2851–2859 (2008).

60. Valenza, M. et al. Cholesterol biosynthesis pathway is disturbed in YAC128 mice and is modulated by huntingtin mutation. Hum. Mol. Genet. 16, 2187–2198 (2007).

61. Valenza, M. et al. Progressive dysfunction of the cholesterol biosynthesis pathway in the R6/2 mouse model of Huntington’s disease. Neurobiol. Dis. 28, 133–142 (2007).

62. Valenza, M. et al. Cholesterol defect is marked across multiple rodent models of Huntington’s disease and is manifest in astrocytes. J. Neurosci. 30, 10844–10850 (2010).

63. Mangiarini, L. et al. Exon 1 of the HD gene with an expanded CAG repeat is sufficient to cause a progressive neurological phenotype in transgenic mice. Cell 87, 493–506 (1996).

64. Van Raamsdonk, J. M. et al. Loss of wild-type huntingtin influences motor dysfunction and survival in the YAC128 mouse model of Huntington disease. Hum. Mol. Genet. 14, 1379–1392 (2005).

65. Van Raamsdonk, J. M., Warby, S. C. & Hayden, M. R. Selective degeneration in YAC mouse models of Huntington disease. Brain Res. Bull. 72, 124–131 (2007).

66. Trushina, E. et al. Mutant huntingtin inhibits clathrin-independent endocytosis and causes accumulation of cholesterol in vitro and in vivo. Hum. Mol. Genet. 15, 3578–3591 (2006).

67. Luthi-Carter, R. et al. SIRT2 inhibition achieves neuroprotection by decreasing sterol biosynthesis. Proc. Natl Acad. Sci. USA 107, 7927–7932 (2010).

68. Reid, P. C. et al. A novel cholesterol stain reveals early neuronal cholesterol accumulation in the Niemann–Pick type C1 mouse brain. J. Lipid Res. 45, 582–591 (2004).

69. Csallany, A. S., Kindom, S. E., Addis, P. B. & Lee, J. H. HPLC method for quantitation of cholesterol and four of its major oxidation products in muscle and liver tissues. Lipids 24, 645–651 (1989).

70. Carr, T. P., Andresen, C. J. & Rudel, L. L. Enzymatic determination of triglyceride, free cholesterol, and total cholesterol in tissue lipid extracts. Clin. Biochem. 26, 39–42 (1993).

71. Carlson, S. E. & Goldfarb, S. A sensitive enzymatic method for determination of free and esterified tissue cholesterol. Clin. Chim. Acta 79, 575–582 (1977).

72. Cattaneo, E., Zuccato, C. & Tartari, M. Normal huntingtin function: an alternative approach to Huntington’s disease. Nat. Rev. Neurosci. 6, 919–930 (2005).

73. Li, S. H. & Li, X. J. Huntingtin–protein interactions and the pathogenesis of Huntington’s disease. Trends Genet. 20, 146–154 (2004).

74. Harjes, P. & Wanker, E. E. The hunt for huntingtin function: interaction partners tell many different stories. Trends Biochem. Sci. 28, 425–433 (2003).

75. Sanchez, H. B., Yieh, L. & Osborne, T. F. Cooperation by sterol regulatory element-binding protein and Sp1 in sterol regulation of low density lipoprotein receptor gene. J. Biol. Chem. 270, 1161–1169 (1995).

76. Brown, M. S. & Goldstein, J. L. The SREBP pathway: regulation of cholesterol metabolism by proteolysis of a membrane-bound transcription factor. Cell 89, 331–340 (1997).

77. Rawson, R. B. The SREBP pathway—insights from Insigs and insects. Nat. Rev. Mol. Cell Biol. 4, 631–640 (2003).

78. Goldstein, J. L. & Brown, M. S. Regulation of the mevalonate pathway. Nature 343, 425–430 (1990).

79. Suzuki, S. et al. Brain-derived neurotrophic factor regulates cholesterol metabolism for synapse development. J. Neurosci. 27, 6417–6427 (2007).

80. Zuccato, C. & Cattaneo, E. Brain-derived neurotrophic factor in neurodegenerative diseases. Nat. Rev. Neurol. 5, 311–322 (2009).

81. Simons, K. & Toomre, D. Lipid rafts and signal transduction. Nat. Rev. Mol. Cell Biol. 1, 31–39 (2000).

82. Parton, R. G. & Richards, A. A. Lipid rafts and caveolae as portals for endocytosis: new insights and common mechanisms. Traffic 4, 724–738 (2003).

83. Murata, M. et al. VIP21/caveolin is a cholesterol-binding protein. Proc. Natl Acad. Sci. USA 92, 10339–10343 (1995).

84. Thiele, C., Hannah, M. J., Fahrenholz, F. & Huttner, W. B. Cholesterol binds to synaptophysin and is required for biogenesis of synaptic vesicles. Nat. Cell Biol. 2, 42–49 (2000).

REVIEWS

© 2011 Macmillan Publishers Limited. All rights reserved

Page 12: Cholesterol Metabolism in Huntington Disease

572 | OCTOBER 2011 | VOLUME 7 www.nature.com/nrneurol

85. Fu, Y. et al. Expression of caveolin-1 enhances cholesterol efflux in hepatic cells. J. Biol. Chem. 279, 14140–14146 (2004).

86. Usdin, M. T., Shelbourne, P. F., Myers, R. M. & Madison, D. V. Impaired synaptic plasticity in mice carrying the Huntington’s disease mutation. Hum. Mol. Genet. 8, 839–846 (1999).

87. Klapstein, G. J. et al. Electrophysiological and morphological changes in striatal spiny neurons in R6/2 Huntington’s disease transgenic mice. J. Neurophysiol. 86, 2667–2677 (2001).

88. Ortiz, A. N., Kurth, B. J., Osterhaus, G. L. & Johnson, M. A. Dysregulation of intracellular dopamine stores revealed in the R6/2 mouse striatum. J. Neurochem. 112, 755–761 (2010).

89. Johnson, M. A. et al. Catecholamine exocytosis is diminished in R6/2 Huntington’s disease model mice. J. Neurochem. 103, 2102–2110 (2007).

90. Chamberlain, L. H., Burgoyne, R. D. & Gould, G. W. SNARE proteins are highly enriched in lipid rafts in PC12 cells: implications for the spatial control of exocytosis. Proc. Natl Acad. Sci. USA 98, 5619–5624 (2001).

91. Wojcicka, G., Jamroz-Wisniewska, A., Horoszewicz, K. & Beltowski, J. Liver X receptors (LXRs). Part I: structure, function, regulation of activity, and role in lipid metabolism. Postepy Hig. Med. Dosw. (Online) 61, 736–759 (2007).

92. Joseph, S. B., Castrillo, A., Laffitte, B. A., Mangelsdorf, D. J. & Tontonoz, P. Reciprocal regulation of inflammation and lipid metabolism by liver X receptors. Nat. Med. 9, 213–219 (2003).

93. Zelcer, N. & Tontonoz, P. Liver X receptors as integrators of metabolic and inflammatory signaling. J. Clin. Invest. 116, 607–614 (2006).

94. Hindinger, C. et al. Liver X receptor activation decreases the severity of experimental autoimmune encephalomyelitis. J. Neurosci. Res. 84, 1225–1234 (2006).

95. Lefterov, I. et al. Expression profiling in APP23 mouse brain: inhibition of Aβ amyloidosis and inflammation in response to LXR agonist treatment. Mol. Neurodegener. 2, 20 (2007).

96. Morales, J. R. et al. Activation of liver X receptors promotes neuroprotection and reduces brain inflammation in experimental stroke. Circulation 118, 1450–1459 (2008).

97. Repa, J. J. et al. Liver X receptor activation enhances cholesterol loss from the brain, decreases neuroinflammation, and increases survival of the NPC1 mouse. J. Neurosci. 27, 14470–14480 (2007).

98. Sironi, L. et al. Treatment with LXR agonists after focal cerebral ischemia prevents brain damage. FEBS Lett. 582, 3396–3400 (2008).

99. Zelcer, N. et al. Attenuation of neuroinflammation and Alzheimer’s disease pathology by liver x receptors. Proc. Natl Acad. Sci. USA 104, 10601–10606 (2007).

100. Futter, M. et al. Wild-type but not mutant huntingtin modulates the transcriptional activity of liver X receptors. J. Med. Genet. 46, 438–446 (2009).

101. Sapp, E. et al. Early and progressive accumulation of reactive microglia in the Huntington disease brain. J. Neuropathol. Exp. Neurol. 60, 161–172 (2001).

102. Moller, T. Neuroinflammation in Huntington’s disease. J. Neural Transm. 117, 1001–1008 (2010).

103. Allen, J. A., Halverson-Tamboli, R. A. & Rasenick, M. M. Lipid raft microdomains and neurotransmitter signalling. Nat. Rev. Neurosci. 8, 128–140 (2007).

104. Li, L., Murphy, T. H., Hayden, M. R. & Raymond, L. A. Enhanced striatal NR2B-containing N-methyl-d-aspartate receptor-mediated synaptic currents in a mouse model of Huntington disease. J. Neurophysiol. 92, 2738–2746 (2004).

105. Milnerwood, A. J. & Raymond, L. A. Corticostriatal synaptic function in mouse models of Huntington’s disease: early effects of huntingtin repeat length and protein load. J. Physiol. 585, 817–831 (2007).

106. Zeron, M. M. et al. Potentiation of NMDA receptor-mediated excitotoxicity linked with intrinsic apoptotic pathway in YAC transgenic mouse model of Huntington’s disease. Mol. Cell Neurosci. 25, 469–479 (2004).

107. Valencia, A. et al. Mutant huntingtin and glycogen synthase kinase 3-β accumulate in neuronal lipid rafts of a presymptomatic knock-in mouse model of Huntington’s disease. J. Neurosci. Res. 88, 179–190 (2010).

108. Reynolds, C. H., Betts, J. C., Blackstock, W. P., Nebreda, A. R. & Anderton, B. H. Phosphorylation sites on tau identified by nanoelectrospray mass spectrometry: differences in vitro between the mitogen-activated protein kinases ERK2, c-Jun N-terminal kinase and P38, and glycogen synthase kinase-3β. J. Neurochem. 74, 1587–1595 (2000).

109. Sui, Z., Kovacs, A. D. & Maggirwar, S. B. Recruitment of active glycogen synthase kinase-3 into neuronal lipid rafts. Biochem. Biophys. Res. Commun. 345, 1643–1648 (2006).

110. Hetman, M., Cavanaugh, J. E., Kimelman, D. & Xia, Z. Role of glycogen synthase kinase-3β in neuronal apoptosis induced by trophic withdrawal. J. Neurosci. 20, 2567–2574 (2000).

111. Linseman, D. A. et al. Glycogen synthase kinase-3β phosphorylates Bax and promotes its mitochondrial localization during neuronal apoptosis. J. Neurosci. 24, 9993–10002 (2004).

112. Ross, C. A. & Tabrizi, S. J. Huntington’s disease: from molecular pathogenesis to clinical treatment. Lancet Neurol. 10, 83–98 (2011).

113. Okamoto, S. et al. Balance between synaptic versus extrasynaptic NMDA receptor activity influences inclusions and neurotoxicity of mutant huntingtin. Nat. Med. 15, 1407–1413 (2009).

114. McGuinness, B. et al. Statins for the treatment of dementia. Cochrane Database of Systematic Reviews, Issue 8. Art. No.: CD007514. doi:10.1002/14651858.CD007514.pub2 (2010).

115. Ponce, J. et al. Simvastatin reduces the association of NMDA receptors to lipid rafts: a cholesterol-mediated effect in neuroprotection. Stroke 39, 1269–1275 (2008).

116. Butterfield, D. A., Barone, E. & Mancuso, C. Cholesterol-independent neuroprotective and neurotoxic activities of statins: perspectives for statin use in Alzheimer disease and other age-related neurodegenerative disorders. Pharmacol. Res. 64, 180–186 (2011).

117. Oosterveer, M. H., Grefhorst, A., Groen, A. K. & Kuipers, F. The liver X receptor: control of cellular lipid homeostasis and beyond Implications for drug design. Prog. Lipid Res. 49, 343–352 (2010).

118. Lund, E. G. et al. Different roles of liver X receptor α and β in lipid metabolism: effects of an α-selective and a dual agonist in mice deficient in each subtype. Biochem. Pharmacol. 71, 453–463 (2006).

119. Madathil, S. K., Nelson, P. T., Saatman, K. E. & Wilfred, B. R. MicroRNAs in CNS injury: potential roles and therapeutic implications. Bioessays 33, 21–26 (2010).

120. Rayner, K. J. et al. MiR-33 contributes to the regulation of cholesterol homeostasis. Science 328, 1570–1573 (2010).

121. Najafi-Shoushtari, S. H. et al. MicroRNA-33 and the SREBP host genes cooperate to control cholesterol homeostasis. Science 328, 1566–1569 (2010).

AcknowledgmentsThis study was supported by grants from the Canadian Institutes of Health Research (CIHR MOP-106684 and MOP-84438).

Author contributionsJ. M. Karasinska researched data for the article and wrote the text. J. M. Karasinska and M. R. Hayden contributed equally to discussions of content, and review and editing of the manuscript before submission.

REVIEWS

© 2011 Macmillan Publishers Limited. All rights reserved