Volume 110, Issue 6 p. 1737-1765
Free Access

The pathogenic mechanisms of polyglutamine diseases and current therapeutic strategies

Peter O. Bauer

Peter O. Bauer

Laboratory for Structural Neuropathology, RIKEN Brain Science Institute, Saitama, Japan

Search for more papers by this author
Nobuyuki Nukina

Nobuyuki Nukina

Laboratory for Structural Neuropathology, RIKEN Brain Science Institute, Saitama, Japan

Search for more papers by this author
First published: 03 September 2009
Citations: 145
Address correspondence and reprint requests to Nobuyuki Nukina, Laboratory for Structural Neuropathology, RIKEN Brain Science Institute, 2-1 Hirosawa Wako-shi, Saitama 351-0198, Japan. E-mail: [email protected]

Abstract

Expansion of CAG trinucleotide repeat within the coding region of several genes results in the production of proteins with expanded polyglutamine (PolyQ) stretch. The expression of these pathogenic proteins leads to PolyQ diseases, such as Huntington’s disease or several types of spinocerebellar ataxias. This family of neurodegenerative disorders is characterized by constant progression of the symptoms and molecularly, by the accumulation of mutant proteins inside neurons causing their dysfunction and eventually death. So far, no effective therapy actually preventing the physical and/or mental decline has been developed. Experimental therapeutic strategies either target the levels or processing of mutant proteins in an attempt to prevent cellular deterioration, or they are aimed at the downstream pathologic effects to reverse or ameliorate the caused damages. Certain pathomechanistic aspects of PolyQ disorders are discussed here. Relevance of disease models and recent knowledge of therapeutic possibilities is reviewed and updated.

Abbreviations used

  • 17-AAG
  • 17-(allylamino)-17-demethoxygeldanamycin
  • AR
  • androgen receptor
  • BAC
  • bacterial artificial chromosome
  • BDNF
  • brain-derived neurotrophic factor
  • CBP
  • CREB-binding protein
  • CHIP
  • C-terminal Hsp70-interacting protein
  • CK
  • casein kinase
  • CRE
  • cAMP response element
  • CREB
  • cAMP response element-binding protein
  • DRPLA
  • dentatorubropallidoluysian atrophy
  • ER
  • endoplasmic reticulum
  • HD
  • Huntington’s disease
  • Hdj
  • human DNAJ
  • HHR23A
  • human homologs of yeast DNA repair protein Rad23
  • Hip-1
  • huntingtin-interacting protein 1
  • HSP
  • heat-shock protein
  • htt
  • huntingtin
  • IP3R1
  • inositol (1,4,5)-trisphosphate receptor type 1
  • mGluR5
  • metabotropic glutamate receptor 5
  • MPT
  • mitochondrial permeability transition
  • NF-Y
  • nuclear factor Y
  • PGC-1
  • PPARγ coactivator-1
  • polyQ
  • polyglutamine
  • ROS
  • reactive oxygen species
  • SAHA
  • suberoylanilide hydroxamic acid
  • SB
  • sodium butyrate
  • SBMA
  • spinal bulbar muscular atrophy
  • SCA
  • spinocerebellar ataxia
  • siRNA
  • small-interfering RNA
  • SP1
  • specificity protein 1
  • SUMO
  • small ubiquitin-like modifier
  • TAF
  • TBP-associated factor
  • TBP
  • TATA-binding protein
  • TFTC/STAGA
  • TBP-free TBP-associated factor-containing/suppressor of Ty 3 homolog (SPT3)/RNA polymerase II, TBP-associated factor (TAF9)/general control of amino-acid synthesis 5 (GCN5) acetyltransferase complex
  • TG
  • transglutaminase
  • TLS
  • translocated in liposarcoma
  • UPS
  • ubiquitin-proteasome system
  • YAC
  • yeast artificial chromosome
  • Polyglutamine (polyQ) diseases comprise one of the most common groups of inherited neurodegenerative conditions. This category of diseases is characterized by the pathological expansion of CAG trinucleotide repeat in the translated regions of unrelated genes. There are at least nine polyQ-related disorders known to date. The first disease connected with the expansion of CAG repeat and causing progressive motor neuron degeneration, spinal bulbar muscular atrophy (SBMA), was reported in 1991 (La Spada et al. 1991). Eight other disorders including Huntington’s disease (HD), dentatorubropallidoluysian atrophy (DRPLA), and six types of spinocerebellar ataxia (SCA1, 2, 3, 6, 7, and 17) (Gardian et al. 2005) have since been identified, as associated with expanded polyQ (Table 1). The most common polyQ diseases worldwide are HD and SCA3 (Schols et al. 2004), but the incidence of these disorders differs between nations. For example, SCA3 accounts for vast majority of SCA in the Portuguese/Brazilian population but appears rarely in the Italian or is absent in the Czech population (Schols et al. 2004; Bauer et al. 2005b).

    Table 1. List of polyQ diseases with respective genomic loci, protein names, and the ranges of normal and pathological expansion
    Disease Locus Protein PolyQ expansion
    Normal Pathological
    SBMA Xq11-q12 Androgen receptor 6–36 38–62
    DRPLA 12p13 Atrophin-1 3–38 49–88
    HD 4p16.3 Huntingtin 6–35 36–121
    SCA1 6p23 Ataxin-1 6–39 41–83
    SCA2 12q24 Ataxin-2 14–32 34–77
    SCA3 14q24-q31 Ataxin-3 12–40 62–86
    SCA6 19p13 CACNA1A 4–18 21–30
    SCA7 3p21–p12 Ataxin-7 7–18 38–200
    SCA17 6q27 TATA-binding protein 25–43 45–63
    • SBMA, spinal bulbar muscular atrophy; DRPLA, dentatorubropallidoluysian atrophy; HD, Huntington’s disease; spinocerebellar ataxia; polyQ, polyglutamine.

    The amounts of mRNA and protein produced seem to be relatively unaffected in these conditions although de novo allele-specific DNA methylation has been proposed to affect the expression of mutant ataxin-2 in SCA2 patients (Bauer et al. 2004b). The disease phenotypes are usually observed when the number of glutamines exceeds ∼35–45. However, in the case of SCA6, the pathological threshold is ∼20 repeats, and in SCA3 it is closer to 60 repeats (Shao and Diamond 2007). The expanded CAG repeat is unstable and tends to expand further leading to earlier age of onset and a more severe disease course in successive generations, a phenomenon called anticipation (Schols et al. 2004). Anticipation was also observed without an increase in CAG number indicating a complicated pathomechanism of polyQ diseases (Bauer et al. 2005a). In some diseases, such as SCA2 and SCA7, the CAG triplets can be extremely unstable. This instability results in either a very long repeat with juvenile age of onset or a de novo pathological expansion from the normal length allele in the absence of an intermediary length repeat (Mao et al. 2002; Bauer et al. 2004a).

    No cases of HD, SBMA, DRPLA, or SCA1–3 and 7 with deletions or point mutations in their genes were reported, suggesting that these disorders do not result from a loss of gene function but rather from a gain of toxic function. The situation in SCA6 is less clear. Point mutations in the P/Q calcium channel gene cause the channel disorders episodic ataxia type 2, and familial hemiplegic migraine. There is significant overlap between these disorders and SCA6 in their clinical presentations (Jen et al. 1998; Alonso et al. 2003). Recently, polyQ-containing intranuclear inclusions in Purkinje cells and brainstem neurons were found in the brains of SCA8 patients and in a bacterial artificial chromosome (BAC)-transgenic mouse model of SCA8 (Moseley et al. 2006). SCA8 is believed to be caused by the expansion of a CTG repeat sequence, which resides in an untranslated endogenous antisense RNA that overlaps the Kelch-like 1 gene (Chen et al. 2008a). The newly discovered gene, ataxin-8, spans the repeat antiparallely to Kelch-like 1 and encodes a polyQ expansion in the CAG direction (Moseley et al. 2006).

    Several shared features of polyQ diseases indicate a common toxic effect related to the polyQ expansion. They are all unremitting and progressive diseases. Both normal and abnormal proteins are usually expressed at the same level in the tissues. Also, there is no clear relationship between expression pattern and site of pathology, except for SCA6, where the gene product is expressed predominantly in Purkinje cells (Ishikawa et al. 1999). Although most of the proteins associated with polyQ diseases are expressed systemically, the cytotoxicity appears restricted to certain neuronal subtypes in the CNS. This suggests that probably certain specific cellular conditions exist in vulnerable neurons that may cause the selective cytotoxicity by their gene products.

    In first part of this review, we discuss several points of the polyQ-related pathogenesis. The second part addresses different treatment strategies for polyQ diseases (Fig. 1) while providing an update of recent knowledge in this field.

    Details are in the caption following the image

    Summary of therapeutic strategies in polyQ diseases. Molecules targeting different events of polyQ diseases pathogenesis (black boxes) are displayed in red boxes. Blue background highlights processing of the polyQ protein.

    Pathomechanism of polyglutamine diseases

    Misfolding and aggregation of mutant polyglutamine proteins

    It is hypothesized that polyQ diseases are the result of a toxic gain of function that occurs at the protein level. A prominent pathological feature in most of these diseases is the intranuclear and cytoplasmic accumulation of aggregated polyQ proteins inside neurons (Davies et al. 1997; DiFiglia et al. 1997). The role of the aggregation in disease pathogenesis is controversial. It is not completely clear whether the toxicity of the expanded polyQ proteins results from the presence of visible aggregates or from smaller intermediary species generated during the aggregation process. Aggregates may merely represent end products of the upstream toxic event. Some studies have suggested that the inclusions serve a protective role (Saudou et al. 1998). Several cellular models show the discrepancy between inclusion formation and cell death. For example, in rat primary striatal neurons, the inclusions were not a prerequisite for cell death but mutant huntingtin (htt) had to be present in the nucleus to induce apoptosis (Saudou et al. 1998). Apoptosis in neuroblastoma cell lines was increased in the presence of mutant htt however it did not correlate with aggregate formation (Lunkes and Mandel 1998; Lunkes et al. 1999). The first transgenic mouse modeling HD had a phenotype reminiscent of HD and extensive striatal intranuclear inclusions; however, there was minimal cell death (Mangiarini et al. 1996). Other mice had striatal loss but fewer or no inclusions (Reddy et al. 1998; Aronin et al. 1999; Hodgson et al. 1999). On the other hand, aggregates are usually found in affected areas of the patients’ brains rather than in unaffected areas and the late onset and progressive nature of polyQ diseases can only be best explained by the slow process of protein aggregation. In SBMA, the inclusions are clearly associated with neuronal death. Despite the widely distributed androgen receptor (AR) throughout the CNS (Li et al. 1998), nuclear aggregates were observed only in affected motor neurons in the spinal cord and brainstem (Apostolinas et al. 1999). This may not be true in all polyQ diseases because the aggregates are also found in the dentate nucleus of HD cerebellum, a region of the brain unaffected by neurodegeneration in HD (Becher et al. 1998). Another exemption is, based on the observation in human necropsied cases of SCA1, SCA2, SCA3, and DRPLA, the absence of intranuclear inclusions in cerebellar Purkinje cells, which are targets of neurodegeneration in these polyQ diseases (Koyano et al. 2002).

    In recent years, it has been suggested that oligomeric species such as protofibrils and microaggregates are the direct source of polyQ toxicity and that large aggregates are cytoprotective (Arrasate et al. 2004). It was proposed that above a certain length threshold, polyQ sequences form oligomers stabilized by hydrogen-bonded polar zippers and associate via formation of hydrogen bonds (Perutz et al. 1994). Non-covalent interactions with other proteins and effect of transglutaminase (TG) were also considered to play roles in polyQ proteins oligomerization (Kahlem et al. 1996; Green 1993). Expanded polyQ tracts are good substrates for TGs in vitro (Kahlem et al. 1998), and the presence of TG inhibitors prevented aggregate formation in COS-7 cells (these cells obtained by immortalizing a cell line derived from kidney cells of the African green monkey with a version of the SV40) over-expressing truncated forms of mutant atrophin-1 or htt (Saudou et al. 1998; Igarashi et al. 1998). Formation of a β-sheet structure may also cause generation of fibrillar and non-fibrillar aggregates (Tanaka et al. 2001; Perutz et al. 2002). In vitro, fibrillary appearances of inclusions under electron microscopy and double refraction after Congo red staining are reminiscent of amyloid and are consistent with the polar zipper hypothesis (Wanker 2000; Hollenbach et al. 1999). Tanaka et al. (2003) reported another form of aggregate named quasi-aggregate, which exposed polyQ tracts on the surface of non-fibrillar aggregates. Recently, it was reported a toxicity of even a monomeric β-sheet conformer of polyQ proteins (Nagai et al. 2007). Analysis of the effect of polyQ protein conformation on cytotoxicity revealed that mutant polyQ β-strand/β-turn structure results in aggregation and toxicity in neuronal cells while polyQ protein analogs unable to assume β-strand/β-turn structure do not aggregate and are not toxic (Poirier et al. 2005).

    Modulation of polyQ aggregation by sequences flanking polyQ stretch has been reported in several studies showing this phenomenon in ataxin-1, ataxin-3, and htt (Ellisdon et al. 2006; Masino et al. 2004; Bhattacharyya et al. 2006; de Chiara et al. 2005; Thakur et al. 2009). In case of mutant htt, flanking sequences on both C- and N-terminal sides of the polyQ in exon 1 influence the aggregation. Presence of the C-terminal proline-rich domain reduces the htt exon 1 aggregation kinetics without changing the aggregation mechanism by favoring more aggregation-resistant conformations (Bhattacharyya et al. 2006). The N-terminal domain consisting of first 17 amino acids of the htt exon 1, on the other hand, determines the aggregation mechanism. This domain unfolds in a polyQ length-dependent manner and self-aggregates to form oligomers with cores composed of the N-terminal domains. PolyQ sequences stretch out during this process and during potential further nucleation. Monomers of expanded htt exon 1 can be then added to these prefibrillar aggregates forming amyloid-like structures (Thakur et al. 2009). The ability of the N-terminal sequence of htt exon 1 to remodel the polyQ aggregation mechanism indicates that polyQ proteins may, depending on flanking sequences, aggregate by distinct mechanisms resulting in variable aggregate morphologies and having different cellular effects.

    Recently, Furukawa et al. reported cross-seeding mechanism between polyQ- and RNA-binding protein T cell intracellular antigen-1 (TIA-1), in which the seed of polyQ aggregates induces T cell intracellular antigen-1 (TIA-1) aggregates. The result suggested that the different protein aggregates are induced through cross-seeding from the primary polyQ aggregates. This mechanism may also explain the diversity of pathology of polyQ diseases (Furukawa et al. 2009).

    Post-translational modifications of mutant polyglutamine proteins

    Proteolytic cleavage of polyglutamine proteins

    Pathogenesis of several polyQ disorders including HD, SBMA, and SCA3, appears linked to proteolytic cleavage resulting in production of toxic polyQ-containing fragments. In other diseases, such as SCA1, the proteolysis of the mutant protein has been not confirmed. The conformational changes in the proteins with expanded polyQ lead to misfolding and proteolytic cleavage into smaller toxic fragments (Paulson 1999, 2000; Lunkes et al. 2002; Gardian et al. 2005). The use of several antibodies directed against N- and C-terminal parts of the protein showed that only a truncated version, including the polyQ expansion, aggregates in the vast majority of the polyQ diseases (Wellington and Hayden 2000). Caspase-mediated cleavage sites were identified or predicted in htt, atrophin-1, ataxin-3, and AR (Wellington et al. 1998). As a consequence of the cleavage, the resulting fragment could have an enhanced toxic effect (Ikeda et al. 1996; Ellerby et al. 1999) and/or could more easily enter the nucleus (Igarashi et al. 1998; Hackam et al. 1999) and/or more rapidly aggregate (Igarashi et al. 1998; Cooper et al. 1998). In several studies, truncated proteins with polyQ expansions appeared more prone than full-length proteins to form inclusions or cause cell death by apoptosis (Paulson et al. 1997; Ikeda et al. 1996; Martindale et al. 1998; Merry et al. 1998). Inclusions found in the brains of HD patients were stained by antibodies against sequences located close to polyQ stretch but not by those against more C-terminal parts of htt (Sieradzan et al. 1999; DiFiglia et al. 1997). The N-terminal fragments in the patients’ brains have been identified as generated by caspases (Wellington et al. 2002). On the other hand, the presence of full-length mutant htt in the sodium dodecyl sulfate and urea insoluble material from HD patients’ brains was reported (Dyer and McMurray 2001). A more recent study has shown that the inclusions isolated from HD brains contain broad ranges of N-terminal fragments of expanded htt with sizes of 50–150 kDa rather than intact mutant protein (Hoffner et al. 2005). These fragments may result from multiple proteolytic activities with little specificity. It is however possible, that the smaller products of specific cleavage seed the inclusions followed by recruitment of larger non-specific fragments. This is supported by the fact that the cleavage sites for caspases 2, 3, and 6 have been identified in human htt at amino acids 552, 513 (and 552), and 586, respectively, while the cleavage at amino acid 552 is an early event in the course of the disease (Wellington et al. 2002). The cleavage of mutant htt at the 586 amino acid by caspase 6 has been shown to be crucial in the HD pathogenesis as it was essential for the HD-related behavioral phenotype and selective neuropathology in yeast artificial chromosome HD mouse model expressing the whole human htt with 128Q (YAC128) (Graham et al. 2006). Modulating the mutant htt cleavage by caspase 3 at the amino acids 513 and 552, on the other hand, had no effect on disease progression and neurodegeneration in these mice. Caspase 6- but not caspase 3-resistant mutant htt protected neurons from excitotoxic stress suggesting that not all htt fragments contribute to excitotoxic neuronal death equally. Moreover, nuclear translocation of expanded htt was delayed in mice expressing mutant htt resistant to caspase 6 cleavage (Graham et al. 2006). This study strongly supports the hypothesis that generation of a specific htt fragment may represent an initial event in HD pathogenesis.

    It has been reported that ataxin-3 is the target of caspase 1 at 241, 244, and 248 amino acids, and that the brains from SCA3 transgenic mice contained a C-terminal fragment (Goti et al. 2004; Berke et al. 2004; Colomer Gould et al. 2007). It has also been shown that normal ataxin-3 was extensively proteolysed in COS-7 cells with cleavage occurring at six other sites throughout the protein. Interestingly, mutant ataxin-3 underwent proteolysis to a much lesser extent with expanded polyQ masking the C-terminal cleavage sites (Pozzi et al. 2008) suggesting that SCA3 pathology can arise also through the accumulation of uncleaved protein. On the other hand, toxic fragment hypothesis has been introduced by Haacke et al. (2006), stating that expanded polyQ fragment must be released from the protective sequence context of full-length protein to be able to trigger the aggregation process. It is therefore possible that a toxic fragment initiates the aggregation process and later, as the intracellular levels of uncleaved expanded ataxin-3 increase, this process becomes independent of the presence of toxic fragments (Pozzi et al. 2008).

    Mutant htt alters calcium signaling and increases Ca2+ levels in different cells (Bezprozvanny and Hayden 2004). Consequently, a calcium-dependent protease, calpain is activated in the brains of HD patients, where calpain co-localizes with htt aggregates and calpain cleavage products are present (Gafni and Ellerby 2002). Another cleavage of mutant htt is performed by an aspartyl protease between amino acids 104 and 114 (Lunkes et al. 2002). It is not known yet whether this cleavage is specific to mutant htt.

    Proteolytic cleavage may be not an essential step in pathogenesis of all polyQ diseases. For polyQ protein normally localized in the cytoplasm, such as htt, ataxin-3, or atrophin-1, proteolysis appears to facilitate their translocation into nucleus enhancing their toxicities. For proteins already localized in the nucleus, for example, TATA-binding protein (TBP) or ataxin-7, proteolytic cleavage may not have so important role.

    Phosphorylation of polyglutamine proteins

    Phosphorylation may affect proteolytic cleavage, the initial conversion of mutant polyQ proteins to pathogenic conformations and nuclear transport (Warby et al. 2009). The impact of htt phosphorylation on HD pathogenesis has been demonstrated in several studies. Phosphorylation of htt at S421 by protein kinase Akt or serum and glucocorticoid-induced kinase SGK is neuroprotective against HD cellular toxicity (Humbert et al. 2002; Rangone et al. 2004). Recently, it has been reported that S421 phosphorylation of htt decreased the accumulation of both full-length htt and htt fragments (including those generated by caspase 6) in the nucleus (Warby et al. 2009). This observation is consistent with a previous study showing the effect phosphorylation status of S421 on vesicular transport in neurons. When phosphorylated, htt recruited kinesin-1 to the dynactin complex on vesicles and microtubules and promoted the anterograde transport of vesicles away from nucleus. Conversely, when htt was not phosphorylated, kinesin-1 was released and vesicles were more likely to undergo retrograde transport back to nucleus (Colin et al. 2008). Calcineurin (calcium/calmodulin-regulated serine/threonine protein phosphatase) dephosphorylates S421 in vitro and in cells. Using a rat model of HD with lentiviral-mediated expression of a polyQ htt fragment in the striatum, it has been shown that inhibition of calcineurin activity either by over-expression of a dominant-interfering form, by RNA interference, or by the specific inhibitor FK506, leads to an increased phosphorylation of S421 and prevents mutant htt-mediated death of striatal neurons (Pardo et al. 2006).

    Phosphorylation of mutant htt at S434 by serine-threonine kinase Cdk5 reduced htt cleavage at 513 amino acid by caspase 3 and inhibited polyQ aggregation and cytotoxicity (Luo et al. 2005). The aggregation and cellular toxicity of mutant htt has also been shown to decrease upon phosphorylation at S536, which inhibits the cleavage of htt by calpain at this amino acid (Gafni et al. 2004; Schilling et al. 2006). Many phopshorylation sites were identified throughout whole htt sequence and several kinases have been proposed to be involved including MAPK/ERK kinase 1/2 or extracellular signal regulated kinase 1 (Schilling et al. 2006). Further studies are needed to understand how all the phosphorylation events affect the function and regulation of htt and their role in HD pathogenesis.

    As mentioned in the section `Phosphorylation of polyglutamine proteins', htt phosphorylation by Akt has been shown neuroprotective. In contrary, phosphorylation of mutant ataxin-1 by Akt at S776 promotes its binding to 14-3-3, which in turn leads to ataxin-1 accumulation and neurodegeneration (Chen et al. 2003). Abnormal nuclear stabilization of mutant ataxin-1 by Akt was shown to have critical role in SCA1 pathogenesis in ataxin-1[82Q]-A776 transgenic mice (Emamian et al. 2003). In a recent study, however, inhibition of Akt either in vivo or in cerebellar extract did not decrease the phosphorylation of ataxin-1 at S776 arguing against involvement of Akt as the key kinase in SCA1 pathogenesis. The same study suggested that cAMP-dependent protein kinase is responsible for the S776 phosphorylation in the cerebellum (Jorgensen et al. 2009). Other phosphorylation sites that may play a role in ataxin-1 localization and interactions in SCA1 are S239 and probably T236. S239 is within the consensus sequences targeted by casein kinase 1 (CK1), Cdc2/Cdk5, and extracellular signal regulated kinase (Vierra-Green et al. 2005).

    Ataxin-3 has been reported to undergo phosphorylation at S256 by glycogen synthase kinase 3β and this phosphorylation resulted in reduced mutant ataxin-3 aggregation in vitro (Fei et al. 2007). Another kinase, CK2 has been shown to be a regulator of nuclear localization of ataxin-3. CK2-dependent phosphorylation of ataxin-3 at S236 and S340/S352 decreased the appearance of nuclear inclusions and controlled the nuclear translocation of ataxin-3 providing a reasonable therapeutic approach for SCA3 (Mueller et al. 2009).

    The polyQ expansion in AR activates major mitogen-activated protein kinase pathways, from which the p44/42 pathway in turn phosphorylates at S514. This phosphorylation has been shown to promote the mutant AR cytotoxicity (LaFevre-Bernt and Ellerby 2003). In atrophin-1, S734 has been identified as a phosphoacceptor in cell cultures and rat brain. Both normal and mutant protein forms are phosphorylated by c-Jun N-terminal kinase (JNK), but precise analyses revealed a reduced affinity of c-Jun N-terminal kinase (JNK) to expanded AR (Okamura-Oho et al. 2003). Biological significance of atrophin-1 phosphorylation needs to be evaluated.

    With the development of specific inhibitors of kinases and proteases, modulating phosphorylation and proteolytic processing of mutant polyQ proteins could be a promising treatment strategy.

    Sumoylation of polyglutamine proteins

    Small ubiquitin-like modifier (SUMO) proteins bind covalently to specific lysine residues in target proteins regulating their cellular localization, protein–protein interactions, and transcription factors transactivation (Sarge and Park-Sarge 2009). Antagonistic relationship between sumoylation and ubiquitination has been proposed as they share common consensus sequence (Muller et al. 2001). Enhanced immunoreactivity of SUMO1 has been observed in affected neurons of HD, SCA1, SCA3, and DRPLA patients (Ueda et al. 2002).

    Sumoylation of N-terminal fragment htt at K6, K9, and K15 was suggested to enhance the stability and reduce the aggregation of mutant htt, which however seemed to result in toxic intermediate polyQ oligomers accumulation and transcriptional repression (Steffan et al. 2004). Importantly, the same lysines are also ubiquitinated, however generally, sumoylation made pathology substantially worse while ubiquitination made it modestly better in Drosophila model of HD suggesting that (Steffan et al. 2004).

    In ataxin-1, five major sumoylation sites have been identified (K16, K194, K610, K697, and K746). Sumoylation of ataxin-1 was dependent on its nuclear localization, phosphorylation at S776, self-association domain and polyQ length, all having role in the subcellular distribution of ataxin-1 in COS-1 cells. Nuclear localization, on the other hand, did not appear dependent on sumoylation, although it could have role in the efficiency of nucleocytoplasmic shuttling of ataxin-1. The lack of nucleus-to-cytoplasmic traffic of mutant ataxin-1 with disrupted interactions of ataxin-1 with other proteins leading to transcription dysregulation could be attributed to the decreased sumoylation (Riley et al. 2005).

    Neuronal intranuclear inclusions in brains of DRPLA patients and mutant atrophin-1 aggregates in DRPLA cellular model were highly sumoylated. The results of this study showed active role of sumoylation in the pathogenesis of DRPLA accelerating aggregate formation and cell death (Terashima et al. 2002).

    Androgen receptor is another target of sumoylation and this modification inhibits AR activity while significantly reducing mutant AR aggregation without affecting the levels of the receptor. Interestingly, SUMO inhibited AR aggregation through a unique mechanism that does not depend on critical features essential for its interaction with canonical SUMO-binding motifs (Mukherjee et al. 2009). Precise role of sumoylation in the polyQ-related neurodegenration is still not completely understood. Interestingly, while sumoylation appears neuroprotective in SBMA and SCA1, in case of HD and DRPLA it seems to accentuate the pathological process.

    Role of nuclear localization of mutant polyglutamine proteins

    Nuclear accumulation of expanded polyQ proteins has been shown to contribute to pathogenesis of several polyQ diseases (Saudou et al. 1998; Peters et al. 1999; Ross et al. 1999; Orr 2001; Klement et al. 1998) by affecting gene expression (Zoghbi and Orr 2000) or by disrupting nuclear organization and function (Sun et al. 2007). Yang et al. (2002) showed that polyQ aggregates were toxic only when localized in the nucleus but not in the cytoplasm. Proteolytic cleavage of mutant polyQ proteins with predominantly cytoplasmic localization of the normal copies may enhance the nuclear access as the upper limit for passive nuclear translocation has been proposed at about 46 kDa (Rubinsztein et al. 1999). Putative nuclear localization signal sequences have been found in some polyQ proteins including ataxin-1, ataxin-7, and atrophin-1 (Schilling et al. 1999b; Kaytor et al. 1999; Klement et al. 1998). Similarly, nuclear export signal sequences have been identified in htt, ataxin-7, and atrophin-1 with expanded polyQ tracts reducing nuclear export of these proteins (Cornett et al. 2005; Nucifora et al. 2003; Xia et al. 2003; Taylor et al. 2006). Nuclear accumulation of mutant proteins and inclusions have been identified as predominant in HD, SCA1, SCA3, SCA7, SCA17, DRPLA, and SBMA patients (Schols et al. 2004); however, cytoplasmic inclusions have also been found in affected brain regions of HD and SCA2 patients (Huynh et al. 2000; DiFiglia 2002).

    Vast majority of htt is normally localized to cytoplasm (Gafni et al. 2004) but in HD, htt migrates to nucleus (DiFiglia et al. 1997) where it disrupts the activities of transcription factors and alters the normal transcriptional profile of neurons (Chan et al. 2002; Panov et al. 2002). Nuclear translocation of mutant htt is required to induce neurodegeneration. Inhibiting this translocation by blocking htt cleavage reduces the toxicity and retards the disease progression (Saudou et al. 1998; Wellington et al. 2000; Gafni et al. 2004). On the other hand, cytoplasmic mutant htt can inhibit the axonal transport and disrupt synaptic function and glutamate release (Gunawardena et al. 2003; Li et al. 2000a, 2001, 2003; Szebenyi et al. 2003) suggesting that both nuclear and cytoplasmic expanded htt polyQ is able to trigger certain pathological events. Htt contains two nuclear export signal sequences, one near the C-terminal and the other one being the N-terminal 17 amino acids. PolyQ expansion together with htt cleavage (e.g. by caspase 6) impedes their function resulting in accumulation of mutant htt in the nucleus (Cornett et al. 2005; Xia et al. 2003). The first 17 amino acids have also been identified as directing htt to various subcellular compartments including plasma membrane, autophagic vesicles, endoplasmic reticulum (ER) and Golgi apparatus (Kegel et al. 2005; Atwal et al. 2007; Rockabrand et al. 2007).

    Ataxin-1 contains a C-terminal nuclear localization signal and is therefore localized to the nucleus (Skinner et al. 1997). The nuclear localization of ataxin-1 is essential for neurotoxicity and SCA1 disease as the disruption of nuclear localization signal inhibited the formation of intranuclear inclusions and reduced the Purkinje cells pathology and SCA1 phenotype in a transgenic mouse model (Klement et al. 1998).

    Ataxin-2 is a cytoplasmic protein with highest expression in Purkinje cells, the most affected neurons in SCA2 (Huynh et al. 1999). Ubiquitinated intranuclear inclusions have been found only in few pontine neurons of SCA2 patients, but not in Purkinje cells (Koyano et al. 1999). Cytoplasmic localization with microaggregation or accumulation of ataxin-2 has been shown to be sufficient to cause SCA2 pathology in humans and mice. The 42 kDa, N-terminal fragments produced by proteolytic cleavage of ataxin-2 are believed to act as seeds of cytoplasmic aggregation in the Purkinje cells of SCA2 patients (Huynh et al. 1999, 2000).

    Ataxin-3 is mostly cytoplasmic protein, but, depending on cell type, it displays subcellular distribution involving both the cytoplasm and the nucleus (Wang et al. 1997; Paulson et al. 1997; Tait et al. 1998; Trottier et al. 1998). Intranuclear inclusions have been observed in human SCA3 brains (Paulson et al. 1997; Schmidt et al. 1998). Recently, it has been found that full-length ataxin-3, regardless of polyQ tract length, is also localized to mitochondrial matrix and membrane and that the polyQ expansion promotes binding to the mitochondrial membrane. Although lacking predicted nuclear or mitochondrial localization signal, first 27 amino acids of ataxin-3 might play role in cell sorting as cleavage at this site caused absence of ataxin-3 in the nucleus or mitochondria (Pozzi et al. 2008). It is apparent that the disease results from the presence of mutant ataxin-3 in several cellular compartments such as nucleus, cytosol and mitochondria although it remains to clarify which localization causes primary abnormalities in SCA3 pathogenesis.

    Ataxin-7 contains a functional nuclear localization signal and is primarily localized in the nucleus (Kaytor et al. 1999) where it is a subunit TFTC/STAGA transcriptional complex (Helmlinger et al. 2004, 2006). Nuclear localization of mutant ataxin-7 appears to be essential for toxicity with profound effect on transcriptional dysregulation (Helmlinger et al. 2004; La Spada et al. 2001). Similarly in SCA17, the localization of TBP as a transcription factor is nuclear, where the polyQ expansion results in transcriptional dysregulation (Friedman et al. 2007).

    Atrophin-1 contains both nuclear localization (at N-terminal) and export (at C-terminal) signal sequences resulting in ubiquitous cellular localization (Schilling et al. 1999b; Nucifora et al. 2003). Neuronal inclusions in DRPLA are predominantly nuclear, but they do not seem to correlate with the sites of neurodegeneration (Sato et al. 1999; Koyano et al. 2002). On the other hand, an N-terminal fragment of mutant atrophin-1 was found to accumulate in the neuronal nuclei of DRPLA patients and model mice while lacking its C-terminal with nuclear export signal. This accumulation occurred before the first symptoms of the disease (Schilling et al. 1999b; Nucifora et al. 2003). In addition, mutation within the nuclear localization signal resulted in retention of the protein in the cytoplasm and reduced the toxicity of atrophin-1 (Nucifora et al. 2003).

    From molecular point of view, SBMA is a unique type of polyQ disease, because a specific ligand can alter the subcellular localization of AR. The unliganded AR is localized to the cytoplasm while binding of an androgen ligand, such as testosterone, translocates AR to the nucleus and triggers the expression of androgen-responsive genes (Stenoien et al. 1999; Katsuno et al. 2002). Intranuclear inclusions have been observed in SBMA patients in motor neurons in brainstem and spinal cord (Li et al. 1998). SBMA affects only males, with female carriers remaining largely asymptomatic as binding of testosterone accelerates AR transportation to the nucleus (Brooks and Fischbeck 1995). Nuclear localization of AR seems to be the main site of SBMA-related pathogenesis (Brooks and Fischbeck 1995; Katsuno et al. 2002) further supporting the concept that toxicity of polyQ proteins depends on their localization.

    Role of heat-shock proteins in preventing aggregation

    The presence of expanded polyQ proteins causes a cellular stress response, which may include the up-regulation of heat-shock proteins (HSPs). HSPs levels have been shown to have a biphasic response to expanded polyQ expression where the initial phase of increased expression of HSPs is induced by cellular stress response, followed by progressive reduction of HSPs levels during later stages of neurodegeneration (Huen and Chan 2005; Hay et al. 2004). HSPs recognize misfolded proteins, stabilize them in monomeric conformation and suppress their aggregation (Muchowski et al. 2000). By interacting with the mutant proteins, chaperones may prevent abnormal interactions with other cellular proteins. The critical role of chaperones in polyQ diseases has been evidenced in many studies (Muchowski 2002). The most investigated chaperones have been Hsp70 and heat-shock cognate 70. A co-chaperone class known as Hsp40, which includes human DNAJ1 (Hdj1) and Hdj2, facilitates the action of Hsp70. Both Hsp40 and Hsp70 are involved in the clearance of misfolded proteins via the ubiquitin-proteasome system (UPS) pathway (Bercovich et al. 1997).

    The members of Hsp40 and/or Hsp70 families co-localize with nuclear aggregates in human brain tissue of SCA1 (Cummings et al. 1998), SCA3 (Schmidt et al. 2002), and SCA7 (Zander et al. 2001) patients along with SCA1 (Cummings et al. 1998) and SCA7 (Yvert et al. 2000) transgenic mouse brains. Hdj2 and heat-shock cognate 70 were associated with nuclear aggregates in HD cell culture as well as in transgenic mice models and their over-expression reduced the aggregate formation (Jana et al. 2000). An Hsp90 family member, Hsp84, was found to co-localize with nuclear aggregates in HD transgenic mouse brains, and its over-expression reduced both polyQ aggregation and toxicity (Mitsui et al. 2002). Hsp105α in SBMA patients’ and transgenic mouse brains suppresses polyQ toxicity (Ishihara et al. 2003).

    The over-expression of Hsp40 and/or Hsp70 chaperones in cell culture, or in in vivo polyQ diseases models reduced the inclusion formation of mutant htt, AR, ataxin-3, or polyQ-green fluorescent fusion proteins and/or suppressed cell death (Kobayashi et al. 2000; Chai et al. 1999a; Kazemi-Esfarjani and Benzer 2000; Stenoien et al. 1999; Warrick et al. 1999; Jana et al. 2000; Muchowski et al. 2000; Cummings et al. 1998). Over-expression of Hsp70 in SCA1 transgenic mice also suppressed neurodegeneration and improved motor function (Cummings et al. 2001). The aggregation reduction by HSPs probably results from increased solubility of polyQ proteins and improved access to the degradation machineries. On the other hand, it has been shown that the suppression of the polyQ cytotoxicity did not depend on the aggregation reduction by Hsp40 and Hsp70 (Zhou et al. 2001). Moreover, a small chaperone Hsp27 could reduce the mutant htt toxicity by inhibition of the accumulation of the reactive oxygen species (ROS) without affecting the aggregation (Wyttenbach et al. 2002). Thus, whether chaperones are directly involved in aggregate formation itself remains to be determined.

    The C-terminal Hsp70-interacting protein (CHIP) links the molecular chaperones with UPS. CHIP interacts with Hsp70 and Hsp90 through its tetratricopeptide repeat domains and with proteasome through an E4/U-box domain (Jiang et al. 2001; Ballinger et al. 1999). CHIP has been demonstrated to reduce the expanded polyQ aggregation, to enhance the clearance of mutant polyQ proteins and to reduce their cytotoxicity in different polyQ diseases in vitro and in vivo models (Jana et al. 2005; Al-Ramahi et al. 2006; Miller et al. 2005; Williams et al. 2009). These studies provided further evidence for the protein misfolding and aggregation model for polyQ toxicity.

    Apoptosis and polyglutamine diseases

    The mechanism of cell death in polyQ diseases is very complex as many processes triggered by the presence of expanded polyQ proteins can lead to the cell death, such as direct activation of cell death pathways, mitochondrial abnormalities, transcriptional dysregulation, proteasome impairment, defects in axonal transport, or unfolded protein response. Other pathophysiological events such as excitotoxicity, metabolic stress, or accumulation of free radicals may promote the cell fate toward death by further enhancing the mitochondrial dysfunction. Many proteins usually involved in apoptosis are sequestered, redistributed or activated by expanded polyQ proteins. Caspases are activated (Wang et al. 1999; Li et al. 2000b), and can be recruited in inclusions (Sanchez et al. 1999). Cytochrome c release was observed in cells transfected with mutant htt (Li et al. 2000b). Cytochrome c released from mitochondria interacts with apoptotic peptidase activating factor 1 (Apaf-1) and procaspase 9, which in turn activate caspase 3 and the caspase cascade (Li et al. 1997). On the other hand, certain apoptotic events may be inhibited by the presence of mutant htt. For example, htt aggregates were shown to sequester the proapoptotic protein kinase Cδ, and therefore block the protein kinase Cδ-dependent DNA fragmentation (Zemskov et al. 2003).

    The expression of expanded polyQ protein in rat neurons resulted in the activation and recruitment of caspase 8 to the aggregates (Sanchez et al. 1999). In this apoptotic pathway, the caspase cascade is initiated and caspase 8 deficiency or expression of a dominant-negative mutant suppressed neurodegeneration. The sequestration of caspase 8 to polyQ aggregates required Fas-associated protein with death domain, because both caspase 8 binding and neuronal death were blocked by dominant-negative Fas-associated protein with death domain (Sanchez et al. 1999).

    Other caspases may also have a role in polyQ toxicity. Activation of caspases 1 and 3 in a HD transgenic mouse model has been demonstrated. Inhibition of these two caspases by minocycline, had neuroprotective effect, delayed the symptoms and extended the lifespan of these mice (Chen et al. 2000).

    Calcium increases outer mitochondrial membrane permeability resulting in release of proapoptotic proteins into the cytosol including cytochrome c and apoptosis inducing factors. Caspases 1 and 9 were also activated through disruption of mitochondrial membrane potential and released cytochrome c into the cytosol in mouse neuroblastoma cells expressing N-terminal htt with expanded polyQ (Jana et al. 2001). This process was associated with impaired proteasomal function. It has been also shown that the altered Ca2+ signaling is strongly involved in the mitochondria-mediated activation of the apoptotic caspases and calpain in HD and SCA3 models (Tang et al. 2003; Chen et al. 2008b). Expanded polyQ protein causes increased mitochondrial apoptosis activation by at least three synergistic mechanisms. Mutant htt elevated cytosolic Ca2+ levels by enhancement of NMDA receptor function and by strong binding to inositol (1,4,5)-trisphosphate receptor type 1 (IP3R1). IP3R1 is then sensitized to activation by IP3 in medium spiny neurons resulting in abnormal Ca2+ influx via NMDA receptor and Ca2+ release from ER through IP3R1. The presence of expanded htt may also directly affect the permeability of the mitochondrial membrane (Bezprozvanny and Hayden 2004).

    An effect of expanded ataxin-3 on the expression of Bcl-2 family proteins has been reported (Chou et al. 2006; Tsai et al. 2004). The expression of mutant ataxin-3 in cerebellar, striatal, or substantia nigra neurons specifically up-regulated proapoptotic Bax and down-regulated Bcl-xL proteins which may lead to mitochondrial release of apoptogenic proteins and apoptotic cell death (Chou et al. 2006). Similar effect on Bax and Bcl-xL had the expression of expanded ataxin-7 in primary neuronal culture from cerebellum or neocortex (Wang et al. 2006). These studies further underlined the complex effect of expanded polyQ on apoptotic cell death.

    A unique aspect of the expression of mutant ataxin-2 was observed in PC12 (cells derived from pheochromocytoma of rat adrenal medulla) and COS-7 cells, where the apoptotic cell death through caspase 3 activation was correlated with disruption of the Golgi complex. This is probably because of the primary localization of normal but not expanded ataxin-2 in Golgi apparatus (Huynh et al. 2003). These data supported previous observations that ataxin-2 co-localized with its binder, the ataxin-2-binding protein 1, which was also co-localized with Golgi proteins (Shibata et al. 2000). When compared with most of the other polyQ proteins, the ataxin-2-induced cell death does not require nuclear localization (Huynh et al. 2000).

    Polyglutamine aggregation can also contribute to apoptosis activation by reduced binding to proapoptotic factors, which are normally sequestered by wild type polyQ proteins. Huntingtin-interacting protein 1 (Hip-1) is a protein with a pseudo death effector domain (Hackam et al. 2000) and was originally identified as an Hip. The affinity of Hip-1 to htt is reduced by the presence of expanded polyQ, resulting in elevated free Hip-1, which forms heterodimers with a Hip-1 interactor, Hippi, and they are able to activate apoptosis through binding and activating caspase 8 (Gervais et al. 2002).

    Cell death in SCA6 has been attributed to CaV2.1 Ca2+ channel dysfunction, which normally prevents cell death by regulating Ca2+ influx, suggesting that the polyQ expansion in α1A subunit of voltage-dependent P/Q type calcium channel (CACNA1A) results in a channelopathy (Matsuyama et al. 2004).

    In SCA17, the polyQ expansion in the TBP affects the interaction of TBP with its binding partners impairing the transcription of the target genes (Friedman et al. 2007). It was demonstrated, that the mutant TBP displays reduced binding to TATA box DNA in vitro. Interestingly, the same study also showed the inhibition of the TATA-dependent transcriptional activity by a mutant TBP fragment lacking the C-terminal DNA-binding domain. This fragment formed intranuclear aggregates and caused severe neurological phenotype in the SCA17 TBP-105Q-T transgenic mice indicating that the mutant TBP can induce neurotoxicity independent of its association with DNA (Friedman et al. 2008).

    A recent study defined the pathway by which the AR with expanded polyQ activates neuronal apoptosis in SBMA. The expression of mutant N-terminal AR induced intrinsic pathway-mediated apoptosis. It is initiated by activation of c-JUN by Jun N-terminal kinase resulting in death protein 5/hara-kiri [member of the proapoptotic Bcl-2 homology domain 3 (BH3)-only family] up-regulation leading to Bax activation (Young et al. 2009).

    Another polyQ-related insult leading to cell death is ER stress with consequent activation of apoptosis signal-regulating kinase 1 (Nishitoh et al. 2002). This hypothesis was supported by a study using a peptide inhibitor of apoptosis signal-regulating kinase 1, as it was able to reduce the apoptosis in a cellular model of DRPLA (Kariya et al. 2005).

    Mitochondrial involvement in polyglutamine diseases

    Impairment of mitochondrial functions is one of the key events in polyQ diseases leading to cell death via activation of apoptotic cascades. The process of mitochondrial dysfunction is accompanied by impaired respiration, stress-induced mitochondrial depolarization, increased free radical production with oxidative damage, and abnormal energy metabolism in polyQ diseases (Grunewald and Beal 1999; Panov et al. 2002; Browne and Beal 2006).

    The importance of mitochondria in polyQ diseases pathogenesis and related cell death had been suggested by mimicking the HD phenotype using the complex II respiratory chain inhibitors 3-nitropropionic acid or malonate (Brouillet et al. 1995; Beal et al. 1993). In toxin models of HD, the activities of complex II and III were reduced and energy metabolism was impaired as observed in human HD brains and in cells derived from HD model mice, respectively (Gu et al. 1996; Milakovic and Johnson 2005). Another study supporting the direct effect of expanded polyQ proteins on mitochondria showed that polyQ proteins increased ROS in isolated mitochondria (Puranam et al. 2006). Mutant htt was also observed on neuronal mitochondrial membranes by electron microscopy and the incubation of normal mitochondria with fusion protein with expanded polyQ reduced their Ca2+ retention capacity (Panov et al. 2002). It has been demonstrated that recombinant mutant htt directly induced mitochondrial permeability transition (MPT) pore opening in isolated mouse liver mitochondria. Mutant htt decreased the Ca2+ threshold necessary to trigger MPT pore opening accompanied by a significant release of cytochrome c (Choo et al. 2004).

    One study showed that the Ca2+ loading capacity of isolated mitochondria from YAC128 and R6/2 HD mouse models was increased compared with control animals, however this was not true in the mitochondria from an Hdh150 knock-in mouse model (Oliveira et al. 2007). In the knock-in mouse models, mitochondria were likely not to function at maximal capacity under resting conditions suggesting that only under conditions of neuronal stress does the mitochondrial impairment contribute to HD pathogenesis (Milakovic and Johnson 2005; Oliveira et al. 2007). No significant defects in basal respiratory capacity, basal ATP synthesis, or uncoupling were observed in Hdh150 striatal neurons, suggesting that there was no major bioenergetic defect in these neurons. Under neuronal stress though, some neurons may be unable to meet the increased ATP demand because of the defects in ATP synthesis and/or export and they become more vulnerable to Ca2+ deregulation (Oliveira et al. 2007). Altered Ca2+ buffering capacity of mitochondria is therefore probably a cause of enhanced NMDA-induced apoptosis occurring in HD brains (Fei et al. 2007). In a cellular and a Caenorhabditis elegans HD model, it has been shown that the over-expression of mutant htt diminished the normal dynamics of mitochondria fusion and fission interfering with mitochondrial ATP generation providing another potential therapeutic target (Wang et al. 2009).

    The expression of mutant AR caused mitochondrial abnormalities leading to caspase activation. Surprisingly, a polyQ-dependent decrease in mitochondrial number and impaired mitochondrial membrane potential was observed with increased ROS. Mutant AR also altered the expression of genes important for mitochondrial function including one of the key regulators of mitochondrial biogenesis and function, peroxisome proliferator-activated receptor γ coactivator-1 (PGC-1β), and its target genes peroxisome proliferator-activated receptor γ, and mitochondrial transcription factor A. The mitochondrial genes regulated by mitochondrial transcription factor A and several cellular antioxidant genes were also down-regulated suggesting a profound role of mitochondrial dysfunction in SBMA pathogenesis (Ranganathan et al. 2009).

    Similarly to previously cited studies, PGC-1α, another member of PGC-1 complex, is a mitochondrial biogenesis and function co-activator which regulates mitochondria response to oxidative stress (St-Pierre et al. 2006). Mutant htt was reported to directly bind to PGC-1α and interfere with its function (Weydt et al. 2006), supporting the connection between oxidative stress and mitochondrial dysfunction in HD. Moreover, PGC-1α is regulated by cAMP response element (CRE)-binding protein (CREB), a transcription factor down-regulated by expanded polyQ htt (Steffan et al. 2000) and mutant htt was shown to interfere with transcription of PGC-1α (Cui et al. 2006).

    Expanded polyQ proteins were reported to impair axonal trafficking, which may lead to defective mitochondrial distribution and function providing another causal link between mitochondria and polyQ diseases (Szebenyi et al. 2003; Chang et al. 2006). This interaction correlates with decreased distribution and transport rate of mitochondria in the processes of cultured neuronal cells and reduced level of ATP in synaptosomal fractions (Orr et al. 2008). Mutant ataxin-3 has been recently reported to decrease the activities of antioxidant enzymes levels causing increased damage of mitochondrial DNA (Yu et al. 2009). Based on number of studies, it can be stated for almost a certainty that mitochondrial impairment is a common feature in the pathogenesis of polyQ diseases.

    Involvement of proteasome in polyglutamine diseases

    The UPS is a major cellular protein degradation pathway clearing short-lived and damaged proteins (Goldberg 2003; Chou et al. 2006). UPS degrades many regulatory proteins having roles in distinct cell signaling pathways and synaptic function and plasticity (Bingol and Schuman 2006). Impairment of the UPS, as a major regulator of normal cellular functioning and as a detoxification machinery targeting damaged proteins for degradation, may have therefore lethal consequences on the affected cells.

    The impairment of UPS by mutant htt and other expanded polyQ proteins has been proposed and shown in several studies (Bence et al. 2001; Jana et al. 2001; Zemskov and Nukina 2003; Khan et al. 2006; Bennett et al. 2007) although it is still considered controversial. The relationship between UPS- and polyQ-related diseases is evidenced by the ubiquitination of several aggregated polyQ proteins (Ross and Poirier 2004). Indeed, expanded ataxin-1, -3, and -7, as well as atrophin-1, AR, and htt colocalize with the proteasome and several studies have shown a redistribution of the proteasome complex to the inclusions (Paulson et al. 1997; Chai et al. 1999b; Stenoien et al. 1999; Yvert et al. 2000; Cummings et al. 1998). Ubiquitinated ataxin-3 inclusions were observed in brains of SCA3 patients (Paulson et al. 1997). Ubiquitinated inclusions have been also reported in SCA2 and SCA7 brains (Holmberg et al. 1998; Koyano et al. 1999), although the exact distribution and relationship to disease remains unclear (Huynh et al. 2000). SCA7 intranuclear inclusions are ubiquitinated in severely degenerated areas such as the inferior olive, yet also in the cerebral cortex, which is unaffected (Holmberg et al. 1998; Yvert et al. 2001). In SCA17, ubiquitin- and TBP-positive inclusions were found in the putamen, which shows pattern of neurodegeneration in affected patients (Nakamura et al. 2001). The ubiquitination of the polyQ inclusions suggests that the UPS is attempting to clear the mutant proteins. Consistent with this, the 20S proteasome was shown to colocalize with ataxin-1 (Cummings et al. 1998) and ataxin-3 aggregates (Chai et al. 1999b).

    Preventing the ubiquitination of mutant ataxin-1 resulted in increased cellular toxicity indicating that the ubiquitination of expanded polyQ may play an important detoxifying role (Cummings et al. 1999). Ataxin-3 was shown to interact with HHR23A and HHR23B, the human homologs of yeast DNA repair protein Rad23 (Wang et al. 2000). The recruitment of HHR23A to the intranuclear inclusions by mutant ataxin-3 probably impairs its normal interaction with UPS as Rad23 was shown to recruit polyubiquitinated proteins to the proteasome and facilitate their degradation in yeast (Verma et al. 2004; Elsasser et al. 2004). Proteasome subunits were also found colocalized with mutant ataxin-7 intranuclear inclusions in SCA7 transgenic mice (Yvert et al. 2001), transfected cells, and human brain (Zander et al. 2001). Moreover, S4 subunit of the 19S proteasome is specifically depleted in the regions of the brain affected in SCA7. Furthermore, a yeast two-hybrid assay has demonstrated interaction between ataxin-7 and this subunit (Matilla et al. 2001).

    Two basic mechanisms are likely to account for the UPS impairment by expanded polyQ proteins. First, sequestration into polyQ aggregates and/or altered subcellular localization of the proteasome components may be responsible for the UPS dysfunction. This hypothesis is supported by sequestration of ubiquilin 1 and 2 as well as Tollip, which are thought to modulate UPS, in polyQ aggregates (Doi et al. 2004). Second, proteins with expanded polyQ may block the proteasome preventing other substrates to enter. Number of studies have been published supporting or disputing these theories and are discussed elsewhere (Davies et al. 2007). It has been also proposed that the reduction of UPS activity may result from caspase-dependent cleavage of proteasome components while the aggregated proteins induce apoptosis (Sun et al. 2004).

    Although the precise role of UPS in polyQ diseases pathogenesis remains to be clarified, it is evident that the alteration of its activity is one of the common features in this group of disorders.

    Transcriptional dysregulation in polyglutamine diseases

    The mutant polyQ proteins interact with many other cellular proteins and may sequester them into cytoplasmic or nuclear aggregates (Mitsui et al. 2002; Doi et al. 2008). Htt, ataxins, atrophin-1, and AR have several interactions in common supporting the recent knowledge of pathogenic effect related to expanded polyQ stretch. As many of the aberrantly bound proteins are transcription-related factors, their dysbalance may have profound effects on gene expression with potentially toxic effects. Analyses of postmortem tissues of HD patients and HD or DRPLA models have revealed altered expression of many genes (Kotliarova et al. 2005; Chan et al. 2002; Kuhn et al. 2007; Luthi-Carter et al. 2002). Oyama et al. found that expression of sodium channel subunit β4 is severely reduced in the brains at an early stage in HD model mice, which might be involved in dendritic abnormalities of neurons observed in HD (Oyama et al. 2006; Miyazaki et al. 2007).

    A variety of nuclear proteins relevant to transcription have been suggested to interact with different mutant polyQ proteins (Okazawa 2003). Some of them bind to more polyQ disease proteins while some are associating specifically, what could be one of the underlying mechanisms for the selective neurodegeneration in polyQ diseases. For example, the interaction of expanded ataxin-1 with leucine-rich repeat acidic nuclear protein 32 (Matilla et al. 1997), PolyQ-binding peptide-1 (Okazawa et al. 2002), Gfi-1/Senseless (Tsuda et al. 2005), and Boat (Mizutani et al. 2005) might lead to selective degeneration in cerebellum. Protein phosphatase 2 activity is regulated by leucine-rich repeat acidic nuclear protein 32, the factor predominantly expressed in Purkinje cells and therefore the dysregulation of protein phosphatase 2 is a possible mechanism of neurodegeneration in SCA1 (Matilla and Radrizzani 2005). Ataxin-7 is a subunit of TFTC/STAGA transcriptional complex, which interacts with the photoreceptor-specific transcriptional activator cone–rod homeobox protein (Helmlinger et al. 2004; Palhan et al. 2005; La Spada et al. 2001). Mutant ataxin-7 sequesters cone–rod homeobox protein and suppresses the acetyltransferase activity of the TFTC/STAGA complex, which results in down-regulation of genes vital for retinal function and accounts for the retinal degeneration seen in SCA7 (Palhan et al. 2005; La Spada et al. 2001).

    A common transcriptional activator, CREB, along with its co-activator CREB-binding protein (CBP), has been strongly implicated in expanded polyQ-induced gene repression (Steffan et al. 2000). CBP is an important mediator of survival signals in neurons. It has histone acetyltransferase activity, which is important for allowing transcription factors access to DNA. In the presence of mutant htt or atrophin-1, CBP is sequestered into aggregates (Nucifora et al. 2001). CBP recruitment to intranuclear inclusions or the interaction with mutant polyQ proteins also occurs in SBMA, SCA3, and SCA7 (McCampbell et al. 2000; Li et al. 2002a; Strom et al. 2005). In SCA1, CBP was not irreversibly trapped into nuclear aggregates but it was rapidly exchanged. Even short residence time of CBP within inclusions, however, may be sufficient to disrupt its normal function in maintaining cellular homeostasis (Stenoien et al. 2002). It appears that the disruption of CREB/CBP-mediated gene expression may be a common mechanism of neurodegeneration in polyQ repeat diseases. On the contrary, it was reported that in R6/2 HD mouse model, the CREB-mediated transcription was increased (Obrietan and Hoyt 2004) and CBP was not depleted in HD model mice brains (Yu et al. 2002).

    TATA-binding protein was found to colocalize with htt, ataxin-2, ataxin-3, and atrophin-1 inclusions in human brains (Perez et al. 1998; Uchihara et al. 2001; van Roon-Mom et al. 2002), but similarly to CBP, the levels of TBP were not reduced in HD model mice (Yu et al. 2002). Specificity protein 1 (SP1), contains a glutamine-rich activation domain, which is responsible for regulation of the transcriptional machinery of transcription factor II D (TFIID), a protein complex composed of TBP and multiple TBP-associated factors (TAFIIs), including TAFII130 (Tanese and Tjian 1993). The interaction of SP1 with mutant htt leads to suppression of SP1 transcriptional activity resulting in down-regulation of dopamine D2 or nerve growth factor receptors (Li et al. 2002b; Dunah et al. 2002). The over-expression of SP1 and TAFII130 was able to ameliorate the mutant htt toxicity and recover the dopamine D2 receptor activity (Dunah et al. 2002) but it was also demonstrated in HD models that SP1 suppression can be neuroprotective (Qiu et al. 2006). The role of SP1 in polyQ toxicity is thus not clear. TAFII130 itself also interacts with mutant htt and it was shown to colocalize with atrophin-1 and ataxin-3 (Shimohata et al. 2000; Dunah et al. 2002). Another TAF, TAFII30, was reported to be sequestered in nuclear inclusions formed by mutant ataxin-7 (Yvert et al. 2001).

    Expanded polyQ reduces the cytoplasmic interaction of htt with the repressor element-1 transcription factor/neuron restrictive silencer factor leading to nuclear enrichment of this factor and resulting in transcriptional repression of the gene encoding brain-derived neurotrophic factor (BDNF) (Zuccato et al. 2003).

    Sequestration of p53 into mutant htt aggregates was originally reported to repress the p53-mediated transcription (Steffan et al. 2000). It was later shown that the binding of mutant htt to p53 increased intranuclear levels of p53 and enhanced its transcriptional activity in HD cellular and in in vivo models as well as brain autopsies of HD patients. Genetic deletion of p53 then suppressed mutant htt-induced neurodegeneration in Drosophila (Bae et al. 2005).

    Recently, Yamanaka et al. identified nuclear factor Y (NF-Y) to be associated with htt aggregates in cellular and mouse models of HD. This interaction led to sequestration of NF-Y components resulting in its decreased activity (Yamanaka et al. 2008). NF-Y has been reported to regulate Hsp70 gene transcription (Imbriano et al. 2001). Presence of mutant htt reduced the NF-Y-mediated expression of Hsp70 revealing the mechanism of progressive decrease of Hsp70 in model mice reported previously (Hay et al. 2004; Merienne et al. 2003).

    Furthermore, Doi et al. (2008) found that TLS (translocated in liposarcoma) is a major polyQ aggregate-interacting protein. Mutations in fused in sarcoma/TLS were recently found to cause a familial form of amyotrophic lateral sclerosis (Kwiatkowski et al. 2009), thus TLS-related neuronal degeneration might play a significant role in polyQ diseases.

    In addition to modulating the activities of transcription factors, a direct interaction of mutant htt with DNA was recently shown to perturb gene expression in vitro and in vivo. The enhanced binding of mutant htt to genomic DNA alters the DNA conformation and together with consequent disrupted binding of transcription factors alters the normal mRNA expression (Benn et al. 2008). The relationship between polyQ pathogenesis and transcription factors is however not clear yet and needs further investigation to resolve the above-mentioned conflicting observations.

    In vivo models of polyglutamine diseases

    Appropriate disease model system is an essential tool in understanding the disease pathomechanism and in the development of therapeutic strategies. For example, the formation of neuronal intranuclear aggregates was first observed in a HD transgenic mouse model (Davies et al. 1997) and this discovery led to identification of similar inclusions in brains of patients with polyQ diseases. Gain of function of the polyQ disease proteins enables relatively simple establishments of cellular model systems for these diseases requiring simply (over)expression of the mutant protein. Cellular models are especially useful in high throughput screenings. These include either random chemical libraries or different gene-silencing libraries with small-interfering RNAs (siRNA) or short-hairpin RNAs. Modern imaging equipments, together with different biochemical methods, enable to analyze many altering effects of chemical compounds or gene modifications, on polyQ proteins accumulation and aggregation, cytotoxicity, etc. Also, it is very important to determine the mechanism of action of the potential drugs and the cellular models offer an ideal starting point in this effort.

    After the identification of the molecular genetic background, the modeling of neurodegenerative diseases in transgenic or mutant animal models has provided a great instrument for investigating the disease pathogenesis and testing treatment strategies. The well-characterized genetics, development and anatomy together with short lifespans of invertebrate organisms make them excellent model systems of polyQ diseases. The development and advantages of nematode C. elegans and the fruitfly Drosophila melanogaster models have been reviewed previously (Parker et al. 2004; Cauchi and van den Heuvel 2006; Driscoll and Gerstbrein 2003). The invertebrate models are also suitable for a large-scale screening studies because of relatively low cost, modest time requirements or absence of the complications of the blood–brain barrier (Bates and Hockly 2003; Hughes and Olson 2001). Invertebrate models are especially useful in modifying genetic screens and have revealed several candidate genes having role in protein folding and clearance and affecting the course of SCA1 and HD (Kazemi-Esfarjani and Benzer 2000; Fernandez-Funez et al. 2000).

    Mouse models

    The generation of mammalian models of polyQ diseases is crucial for providing researchers with an alternative approach to study molecular mechanism of respective disorders although some of these models do not completely simulate the neuropathological changes in humans. Generally, the CAG repeat expansions responsible for the diseases in humans are not sufficient to trigger disease in mice. Therefore, very large expansion or multiple gene copies have been used in mice to mimic the humans’ conditions but there is still no model available which would fully reproduce the symptoms seen in patients. Nevertheless, some of the mouse models have been very useful in therapeutic testing. Number of various mouse models for SCA1, 2, 3, 7, DRPLA, SBMA, and HD have been generated and were, together with their potential utilization in therapeutic trials, reviewed recently (Ferrante 2009; Bates and Gonitel 2006; Heng et al. 2008; Yamada et al. 2008; Marsh et al. 2009).

    The transgenic mouse models can be generally divided to fragment and full-length genetic models. Knock-in mice represent the most precise genetic models for polyQ diseases so far. An expanded CAG triplet repeat is inserted into the endogenous mouse gene and the expression is driven by the respective endogenous promotor. Alternatively, a whole mouse exon can be replaced by a human exon carrying expanded polyQ.

    Fragment transgenic Huntington’s disease mouse models

    As mentioned in the section `Mouse models', there is a problem with inverse correlation of genetic accuracy of the model and the phenotype or the extent of neuropathological lesions observed in the animals. A study comparing the knock-in HdhQ150/Q150 mice and N-terminal exon 1 fragment model, R6/2, revealed that both models had similar phenotypic and molecular characteristics, but they appeared much later in the knock-in mice (Woodman et al. 2007). Robust phenotype and pathology display model mice expressing the fragment of the gene containing the expanded polyQ. In case of HD, three mouse models with an N-terminal segment of mutant htt have been used in HD research, including R6/2, R6/1, and N171-82Q. The first and best characterized is the R6/2 line, which has an N-terminal end of exon 1 originally with ∼150Q (Mangiarini et al. 1996). This model revealed that expression of only exon 1 containing expanded polyQ is sufficient to produce mice with many neuropathological features of HD. It is one of the most commonly used genetic models of HD because of a progressive and homogenous phenotype. The short survival of about 3 months and well-quantifiable phenotype makes it a great tool for experimental therapeutic interventions. Moreover, the behavioral impairment and neuropathological findings suggested that the R6/2 model corresponds to human HD to a large extent (Stack et al. 2005). The effects on lifespan extension, body weight, and improvements in motor performance have been used as indicators of the treatment outcome in many studies. Other tests, such as limb clasping score, grip strength, or general activity became also very useful in evaluating and comparing different therapeutic approaches. It has been found that the variability in survival and the amelioration of the phenotype with increased CAG repeat size reduces the utility in therapeutic trials and may corrupt the results (Stack et al. 2005). In our experience, R6/2 mice with repeat size between 125 and 145 CAG were very appropriate for testing different treatments.

    The R6/1 mouse model harbors exon 1 of the human HD gene with ∼116 CAG repeats (Mangiarini et al. 1996). This model has been not studied so extensively as R6/2. The disease starts later and has milder course than that in R6/2 with which it shares the hindlimb clasping, gait abnormalities and other motor deterioration. The later disease onset correlates with delayed appearance of htt aggregates and brain atrophy. Also, R6/1 mice usually live more than 1 year (Naver et al. 2003). In several treatment studies, N171-82Q mice expressing N-terminal fragment of htt gene containing exons 1 and 2 with 82 CAG was used (Schilling et al. 1999a). These mice have similar but less severe and more variable phenotype than R6/2. The lifespan varies between ∼4 and 6 months. Interestingly, the exposure of R6/1and R6/2 mice to enriched environment delayed the disease onset and ameliorated the clasping phenotype and the loss of peristriatal cerebral volume (Hockly et al. 2002; van Dellen et al. 2000; Glass et al. 2004), probably because of restoration of the neurogenic process (Lazic et al. 2006).

    More recently, HD190Q and HD150Q mice were generated expressing exon 1 htt with 190 and 150 CAG under htt promotor, respectively. The transgene also contains enhanced green fluorescent protein, which makes this model a simple and sensitive tool for in vivo testing of therapeutic molecules for inhibiting aggregate formation that can be visualized by a fluorescent imager. Down-regulation of mRNAs for a number of hypothalamic peptides that are known to be involved in feeding behavior and energy homeostasis was observed in these mice. The median survival of HD190Q is ∼21 weeks and that of HD150Q mice 32 weeks (Kotliarova et al. 2005). This model however will need further phenotypic analysis.

    Full-length transgenic Huntington’s disease mouse models

    An HD mouse model with full-length human htt with 48 or 89Q manifested progressive behavioral and motor dysfunction with neuronal loss and gliosis (Reddy et al. 1998). A yeast artificial chromosome HD mouse model expresses the whole human htt with 128Q (YAC128) and replicates the neuropathology in HD patients (Hodgson et al. 1999). First symptoms, such as hindlimb clasping or gait abnormalities occur at about 3 months (Slow et al. 2003). Apoptosis is activated and mitochondrial dysfunction occurs in these mice as well (Fei et al. 2007). YAC128 appears to be a very useful model for therapeutic studies (Tang et al. 2009). A newer model, BAC transgenic mouse model of HD, was generated by introduction of BAC consisting of whole human htt locus of 170 kb with an expansion of 97 CAG (Gray et al. 2008). BAC transgenic mouse model of HD revealed that the progressive course of HD and the selective pathology of mouse brains might occur without early nuclear accumulation of aggregated htt. Synaptic pathology was observed prior to neuronal degeneration in these mice with first signs of synaptic dysfunction detectable at 3 months and a progressive synaptic pathology at 6 months (Spampanato et al. 2008).

    Knock-in Huntington’s disease mouse models

    Four knock-in HD mouse models have been generated and are one of the most precise genetic replicas of conditions in humans and were expected to mimic the pathogenesis of HD better than the transgenic models. The symptoms and neuropathological findings of these mice turned out to be quite subtle with no cerebral atrophy and only rare nuclear inclusions occurring later than in transgenic mice. No shortening of the lifespan was observed in knock-in models (Shelbourne et al. 1999; Wheeler et al. 2000; Lin et al. 2001; Levine et al. 1999), therefore survival is not suitable as an endpoint in therapeutic studies. On the other hand, these mice display measurable neuropathological features and behavioral symptoms that can be used for validation of experimental treatments.

    The HdhQ111 line is a knock-in model with 111 CAG repeats inserted into the mouse HD gene (Wheeler et al. 2000). Nuclear htt immunoreactivity in ventral striatal neurons at about 6 weeks of age and later, 4.5 months, formation of N-terminal htt intranuclear inclusions in medium spiny neurons were observed in these mice. Late-onset motor deficits comprising mild gait disturbances were detected by footprint analysis at 24 months of age accompanied by reactive gliosis in affected areas of the brain. Lack of neuronal apoptosis in the striatum at this age indicates slow progression of the disease (Wheeler et al. 2002).

    Similarly to HdhQ111, in HdhQ94 and HdhQ140 knock-in mice, exon 1 of the mouse htt is replaced by mouse/human chimeric exon 1, in these cases with 94 and 140 CAG repeats, respectively. Mice expressing 94Q protein show increased repetitive rearing at night at 2 months followed by decreased locomotion activity at 4 and 6 months of age. Intranuclear microaggregates were observed at 4 months together with reduced mRNA levels in striatum. The neuronal intranuclear inclusions are widely distributed throughout striatum at 6 months of age (Menalled et al. 2002).

    HdhQ140 mice develop very similar phenotype as HdhQ94, however the disease starts earlier. The increased rearing occurs at 1 month, hypoactivity at 4 months and gait deficits at 12 months of age. Htt inclusions were observed from 1 month of age and there was a more widespread distribution than in HdhQ94 mice with intranuclear aggregates in several brain regions at age of 6 months (Menalled et al. 2003). Neuronal loss with reactive gliosis was found at 20–26 months of age with as much as 38% reduction in striatal volume. Alterations in dopaminergic signaling in HD mice could be at least partly attributed to the decrease in 32 kDa dopamine and cAMP-regulated phosphoprotein expression observed in the striata of HdhQ140 mice from 12 months of age (Heng et al. 2008).

    Another knock-in mouse model is Hdh(CAG)150 with insertion of 150 CAG expansion into exon 1 of the mouse htt homolog. Hdh(CAG)150 mice have late-onset phenotype with neuronal intranuclear inclusions present predominantly in the striatum (Lin et al. 2001; Tallaksen-Greene et al. 2005; Heng et al. 2007). Significant behavioral phenotype does not appear within first year and only after 70–100 weeks of age in homozygous mice, weight loss, decreased activity, diminished rotarod performance, and clasping was observed. At age of 100 weeks, both homozygous and heterozygous Hdh(CAG)150 mice exhibited resting tremor, unsteadiness, and staggering gait. Gliosis was significantly increased by 14 months. By 27 weeks, nuclear htt immunoreactivity and ubiquitin-positive inclusions were initially present within the matrix compartment and later, at 70–100 weeks of age, intranuclear inclusions were observed in most striatal neurons. Striatal dopamine D1 and D2 receptor binding sites were reduced at 100 weeks of age in both homozygous and heterozygous mice with a 50% neuronal loss and a 40% reduction in striatal volume. The D1 and D2 receptors were, however, already diminished at age of 70 weeks with no loss of striatal neurons suggesting that neuronal dysfunction precedes neurodegeneration (Heng et al. 2007).

    Mouse models of other polyglutamine diseases

    Transgenic mouse models bearing either truncated or full-length cDNA of respective genes for SCA1–3 and 7, DRPLA, and SBMA, YAC mouse models for SCA3 and SBMA and knock-in mouse lines for SCA1 and SCA7 have been generated (Yamada et al. 2008; Marsh et al. 2009). Models of all polyQ diseases showed accumulation of mutant proteins in neuronal nuclei except SCA2, where the expanded ataxin-2 was localized in the cytoplasm (Huynh et al. 2000).

    Polyglutamine mouse models and preclinical therapeutic testing

    Recently, concerns about applicability of different mouse models for preclinical therapeutic testing have been raised. Especially, the suitability of the models with truncated forms of mutant polyQ proteins have been discussed because they might represent very artificial systems not reflecting the real conditions observed in human diseases. Usually, these truncated transgenes are expressed from more than one copy of the cDNA per cell resulting in an over-expression and intensified toxicity. However, this should not be a prerequisite for avoiding the fragment models for initial experimental treatments. It has been shown that in the full-length HD mouse models, the cleaved form of mutant htt is the mediator of cytotoxicity (Graham et al. 2006). If a therapy is able to ameliorate the neuropathological and behavioral phenotype of an ‘exaggerated’ state, the more subtle damages observed in full-length or knock-in mouse models would be expected to be manageable with the same therapy and in an optimal situation with more pronounced beneficial outcome or with lower doses of the drug reducing the risk of unwanted side effects. Moreover, some of the fragment models, such as R6/2, are so well characterized and have been employed in many studies that they are the best choice for at least initial in vivo treatment testing and comparing the outcomes with preceding trials. Although this is not true in some therapeutic strategies, such as those targeting the proteolytic cleavage of the intact proteins. For other treatments, such as those targeting polyQ protein aggregation or clearance, the fragments models are very useful because of the robust expression of the transgenes and constant appearance of neuronal inclusions. For an eventual translation of the particular treatment into the clinic, however, the results from a fragment model should be validated in a second mouse model with a full-length transgene or in a knock-in model.

    Therapeutic approaches for polyglutamine diseases

    Experimental treatments of polyQ diseases could be divided into those targeted at the pathological cascades of the disease, such as preventing the cellular damage and to those intercepting downstream deterioration. To the former category belong the therapies decreasing the levels and inhibiting the aggregation of the mutant protein, while the approaches targeting the toxic effects of the polyQ protein such as mitochondrial dysfunction and oxidative stress, transcriptional abnormalities, UPS impairment, excitotoxicity, or apoptotic pathways belong to the latter. Some treatments can target both aspects, for example amiloride or its homolog benzamil decreases the intracellular levels of mutant htt by activating the impaired UPS (Wong et al. 2008), enabling the degradation of other ubiquitinated cellular proteins.

    Despite many experimental therapeutic studies (Beal and Ferrante 2004; Li et al. 2005; Giampa et al. 2009), unfortunately only a few have been translated into clinical trials. This is however not always a result of a failure or toxicity of the therapeutic agent itself in preclinical confirmatory studies. The clinical studies in polyQ diseases patients are difficult to perform because of slow progression of these diseases, low incidence, and inter- and intrafamilial variability in the disease course.

    Therapies aimed at the polyglutamine proteins

    Gene silencing

    Decreasing the levels of the mutant protein and thus preventing the downstream deteriorating effects appears to be one of the best strategies. The therapeutic potential of down-regulating abnormal gene expression has been demonstrated in a tetracycline-regulated mouse model of HD (Yamamoto et al. 2000) and a doxycycline-regulated SCA1 mouse model (Zu et al. 2004). The nuclear inclusions, which formed after induction of the mutant htt expression, disappeared when the expression was shut down. Also the behavioral phenotype was ameliorated, suggesting that therapeutic approaches aimed either at inhibition of mutant htt expression or its degradation might be effective. Several techniques targeting polyQ protein expression have been explored including siRNA or short-hairpin RNA (Bonini and La Spada 2005). This strategy however appeared to be not feasible in humans because of the non-specific ablation of the normal gene copy. On the other hand, this problem could be overcome as single nucleotide polymorphisms have been identified with some of them being suitable to siRNA targeting (Lombardi et al. 2009; Pfister et al. 2009). Recently, microRNAs suppressing the expression of ataxin-1 mRNA by specific binding to its 3′-untranslated region and decreasing the levels of the protein were identified and successfully tested in cells (Lee et al. 2008). For diseases with pathology at limited localizations such as retina in SCA7 this approach is suitable. On the other hand, targeting wide CNS pathology in other polyQ diseases would be more challenging.

    Enhancement of protein degradation

    Enhancing the degradation of mutant proteins is another therapeutic approach and the attempts to increase the autophagic clearance of mutant htt resulted in reduced htt toxicity in N171-82Q transgenic mice and in the Drosophila and the zebrafish HD models (Sarkar et al. 2007b; Williams et al. 2008). The activations of both mammalian target of rapamycin-dependent (e.g. by rapamycin analog CCI-779) or -independent (e.g. by lithium, calpain inhibitors, etc.) pathways modulation were beneficial and the combination treatment resulted in additive protection against polyQ-related neurodegeneration (Sarkar et al. 2008). Lithium also improved the neurological and pathological findings in SCA1(154/2Q) knock-in mice (Watase et al. 2007). There is not much knowledge about regulation of the enzymatic UPS activity. While a relatively large-scale of UPS inhibitors exists, no chemical compound actually activating the UPS activity has been available. Recently, amiloride, a well-known potassium-sparing diuretic drug widely used in clinics, and its derivative, benzamil, have been reported to reduce the polyQ aggregation and toxicity in HD models. Benzamil ameliorated brain pathology, motor deficits, and increased the lifespan of R6/2 mice (Wong et al. 2008). Another compound shown to increase enzymatic UPS activity is Y-27632, a rho-associated kinases inhibitor, which already has been used in clinical trials because of its anti-ischemic, antivasospastic, and antihypertensive effects (Lai and Frishman 2005). Interestingly, this compound also enhanced the macroautophagy activity, and this unique effect of modulating both main cellular degradation pathways led to reduced levels and reduced aggregation of mutant htt, ataxin-3, AR, and atrophin-1 in cell systems (Bauer et al. 2009). Although Y-27632 was previously shown to reduce the polyQ toxicity in Drosophila model of HD (Pollitt et al. 2003), this treatment needs to be tested further in mouse models.

    Inhibition of aggregation

    One of the first therapeutic approaches in polyQ diseases has been aimed at the prevention of aggregation. The therapeutic potential of small molecules able to prevent directly the formation of polyQ aggregates has been shown in several studies. For example treatment of R6/2 mice with Congo Red or trehalose significantly increased the mice survival by 16.4% and 11.3%, respectively (Sanchez et al. 2003; Tanaka et al. 2004). In a recent study, however, chronic administration of Congo Red did not improve the phenotype and lifespan significantly, but this could have been a consequence of limited ability of this compound to cross the blood–brain barrier (Wood et al. 2007). Trehalose has been reported to activate macroautophagy in an mammalian target of rapamycin-independent manner (Sarkar et al. 2007a) so the observed beneficial effect in mice might result from the stabilization of the expanded polyQ protein and its enhanced degradation. Conserving mutant htt in the native non-toxic conformation has been thought to prevent the cytotoxic effect of the mutant protein. PolyQ-binding peptide 1 was shown to prevent conversion of the expanded polyQ into aggregation-prone β-sheet-rich conformation and prevent neurodegeneration in a Drosophila model of HD (Nagai et al. 2003, 2007).

    Cystamine may reduce expanded polyQ aggregation by inhibition of TG that is thought to crosslink expanded polyQ proteins and facilitate their aggregation. In HD brains, increased TG activity was observed (Karpuj et al. 1999). In one study, where the treatment of R6/2 mice started at age of 7 weeks, cystamine did not affect aggregate formation, while improving the mice phenotype (Karpuj et al. 2002). This could have been a result of caspase inhibition and antioxidant activity of cystamine (Lesort et al. 2003). In another study though, where the treatment started at 3 weeks, the htt aggregation decreased by 68% in striatum and by 47% in neocortex (Dedeoglu et al. 2002). The survival improvement was high in both studies with 12% and 19.5% increase, respectively.

    Another strategy to reduce the misfolding, oligomerization, and aggregation of polyQ proteins is to increase the cellular levels of molecular chaperones such as Hsp70. The up-regulation not only interferes with the aggregation process but may also enhance the degradation of the mutant protein through UPS. Beside over-expression of HSPs, several promising compounds have been introduced which induce the expression of HSPs. Geranylgeranylacetone is an acyclic isoprenoid compound, used as an oral anti-ulcer drug, strongly induced HSPs expression in different tissues (Hirakawa et al. 1996). Oral administration of geranylgeranylacetone enhanced the expression of Hsp70, Hsp90, and Hsp105 through induction of heat-shock factor 1 in CNS, where it inhibited nuclear accumulation of mutant AR leading to improved motor behavior and increased survival of transgenic SBMA mice (Katsuno et al. 2005). An Hsp90 inhibitor, 17-(allylamino)-17-demethoxygeldanamycin (17-AAG) has also been demonstrated to ameliorate neurodegeneration in SBMA mice (Waza et al. 2005). Treatment with 17-AAG dissociated p23 (Hsp90 cochaperone) from the Hsp90–AR complex and facilitated the proteasomal degradation of the mutant AR. Recently, it has been reported that 17-AAG suppressed polyQ-induced neurodegeneration in Drosophila models of SCA3 and HD via activation of heat-shock factor 1 which in turn up-regulated molecular chaperones Hsp40, Hsp70, and Hsp90. 17-AAG reduced the lethality of SCA3 and HD Drosophila models by 74.1% and 46.3%, respectively (Friedman et al. 2008). An analog of 17-AAG, 17-(dimethylaminoethylamino)-17-demethoxygeldanamycin has recently been reported to induce expression of Hsp40 and Hsp70, degrade the pathogenic AR protein, and to ameliorate the phenotype in SBMA transgenic mice (Tokui et al. 2009).

    Therapies aimed at downstream effect of the expanded polyglutamine protein

    Besides attempting to cut off the polyQ-related pathology at the level of the mRNA or polyQ protein itself, a number of treatments targeting the downstream pathogenic events have been tested.

    Modification of transcription

    Several compounds modulating the deregulated gene transcription have been tested in different models of polyQ diseases. Especially, targeting histone methylation and acetylation is thought to achieve normalization of transcription disrupted by expanded polyQ proteins. Histone deacetylase inhibitors such as suberoylanilide hydroxamic acid (SAHA) and sodium butyrate (SB) have been demonstrated to ameliorate the phenotype in polyQ disease models. SAHA has been tested in R6/2 mice in which it improved the motor phenotype however no amelioration in body weight was observed and higher doses of SAHA were toxic (Hockly et al. 2003). The effect of SB was investigated in R6/2 mice and besides the improvement in motor phenotype it also delayed the body weight loss and extended the lifespan by more than 20%. Moreover, SB increased not only the acetylation of histones H3 and H4 but also of SP1 resulting in improved transcriptional regulation in R6/2 brains (Ferrante et al. 2003). In DRPLA transgenic mice expressing full-length atrophin-1 with 118Q (Atro-118Q14), SB ameliorated the motor impairments and extended the lifespan (Ying et al. 2006). Treating AR-97Q SBMA mice expressing full-length AR with SB mitigated the neuromuscular phenotype and increased survival (Minamiyama et al. 2004). Another histone deacetylase inhibitor, phenylbutyrate, improved survival, and attenuated gross brain atrophy and ventricular enlargement in N171-82Q HD transgenic mice (Gardian et al. 2005).

    The inhibition of histone H3 methylation with mithramycin resulted in a great extension of lifespan in R6/2 mice accompanied by improvement in motor performance and striatal pathology (Ferrante et al. 2004). The improvement of H3 and H4 modification profile after chromomycin and mithramycin treatment was observed also in N171-82Q mouse model (Stack et al. 2007). Another drug shown to mitigate the transcriptional dysbalance is the phosphodiesterase type IV inhibitor rolipram, which increases the phosphorylation and activity of CREB. Rolipram ameliorated the neuropathological findings, slowed the progression of neurological phenotype and increased the survival in R6/2 mice. Moreover, BDNF, whose expression is impaired in HD (Zuccato et al. 2003), was induced in treated mice through restored function of CREB (DeMarch et al. 2008). In another study, it was shown that rolipram also prevented the sequestration of CBP into nuclear aggregates (Giampa et al. 2009).

    In the AR-97Q SBMA mouse model, a drug specifically disrupting the interaction of mutant AR with CBP, ASC-J9 (dimethylcurcumin), reduced the SBMA symptoms and reversed muscular atrophy in treated mice. This effect was accompanied by restoration of vascular endothelial growth factor 164 expression (Yang et al. 2007).

    In a C. elegans HD transgenic model expressing N-terminal htt with 128Q, treatment with resveratrol resulted in amelioration of mutant polyQ toxicity and it also reduced cell death in neuronal cells derived from HdhQ111 knock-in mice. The underlying mechanism was shown to be an enhanced activation of silent mating type information regulation-2 protein through activation of daf-16, a member of the forkhead box family of Forkhead transcription factors (Parker et al. 2005). Resveratrol treatment appears to be very promising candidate for support therapy of polyQ diseases.

    Mitochondria, energy metabolism, and antioxidants

    Different compounds improving energy metabolism defects or reducing oxidative stress in polyQ diseases have been successfully tested in mouse models. Reduced concentrations of creatine and phosphocreatine were observed in basal ganglia of HD patients (Sanchez-Pernaute et al. 1999). Creatine administration stabilized the MPT, prevented ATP depletion, and increased the protein synthesis. In R6/2 mice, the treatment with 2% creatine ameliorated the brain pathology, improved the phenotype and increased the lifespan by 17.4% (Ferrante et al. 2000). Similarly, in N171-82Q HD mice, the survival extended by 19.3% with 2% dietary creatine treatment (Andreassen et al. 2001a). Other drugs preventing the depletion of ATP, or increasing the activities of mitochondrial complexes II–III seen in HD, such as dichloroacetate and triacetyluridine were successfully tested in R6/2 mice with moderate but significant effects (Andreassen et al. 2001b; Saydoff et al. 2006).

    Antioxidants such as α-lipoic acid, coenzyme Q10, clioquinol, tauroursodeoxycholic acid (also antiapoptotic effect), or BN82451 have been proven effective in R6/2 mouse lines (Giampa et al. 2009). Especially coenzyme Q10, an antioxidant, a cofactor of the mitochondrial electron transport chain and inhibitor of MPT, extended the survival of R6/2 mice by up to 26.3% while greatly improving the motor perfomance and reducing weight loss and nuclear inclusions (Smith et al. 2006). Clioquinol administration reduced the striatal atrophy, decreased nuclear inclusion formation, ameliorated the HD phenotype development, and enhanced the lifespan in R6/2 mice by 20% (Nguyen et al. 2005).

    Excitotoxicity

    The overactivation of NMDA glutamate receptors and subsequent excitotoxicity in striatal neurons leading to their death has been thought to have a role in pathogenesis of HD and electrophysiological studies in brain slices from R6/2 mice indeed suggested complex changes in glutamatergic transmission (Cepeda et al. 2003). Therefore, compounds intercepting excessive glutamate release have been tested for therapeutic purposes. Treatment of R6/2 mice with a glutamate antagonist, riluzole, decreased the body weight loss, affected the initial locomotor activity, significantly increased the mean survival (10.2%), and appeared to modulate the aggregation process (Schiefer et al. 2002). Although the beneficial effects of riluzole were not confirmed in a recent study (Hockly et al. 2006), clinical trials were carried out in HD patients. Unfortunately, the latest 3-year trial showed no beneficial effect (Landwehrmeyer et al. 2007). Two drugs affecting the levels of glutamate in synaptic cleft through stimulation of the pre-synaptic metabotropic glutamate receptor 2 (mGluR2) or inhibition of the post-synaptic mGluR5, LY379268, and 2-methyl-6-(phenylethynyl)-pyridine, respectively were also tested for their ability to combat HD in R6/2 mice (Schiefer et al. 2004). Similar to riluzole, both compounds significantly increased the lifespan of R6/2 mice. In one study, the administration of remacemide (NMDA antagonist) or coenzyme Q10 prolonged the survival of R6/2 mice by 15.5% and 14.5%, respectively, and reduced motor deficits, weight loss, and aggregate formation. When the drugs were combined, the survival extended by 31.8% (Ferrante et al. 2002). This combination was not so effective in N171-82Q mice but the lifespan still improved by more than 20% (Schilling et al. 2001). The clinical trial using either of these drugs or in combination, however displayed no significant effects in HD patients (Huntington Study Group 2001). Another NMDA antagonist, memantine, has been reported to retard the disease progression in HD patients (Beister et al. 2004).

    Apoptosis

    From antiapoptotic drugs, minocycline appeared to be a good candidate for therapeutic trials. Minocycline was shown to inhibit the mitochondrial release of apoptosis inducing factor, proapoptotic protein Smac/Diablo, and cytochrome c, while decreasing the cleavage of proapoptotic factor Bid and activating caspases 1, 3, 8, and 9 (Wang et al. 2003). Intraperitoneal administration of minocycline to R6/2 mice extended the lifespan by 13.5% (Chen et al. 2000). When administered perorally, minocycline failed to improve the symptoms in R6/2 mice (Smith et al. 2003). In another study, minocycline and coenzyme Q10 improved survival of R6/2 mice 11.2% and 14.6%, respectively, and when administered together, the neuropathology and phenotype outcome improved when compared with separate treatments and the lifespan extension was 18.2% (Stack et al. 2006). The results of the clinical studies using minocycline produced promising data (Bonelli et al. 2004) but a long-term clinical trial should be conducted to definitely evaluate the benefits of minocycline in HD patients.

    Inhibition of caspases could comprise double beneficial effect in polyQ pathogenesis. First, decreased generation of caspase-cleaved fragments of mutant proteins may result in reduced toxicity and second, execution of apoptosis may be blocked. A broad caspase inhibitor, z-Val-Ala-Asp-fluoromethylketone, improved the rotarod performance of R6/2 mice and extended the lifespan by 25% (Ona et al. 1999). The combined administration of caspases 1 and 3 inhibitors, Tyr-Val-Ala-Asp-chloromethylketone and Asp-Glu-Val-Asp-aldehyde-fluoromethylketone, respectively, increased the survival by 17.3% (Chen et al. 2000).

    Other treatment possibilities

    Selective serotonin reuptake inhibitors are a group of drugs widely used for the treatment of patients with depression and severe anxiety disorders (Blier and de Montigny 1994). In addition to restoration of serotonin levels in striatum, the treatment of N171-82Q mice with paroxetine reduced brain atrophy, delayed the onset of motor dysfunction, attenuated weight loss, and increased the survival (Duan et al. 2004). Another selective serotonin reuptake inhibitor, sertraline, had similar effects in N171-82Q mice. Upon this treatment, increased BDNF levels, preserved levels of Hsp70 and Bcl-2 in the brains, and enhanced neurogenesis were observed (Duan et al. 2008). The enhanced neurogenesis and increased BDNF levels were also reported in R6/2 mice treated with sertraline (Peng et al. 2008). Importantly, the effective levels of sertraline in mice were comparable to those achievable in humans (Duan et al. 2008).

    Control of Ca2+ homeostasis seems to be a promising therapeutic strategy for polyQ diseases and particularly the modulation of the IP3R1 activity in ER membrane could provide a good target (Bezprozvanny 2007; Chen et al. 2008b). Inhibiting the excessive Ca2+ release from intracellular storages would potentially result in reduced deteriorating events such as calpain activation or mitochondrial dysbalance leading to caspase and apoptosis activation. Although there has been a focus on identifying downstream effectors to address the particular pathogenic event without disturbing other physiological or protective cellular processes, in this case, the relatively proximal component of the intracellular Ca2+-dependent network, down-regulation of IP3R1 activity, could be evaluated more extensively.

    Another strategy to advance toward an effective therapy of polyQ diseases is the combination of drugs targeting different aspects of these disorders. As mentioned in the section `Excitotoxicity', the synchronous utilization of some compounds resulted in increased therapeutic effects in some cases (coenzyme Q10 with remacemide or minocycline). Combinations of a treatment targeting the polyQ protein abnormal processing with another treatment aimed at a downstream pathogenic process are of particular interest. For example, the concomitant treatment of R6/2 mice with cystamine (prevents polyQ aggregation by TG inhibition) and mithramycin (restores the polyQ-mediated transcriptional dysregulation) extended the mean survival by 40% and markedly improved the neuropathological findings more than a single treatment by any of these two drugs (Ryu et al. 2006).

    As described in the section `Role of nuclear localization of mutant polyglutamine proteins', testosterone binds to AR and promotes its uptake by nucleus, the primary site of SBMA pathogenesis (Stenoien et al. 1999; Katsuno et al. 2002). Treatment aimed at testosterone blockage by castration has been shown to be beneficial in SBMA transgenic mice (Katsuno et al. 2002). Chronic adminitration of a luteinizing hormone-releasing hormone analog, leuprorelin, caused decreased serum levels of testosterone in male transgenic AR-97Q mice resulting in amelioration of neuromuscular phenotype and increased lifespan (Katsuno et al. 2003). Importantly, phase 2 clinical trial showed beneficial effects of leuprorelin in SBMA patients and is a good premise for large-scale clinical trials of androgen deprivation (Banno et al. 2009).

    In last decade, a great progress has been achieved in understanding the pathomechanism of polyQ diseases, but, unfortunately, the therapy development does not keep up. The challenge for the next decade will be the translation of promising therapeutic strategies from preclinical trials to clinical use.

    Acknowledgement

    We would like to thank NIH Fellows Editorial Board for corrections in the manuscript.

        The full text of this article hosted at iucr.org is unavailable due to technical difficulties.