U.S. flag

An official website of the United States government

NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.

Lo DC, Hughes RE, editors. Neurobiology of Huntington's Disease: Applications to Drug Discovery. Boca Raton (FL): CRC Press/Taylor & Francis; 2011.

Cover of Neurobiology of Huntington's Disease

Neurobiology of Huntington's Disease: Applications to Drug Discovery.

Show details

Chapter 7Mouse Models for Validating Preclinical Candidates for Huntington’s Disease

and .

INTRODUCTION

Clinical Features of Huntington’s Disease

Ever since its original description by George Huntington in 1872, Huntington’s disease (HD) has been known as one of the most devastating inherited neurodegenerative disorders afflicting the human brain. Currently in the United States, there are about 30,000 patients with HD and another 150,000 people who are at a genetic risk of developing the disease. HD is characterized by the clinical triad of late-onset motor disturbances (i.e., chorea and dystonia), psychiatric deficits (i.e., depression, irritability, and psychosis), and cognitive decline (Bates et al., 2002). The majority of HD patients experience onset of symptoms around the age of 40 (adult-onset HD), and the disease relentlessly progresses until the patient’s death, which usually occurs within 10–20 years after disease onset. A small subset of HD patients experience the onset of symptoms before age 20 (juvenile HD), and these patients exhibit slightly different clinical features in that they tend to have more dystonia than chorea, as well as a higher incidence of epilepsy. Although the onset of HD is currently defined by the onset of motor deficits, recent studies using more sensitive motor studies, as well as cognitive studies, indicate that clinical manifestations of HD may occur years to decades before the onset of motor symptoms, and such deficits may correspond to the early and progressive cortical and striatal atrophy seen in presymptomatic HD patients (Aylward et al., 2004; Rosas et al., 2005, 2006).

HD Neuropathology

HD neuropathology is characterized by selective atrophy and neuronal loss, primarily targeting the striatum and cortex (Vonsattel and DiFiglia, 1998). The neostriatum (i.e., caudate and putamen) bears the brunt of the disease. In the striatum, the atrophy is accompanied by robust and highly selective loss of the striatal medium spiny neurons (MSNs), as well as reactive gliosis and microgliosis. Other striatal neuronal types, such as a variety of interneurons, may exhibit evidence of dysfunction but do not appear to degenerate in HD. Current neuropathological studies have provided very good clinical correlations between striatal atrophy and MSN degeneration with motor and certain psychiatric manifestations of HD.

Besides striatum, the HD cortex also suffers from atrophy and the loss of pyramidal projection neurons in the deep cortical layers (Hedreen et al., 1991; Heinsen et al., 1994). Similar to the striatum, the projection neurons in the cortex are more vulnerable to degeneration than the cortical interneurons. Until recently, the pathogenic significance of the cortex in HD remained poorly defined. Imaging studies in patients who demonstrate regional-selective cortical thinning, semi-independent of striatal atrophy, are significantly correlated with cognitive deficits in HD (Rosas et al., 2005). Moreover, the white matter abnormalities, particularly those connecting the cortical and subcortical structures such as the striatum, are affected early in HD (Rosas et al., 2006). These emerging clinical studies highlight HD as a disease selectively targeting multiple components in the corticostriatal-thalamocortical circuit, a circuit mediating many of the basic functions affected in HD, including motor control, motor and reward learning, and cognition.

HD Molecular Genetics

A triumph in HD research is the identification of the HD mutation in 1993 (The Huntington’s Disease Collaborative Research Group, 1993). HD is caused by a CAG repeat expansion mutation (>36 repeats) in the Huntingtin (HTT) gene, which translates into an expanded polyglutamine (polyQ) repeat. In HD and in eight other neurodegenerative disorders caused by the polyQ repeat expansion, the onset of clinical symptoms is inversely correlated with repeat length (Zoghbi and Orr, 2000). Although the mutant huntingtin (mhtt) protein is widely expressed in the brain and in other tissues outside the nervous system (Schilling et al., 1995; Sharp et al., 1995), the MSNs in the striatum and pyramidal neurons in the cortex (particularly in the deep cortical layers) are most vulnerable to degeneration (Vonsattel and DiFiglia, 1998). The cellular and molecular mechanisms underlying this selective neuropathology in HD remain unknown. Resolving this and other key HD pathogenic mechanisms may provide critical novel insights toward therapeutic development in HD.

Advantages of Using Mouse Models to Study HD

A primary challenge in studying the pathogenesis and treatment of HD is to develop disease models that recapitulate both the genetic and phenotypic aspects of the disease, hence allowing one to go from genetic mutations to understanding disease mechanisms and ultimately to the development of effective therapeutics. Genetic models of HD have been generated using different model organisms to elucidate pathogenic mechanisms (Marsh and Thompson, 2006; Sipione and Cattaneo, 2001). Some of these genetic models (i.e., yeast, Caenorhabditis elegans, and Drosophila) are suitable for high-throughput target identification and/or validation and for drug screening (Hughes and Olson, 2001). It is generally believed, however, that an essential step in HD therapeutic discovery, before the pursuit of a full-scale human clinical trial, is the use of mammalian genetic models of HD for preclinical validation of the therapeutic targets and/or compounds.

The mouse is an ideal mammalian genetic model organism for modeling human neurodegenerative disorders. First, compared with the nonmammalian model organisms, mice have a genetic background more closely related to humans. Mice and humans diverged only about 75 million years ago compared with more than 600 million years of evolutionary divergence between Drosophila and humans. Because of such close evolutionary distance, mouse and human genomes are highly similar. About 90% of human genes have direct murine counterparts, and the overall genomic organization and gene expression are also similar between humans and mice (Mouse Genome Sequencing Consortium, 2002). Such genetic similarity increases the likelihood that the genomic response to the HD disease gene and/or to a therapeutic intervention may be comparable between HD mice and patients. Second, mice possess the basic neural circuits and neuronal cell types that are selectively targeted in HD, as well as a rich repertoire of behavior assays that may pinpoint the dysfunction and degeneration of neurons in the HD-relevant neural circuit (Watase and Zoghbi, 2003). This cellular/circuitry context permits the use of mouse models to study the cellular and molecular mechanisms underlying selective neuronal toxicity in HD, which could not be accomplished using other common nonmammalian models. Third, rich genetic resources and genetic manipulation tools are available in mice for sophisticated analyses of HD pathogenesis and treatment (Capecchi, 2005; Heintz, 2001). Mice are well known for their small size and short generation time (3 months). There are also many well-characterized inbred mouse strains, each with identical genetic background and well-defined characteristics. Furthermore, the mouse is readily amenable to genetic manipulations, including stable introduction of an exogenous gene into the mouse genome (transgenic mice), deletion of an endogenous gene in the mouse genome (knockout mice), and targeted replacement of an endogenous gene with an exogenous gene/sequence (knockin mice). Finally, more advanced mouse genetic strategies also permit conditional expression of a gene of interest (on or off) in specific cell types and/or at defined time points (Yamamoto et al., 2000). Thus, mice provide a powerful mammalian genetic model for the dissection of neural circuitry and molecular mechanisms underlying HD pathogenesis, as well as for validating and testing therapeutics for HD.

MOUSE GENETIC MODELS OF HD FOR PRECLINICAL STUDIES

Since the publication of the first HD mouse model (R6 mice) in 1996 (Mangiarini et al., 1996), the HD research field has produced more than a dozen different genetic mouse models of the disease (Levine et al., 2004; Menalled and Chesselet, 2002). Together, this rich repertoire of HD mouse models has been essential to unraveling the pathogenic mechanisms elicited by mhtt in mammalian organisms (Gusella and Macdonald, 2006; Landles and Bates, 2004; Orr and Zoghbi, 2007). Because recent excellent reviews on the subject can be found elsewhere, we will not attempt to comprehensively survey all the existing HD mouse models generated (Levine et al., 2004; Li and Li, 2006; Menalled, 2005; Menalled and Chesselet, 2002; Van Raamsdonk et al., 2007). Instead, we will focus on the few preclinical HD mouse models that are already being used or are intended for use in preclinical studies, that is, target validation and therapeutic compound testing (Table 7.1). We will address the genetic and phenotypic characteristics of each model, as well as important issues related to the use of these models for validating therapeutic targets in HD. The general principles discussed in this chapter will apply not only to the few HD mouse models discussed here but also to other future mouse models that may be used for HD preclinical studies.

TABLE 7.1. Summary of Preclinical Mouse Models of Huntington's Disease.

TABLE 7.1

Summary of Preclinical Mouse Models of Huntington's Disease.

Validating Preclinical Mouse Models of HD

The rationale for generating mouse genetic models of HD is based on the belief that these models can recapitulate, at least in part, some of the key disease phenotypes and/or pathogenic mechanisms that occur in patients; therefore, positive preclinical studies in these models may predict positive outcomes in clinical trials. In generating and analyzing these models, a key question that needs to be addressed is how valid these models are in recapitulating the salient genetic and clinical features of the human disease (Watase and Zoghbi, 2003). In this section, we will discuss some of the factors one should consider to determine the validity of a mouse genetic model of HD.

Construct Validity

Construct validity refers to the similarity of the genetic context used to develop the mouse model to that found in HD patients (Fleming et al., 2005). The genetic context should be viewed at multiple levels. At the DNA level, one should consider whether the types of mutations introduced (i.e., CAG repeat length) resemble those in the patients. In this regard, the targeted insertion of expanded CAG repeat and/or mhtt exon 1 into the endogenous murine huntingtin disease homolog (Hdh) locus in the mouse genome (i.e., knockin models) is the most genetically precise model of HD (Gusella and Macdonald, 2006). In the transgenic models of HD (with random insertion of transgenes into the mouse genome), one also needs to consider whether the genomic DNA context of the mutant transgene is similar to that in the patients. At the RNA level, one should consider whether the regulation of mhtt transcription is similar to the endogenous human htt, whether it is driven by an exogenous promoter, and whether endogenous mhtt mRNA processing such as splicing is preserved (i.e., models with an intact htt genomic locus). Finally, at the protein level, one needs to consider how similar the expressed mhtt protein in a given mouse model is to that in patients. Any divergence from the full-length human mhtt protein sequence may also reduce the construct validity of a model. For example, mouse models only expressing mhtt N-terminal fragments may have reduced construct validity at the protein level. Moreover, because the murine Hdh differs from its human counterpart in the polyproline region (polyP, a known protein–protein interaction domain of htt) and in another 273 amino acids outside the polyQ and polyP regions, these amino acid differences may subtly modify the disease processes mediated by the mutant Hdh compared with those by human mhtt.

Face Validity

Face validity refers to whether the overall phenotypes of a disease mouse model, both behavioral deficits and neuropathology, faithfully recapitulate those in HD patients. As described in more detail below, the phenotypes of the current preclinical HD mouse model can be studied at multiple levels to fully uncover its disease-relevant phenotypes. Commonly used behavioral phenotyping tools in HD mice include longitudinal behavioral observations (i.e., SHIRPA) (Rogers et al., 1997); motor behaviors such as rotarod performance (the ability to run a rotating wheel) and open field activity (automated measurement of spontaneous activity in an open field); cognitive tests that include a variety of learning tasks that may depend on the hippocampus or the corticostriatal circuit, such as swimming T mazes (Holmes et al., 2002) and instrumental conditioning (Balleine, 2005); and psychiatrically related behaviors such as anxiety and depressive behaviors and prepulse inhibition (PPI) (Arguello and Gogos, 2006; Holmes et al., 2002). Commonly used neuropathological studies in HD mice include stereological measurement of striatal and cortical volume and stereological counting of the striatal neurons (Menalled et al., 2002), counting striatal degenerating dark neurons as measured by toluidine blue staining of semithin brain sections (Gray et al., 2008; Hodgson et al., 1999; Turmaine et al., 2000), gliosis (Gu et al., 2005), and microgliosis (Simmons et al., 2007). In addition, weight loss and premature death are also used as secondary readouts in several preclinical mouse models. Each mouse model of HD, before entering the preclinical studies, has been extensively characterized to determine the extent of its phenotypic similarities to HD. Those models with relatively early-onset, progressive, and robust disease-like phenotypes (Table 7.1) are now being pursued for preclinical studies in HD.

Predictive Validity

An ultimate test for the validity of an HD mouse model is its predictive validity, which refers to whether key mechanistic findings and therapeutic efficacy in a given HD mouse model can predict positive clinical outcomes in human patients and vice versa. Currently, there is no effective disease-modifying therapy available for HD that can be used to validate the existing HD mouse models. Conversely, because the mouse genetic models of HD are widely used to identify pathogenic mechanisms and to test preclinical candidates, future clinical studies in HD patients using therapeutic leads originating from the mouse studies may be crucial to reveal the predictive validity of these models.

The concepts of construct validity, face validity, and predictive validity will be crucial to guide us in appreciating the strengths but also recognizing the weaknesses of the existing preclinical mouse models of HD. They may also be helpful in choosing the appropriate mouse models for preclinical studies. In the following section, we will describe the three major types of HD mouse models that are currently used or being prepared for use in HD preclinical studies. They include those expressing a small toxic N-terminal fragment of mhtt, those with precise genetic modification of the endogenous murine Hdh, and those human genomic locus transgenic mice expressing full-length human mhtt under its endogenous regulatory elements (Table 7.1). We will describe for each model its genetic construct and key phenotypic outcomes that may be used in preclinical studies.

N-Terminal Mutant Huntingtin Fragment Models

R6/2 Mice

Genetic Construct and Expression

The R6/2 model is the first and one of the most influential HD mouse models for preclinical studies to date (Mangiarini et al., 1996) (Table 7.1). It was created on the (C57BL/6J × CBA) F1 mixed genetic background. R6/2 mice carry about 1 kilobase (kb) of the human huntingtin promoter, which drives the expression of mhtt exon 1 with ~150 CAG repeats. The transgene also contains ~262 base pair (bp) of human htt intron 1 sequence after the exon 1. R6/2 mice express the mhtt-exon 1 transgene at about 75% of the endogenous Hdh level (Mangiarini et al., 1996). One issue with R6/2 is its genetic heterogeneity as a result of the CAG repeat instability in the germ line, which can range from low 100s to >300 repeats. Such dramatic repeat alterations may significantly alter R6/2 mouse phenotypes. Therefore, in a preclinical study, the repeat length for the R6/2 mice should be closely monitored to avoid genetic drift of the repeat lengths during the study.

Behavioral Phenotypes

The earliest motor deficits in R6/2 mice were detected at 4.5 weeks and include decreased running wheel activities in mice individually housed, climbing a wired cylinder, and hypoactivity in the open field test (Hickey et al., 2005). At 7–8 weeks, grip strength and rotarod deficits can be readily detected, and these deficits increase until the animal’s premature death around 16 weeks of age (Carter et al., 1999; Lione et al., 1999; Murphy et al., 2000). In addition, R6/2 mice also exhibit a variety of other motor phenotypes, including changes in circadian rhythms (Morton et al., 2005), clasping (a dystonic posture when the mice were suspended by the tail), tremor, involuntary jerky movements, and seizures (Li et al., 2005; Mangiarini et al., 1996).

Before the onset of overt motor deficits (at 8 weeks of age), R6/2 mice already exhibit a variety of cognitive deficits reminiscent of HD patients. For example, HD patients exhibit perseverance of learned tasks indicating frontal-striatal inhibition deficits (Aron et al, 2003). R6/2 mice also demonstrate difficulty in reversing prelearned tasks in the two-choice swim tank, T maze (Lione et al., 1999), and an automated video-based reversal task (Morton et al., 2006). R6/2 mice also exhibit deficits in spatial learning in the Morris water-maze task (Murphy et al., 2000) and in contextual fear conditioning (Bolivar et al., 2003). In the psychiatrically related deficits, R6/2 mice exhibit impairment in PPI, a sensorimotor gating abnormality also seen in the HD patients (Swerdlow et al., 1995) and in patients with schizophrenia (Swerdlow et al., 1994).

Neuropathology

R6/2 mice demonstrate robust brain atrophy (about 20% brain weight loss at 12 weeks). Striatal atrophy in these mice can be readily measured using unbiased stereology (Li et al., 2005). R6/2 mice exhibited only moderate dark neuron degeneration in the late stage (Mangiarini et al., 1996; Turmaine et al., 2000), but the striatal neuronal atrophy can be readily quantified as a preclinical outcome. Another frequently used pathological endpoint in R6/2 mice is the presence of mhtt-containing nuclear inclusions (NIs) and neuropil aggregates (NAs), some of which can be detected with antibodies against ubiquitin or against the N-terminal region of huntingtin, for example, EM48 antibody (Davies et al., 1997; Gutekunst et al., 1999; Li et al., 1999). The aggregates in this model are much more widespread than those observed in patients using the same EM48 antibody (DiFiglia et al., 1997; Gutekunst et al., 1999). Moreover, because the significance of mhtt aggregates to the pathogenesis of HD neuronal dysfunction and degeneration remains unclear (Bates, 2003), and there is accumulating evidence suggesting that the large nuclear inclusions may be neuroprotective (Arrasate et al., 2004), mhtt aggregates should not be used as the sole pathological outcome in a preclinical study.

Survival and Weight Loss

Two commonly used preclinical outcomes in R6/2 mice are weight loss and premature death. Because these phenotypes can be readily modified by environmental enrichment (Hockly et al., 2002), which may confound any therapeutic study, it is now recommended that standardized living conditions with a moderate level of enriched environment should be implemented for this model (Hockly et al., 2003). Because the underlying etiology for weight loss and death in R6/2 mice remains unclear, these phenotypes are often only used as surrogate markers to support findings from the primary behavioral and neuropathological outcomes (Bates and Hockly, 2003; Hockly et al., 2003).

Other Phenotypes

R6/2 mice also have a set of molecular readouts, such as gene expression changes, that could be readily used in preclinical studies. A recent microarray analysis comparing gene expression changes in HD mice and HD patients revealed that many of the gene expression changes seen in the R6/2 striatum significantly overlap with those occurring in patients (Kuhn et al., 2007). This result supports the idea of using gene expression signatures as a biomarker for preclinical efficacy in R6/2 mice. Other promising biomarkers in R6/2 mice include using NMR spectroscopy to detect a robust (53%) nonlinear decrease in in vivo N-acetyl aspartate (NAA) levels beginning at 6 weeks of age, a brain energetic deficit also seen in patients (Jenkins et al., 2000). Finally, 8-hydroxy-2-deoxyguanosine (8-OHDG), a biomarker for oxidized DNA damage, which is elevated in HD patient serum (Hersch et al., 2006), is also elevated in R6/2 and R6/1 mice (a model with similar construct to R6/2) (Bogdanov et al., 2001; Kovtun et al., 2007). These latter biomarkers (NAA and 8-OHDG), once validated in HD patients, may be used as relatively noninvasive and unbiased endpoints for preclinical and clinical studies in HD (Hersch et al., 2006).

N171-82Q Mice

Genetic Construct

Another commonly used N-terminal fragment mouse model of HD in preclinical studies is the N171-82Q mouse model. N171-82Q mice express an N-terminal fragment of human htt with 82 polyglutamine repeats and the first 171 amino acids of htt in all the neurons of the brain driven by a mouse prion protein promoter (Schilling et al., 1999). The N171-82Q line was created on the C3H/HeJ × C57BL/6J background. This model expresses mhtt fragment at about 20% of levels of the endogenous murine Hdh protein (Schilling et al., 1999).

Behavioral Phenotypes

N171-82Q mice exhibit progressive motor deficits starting at 12 weeks of age, including tremors, uncoordination, hypoactivity, abnormal gait, and clasping (Schilling et al., 1999). These symptoms are progressive until premature death at approximately

5–6 months of age. The most used quantifiable motor phenotype in this model is rotarod impairment, beginning at approximately 11 weeks of age (Andreassen et al., 2001; Schilling et al., 2004b). The behavioral deficits in N171-82Q mice appear to be

more variable compared with R6/2 mice; thus, to attain adequate statistical power, a typical preclinical study with N171-82Q mice would require a large number of mice (i.e., minimum of 20) than a study with R6/2 (Hersch and Ferrante, 2004).

Neuropathology

N171-82Q mice have several pathological features mimicking HD. First, they exhibit brain weight loss at 120 days of age (Andreassen et al., 2001; Hersch and Ferrante, 2004), striatal atrophy (as measured by stereological measurement of the ventricular volume), and striatal neuronal atrophy (Gardian et al., 2005). Moreover, these mice exhibit other neuropathological features more similar to HD patients, including more mhtt aggregates in the cortex than in the striatum, robust reactive gliosis at 4–5 months of age, and apoptotic neuronal degeneration can be detected in the cortex and striatum at 4.5 months of age (Yu et al., 2003).

Survival and Weight Loss

Similar to R6/2 mice, N171-82Q mice also exhibit progressive weight loss (beginning at approximately 90 days of age) and shortened lifespan (average survival of about

130 days). These phenotypes are frequently used as surrogate markers of therapeutic efficacy in the preclinical studies. Because environmental enrichment can also significantly modify these phenotypes (Schilling et al., 2004a), standardized housing conditions are critical in preclinical trials using this model.

Other Phenotypes

N171-82Q mice exhibit hyperglycemia (i.e., increased baseline fasting glucose level and impaired glucose tolerance test) at about 80 days of age, which could be used as an outcome measure (Andreassen et al., 2001). The utility of the test is relatively limited because of its unclear etiology and high variability.

Full-Length Murine Huntington’s Disease Homolog KI Mouse Models

Knockin (KI) models of HD were generated by targeting an expanded polyglutamine repeat and/or adjacent human mhtt exon 1 sequences (including the polyP region) to replace the corresponding sequences in the endogenous murine Hdh; hence the mutant Hdh is expressed from the endogenous Hdh locus in a manner similar to the expression of mhtt in patients. Thus, Hdh-KI mice are generally considered the most precise genetic mouse model of HD (Gusella and Macdonald, 2006). Existing Hdh-KI mice carry expanded mutant CAG repeats up to 150 (Menalled, 2005). Compared with the mhtt N-terminal fragment models, Hdh-KI mice exhibit slow progression and relatively mild phenotypes, and their lifespans are usually normal. Three Hdh-KI models—HdhQ111, CAG140, and Hdh(CAG)150—are currently being developed for preclinical studies and will be discussed in more detail here.

HdhQ111 Mice

Genetic Construct

The HdhQ111 model is a KI mouse model of HD generated on a CD1 × 129Sv background. This model targeted a chimeric murine Hdh/human mhtt exon 1 into the endogenous Hdh locus, and the human mhtt portion includes 111 CAG repeat and human polyP region (Wheeler et al., 1999).

Behavioral Phenotypes

The behavioral phenotypes of HdhQ111 are very mild. Gait abnormalities are detected at 24 months of age, and no rotarod, clasping, or open field abnormalities were detected in heterozygous or homozygous HdhQ111 mice up to 17 months of age (Wheeler et al., 2002).

Neuropathology

HdhQ111 homozygous mice display selective accumulation of nuclear mhtt at 2.5 months of age, as stained by EM48 antibody (Wheeler et al., 2000). This phenotype progresses as EM48-positive puncta are detected at 5 months of age. Nuclear inclusion formation is first detected at 10 months of age (Wheeler et al., 2000, 2002). At 24 months, these mice also demonstrate reactive gliosis in the striatum and a small number (about 3.5%) of striatal dark neurons as stained by toluidine blue (Wheeler et al., 2002). Brain atrophy has not been reported in this model.

Other Phenotypes

HdhQ111 mice exhibit a variety of molecular phenotypes, including but not limited to increased expression of a ribosomal signaling protein Rrs1, somatic Hdh CAG- repeat instability in the striatum, and low mitochondrial ATP levels (Fossale et al., 2002; Lloret et al., 2006; Seong et al., 2005; Wheeler et al., 2003).

CAG140 Mice

Genetic Construct

A second preclinical Hdh-KI mouse model has targeted replacement of the endogenous murine Hdh exon 1 with a chimeric mouse and human exon 1 with 140 CAG repeats (Menalled et al., 2003). They were generated on a 129Sv × C57BL/6J background.

Behavioral Phenotypes

CAG140 mice exhibit early hyperactivity as measured by an increase in rearing at 1 month of age, followed by hypoactivity at 4 months of age (Menalled et al., 2003). This pattern of locomotor activity changes is similar but with an earlier onset compared with another Hdh-KI model, CAG94 (Menalled et al., 2002). CAG140 mice exhibit gait abnormality at 12 months of age, with a decrease in stride length (Menalled et al., 2003).

Neuropathology

CAG140 mice exhibit selective striatal and cortical nuclear staining of mhtt microaggregates by EM48-positive antibody (Menalled et al., 2003). Nuclear inclusions are present in the striatum at 4 months of age and do not appear in the cortex until 6 months of age. At this age, nuclear staining and nuclear microaggregates are also seen in the cerebellum, an area relatively spared in HD. Neuropil microaggregates are also present in the striatum, globus pallidus, and piriform cortex at 2 months of age and increase with age. To date, no overt cell loss or brain atrophy has been reported for this model. However, CAG94, which is a model very similar to CAG140, exhibits significant striatal atrophy (15%) but no neuronal loss at 18–26 months of age as measured by unbiased stereology (Menalled et al., 2002). Therefore, it is very likely that CAG140 may exhibit significant striatal atrophy phenotype that could facilitate its utility in a preclinical study.

Hdh(CAG)150 Mice

Genetic Construct

The third preclinical Hdh-KI model is Hdh(CAG)150, which replaced the short CAG repeat in the murine Hdh exon 1 with a repeat of 150 CAG (Lin et al., 2001). These mice, created in a 129/Ola × C57BL/6J background, were analyzed either in this mixed genetic background (Heng et al., 2007) or in a (CBA × C57BL/6) F1 background generated from C57BL/6 and CBA congenic lines (i.e., backcrossed for >12 generations) (Woodman et al., 2007). The latter breeding strategy ensures the homogeneity of the genetic background in the mutant and control mice (Silva, 1997).

Behavioral Phenotypes

Homozygous Hdh(CAG)150 mice were extensively studied using a battery of behavioral paradigms that revealed several slowly progressive motor abnormalities (Heng et al., 2007; Woodman et al., 2007). In the mixed genetic background, a mild exploratory deficit is present in HdhCAG(150) mice at 70 weeks of age, with more robust deficits seen at 100 weeks (Heng et al., 2007). In the rotarod test, homozygous mice display a significant deficit on the rotarod at 40 weeks (Lin et al., 2001) and 100 weeks of age (Heng et al., 2007). At the latter time point, these mice also exhibit gait abnormalities, difficulty in traversing balance beams, and clasping.

In the (CBA × C57BL/6) F1 background (Woodman et al., 2007), homozygous HdhCAG(150) mice also have onset of significant and progressive rotarod deficits at 18 months of age. A grip-strength deficit was observed at 6 months of age, but it is nonprogressive until 10 months of age. No locomotor deficits were detected in the mutant mice up to 18 months of age, suggesting this phenotype may depend heavily on the genetic background of the mutant mice.

Neuropathology

Mutant huntingtin aggregates can be detected in the striatum and hippocampus in the homozygous HdhCAG(150) mice at 6 months of age and become widespread in the brain at 10 months of age (Tallaksen-Greene et al., 2005). Similar to what was observed in the patients, the recent study reveal that at 100 weeks, HdhCAG(150) homozygous mice exhibit reduced striatal D1 and D2 dopamine receptor ligand binding and reduced dopamine transporter (Dat) ligand binding (Heng et al., 2007). Using stereological counting of NeuN(+) striatal neurons, the same group found that the HdhCAG(150) mice exhibit 40% reduction in striatal volume and 43% reduction in striatal NeuN(+) neurons. This latter stereological result is very promising to use striatal atrophy and/or neuronal loss as preclinical outcomes in this model. However, because these results were obtained from a relatively small sample size, it would be important for the study to be replicated using a much larger sample size, and preferably also in recombinant inbred F1 background.

Survival and Body Weight

HdhCAG(150) mice do not have a shortened lifespan. However, in both mixed and F1 genetic backgrounds, the mutant mice exhibit significant and progressive weight loss at 70 weeks of age (Heng et al., 2007) and at 52 weeks in the pure F1 background (Woodman et al., 2007).

Molecular Phenotypes

Brain gene expression profiling in the HdhCAG(150) F1 mice revealed parallel changes in 22-month-old HdhCAG(150) mice and 12-week-old R6/2 mice (Woodman et al., 2007). Decreases in the expression of chaperones (i.e., heat shock protein 70, heat shock cognate 70) can be confirmed by Western blots in both models. Such parallels in the molecular phenotypes between the two different types of HD mouse models support the further exploration of gene expression changes as biomarkers in preclinical studies.

Full-Length Human Huntingtin Transgenic Mouse Models

Multiple full-length human huntingtin transgenic mouse models (fl-mhtt Tg) have been generated to study the dominant toxicity elicited by expanded polyglutamine repeat in the context of full-length mhtt (fl-mhtt). This group of fl-mhtt transgenic models include one cDNA transgenic mouse driven by the CMV promoter (Reddy et al., 1998) and two genomic transgenic models expressing fl-mhtt from the human genomic locus on a yeast artificial chromosome (YAC) (Hodgson et al., 1999; Slow et al., 2003; Van Raamsdonk et al., 2007) or on a bacterial artificial chromosome (BAC) (Gray et al., 2008). As a group, the fl-mhtt Tg mouse models demonstrate earlier and relatively robust motor deficits (i.e., rotarod deficits), as well as selective atrophy and/or neurodegeneration in the striatum and cortex. Therefore, fl-mhtt Tg mouse models, particularly YAC and BAC models, are currently being used or developed as preclinical mouse models in HD. In this review, we will focus our discussion on the two latter models, YAC128 and BACHD.

YAC128 Mice

Genetic Construct

A series of YAC transgenic models of HD expressing full-length human mhtt with 18, 46, 72, and 128 CAG repeats (i.e., YAC18, YAC46, YAC72, and YAC128) were generated in the inbred FvB background (Hodgson et al., 1999; Slow et al., 2003; Van Raamsdonk et al., 2007). In these models, the YAC transgenes carry the entire 170-kb genomic locus of human htt gene plus 25 kb of the 5′ flanking sequence and about 120 kb of 3′ flanking sequence. YAC128 mice are the latest of the series and exhibit by far the most robust phenotypes among all the YAC models; therefore, YAC128 is used as a preclinical model in HD (Slow et al., 2003). At the protein level, the YAC128 line expresses human mhtt at about 75% of the level of the endogenous murine Hdh (Slow et al., 2003).

Behavioral Deficits

YAC128 mice exhibit hyperactivity at 2 months of age and hypoactivity at 8–12 months of age. They also exhibit rotarod deficit initially at 4 months of age but become more prominent at 6 months of age (Graham et al., 2006; Van Raamsdonk et al., 2005c). YAC128 mice demonstrate a variety of cognitive deficits, including motor learning deficits on rotarod beginning at 2 months, a reversal learning deficit in a swimming T maze beginning at 2 months, PPI deficits at 12 months, openfield habituation test at 9 months, and a linear swimming test at 8 months (Van Raamsdonk et al., 2005e).

Neuropathology

The earliest pathological marker in YAC128 mice is the selective nuclear localization of EM48-positive mhtt in the striatum at 1–2 months of age, which intensifies at 3 months of age (Van Raamsdonk et al., 2005a). EM48-positive mhtt nuclear accumulation is present in the cortex and hippocampus at 3 months of age, and EM48-positive nuclear inclusions appear in the striatum by 18 months of age (Van Raamsdonk et al., 2005a). The YAC model is the first HD model to demonstrate selective atrophy in the striatum and cortex but not in the cerebellum, a pattern reminiscent of that in HD (Van Raamsdonk et al., 2005a). Stereological studies reveal striatal volume loss first detected at 9 months of age, as well as a significant decrease in striatal volume (10%), cortical volume (8.6%), and globus pallidus (10%) at 12 months of age but no change in the hippocampus or cerebellar volumes (Van Raamsdonk et al., 2005a). The striatal volume loss is associated with a loss of NeuN positive neurons (18%) (Van Raamsdonk et al., 2005a).

Survival and Body Weight

The lifespan of YAC128 mice is slightly decreased. YAC128 mice gain weight between 2 and 6 months of age (about 27%), and the weight gain occurs in all organs except the brain and testis, where mhtt is particularly toxic and causes weight loss (Van Raamsdonk et al., 2006). The weight gain phenotype appears to be caused by the overexpression of human htt because YAC18 mice overexpressing wild-type htt also exhibit similar weight gain phenotype. Because HD patients exhibit weight loss rather than weight gain, the weight gain phenotype in the YAC128 model is not a feature of the disease and should not be used as a preclinical endpoint in this model.

BACHD Mice

Genetic Construct

BACHD is a novel transgenic mouse model of HD generated and maintained in the FvB inbred background. BACHD mice express full-length human mhtt from its own regulatory elements on a 240-kb BAC, which contains the intact 170-kb human htt locus plus about 20 kb of 5′ flanking genomic sequence and 50 kb of 3′ sequence (Gray et al., 2008). The BAC was engineered to include an mhtt exon 1 containing a mixed CAA/CAG repeat encoding an intact polyglutamine stretch (Kazantsev et al., 1999; Yang et al., 1997). A recent more precise sizing of the BACHD CAA/CAG repeat by Laragen (Los Angeles, CA), using both direct sequencing and GeneMapper methods, shows that the BACHD CAA/CAG repeat length is 97. Unlike Hdh-KI and R6/2 mice, the CAA/CAG repeat length in BACHD mice appears stable in the germline over many generations. Another feature of the BACHD transgene design is the inclusion of two LoxP sites flanking mhtt-exon 1, which permits the selective removal of mhtt expression in any cell types expressing the Cre recombinase (Branda and Dymecki, 2004; M. Gray and X. W. Yang, unpublished data) Therefore, this model is particularly useful to study the cell autonomous toxicity and pathological cell–cell interactions elicited by mhtt (Gu et al., 2005). BACHD mice have five copies of the transgene integrated and express fl-mhtt protein at about 1.5- to 2-fold of the endogenous Hdh level. The expression of mhtt is in an endogenous pattern and functional because the BACHD transgene can rescue the embryonic lethality of the murine Hdh knockout mice (Gray et al., 2008; Zeitlin et al., 1995).

Behavioral Phenotypes

BACHD mice exhibit mild but significant rotarod deficits at 2 months, but repeated testing revealed that the rotarod deficits in these mice are progressive and become very pronounced at 6 and 12 months of age. At 6 months, BACHD mice also exhibit hypoactivity in the open field test. In the cognitive domain, a preliminary study reveals BACHD mice exhibit deficits in an instrumental reward learning paradigm, in which these mice are unable to efficiently learn the relationships between stimulus, action, and the rewarding outcomes (B. Balleine, unpublished data). The BACHD mice also exhibit significant impairment in acquiring the swimming T-maze task at 4–6 months of age (L. Menalled and D. Howland, personal communication). Finally, BACHD mice also exhibit several phenotypes in the psychiatric-like behavioral domain (Holmes et al., 2002) starting at about 6 months of age, including enhanced anxiety in the light-dark box and elevated plus maze, enhanced depressive-like behavior in forced swim test, and altered PPI (Menalled et al., 2009; M. Gray and X. W. Yang, unpublished data).

Neuropathology

At 6 months of age, BACHD mouse brains are indistinguishable from the wildtype controls and have comparable cortical and striatal volumes. At 12 months of age, BACHD brains are visibly atrophic with 20% reduction in the forebrain weight but normal cerebellar weight compared with the wild-type mice (Gray et al., 2008). Stereological measurement reveals a 28% reduction in the striatal volume and 32% reduction in the cortical volume compared with the wild types. Furthermore, using toluidine blue staining of striatal semithin sections, we found the 12-month-old BACHD mice contain about 15% dark degenerating neurons in the lateral striatum compared with only 0.3% in the control littermates. One distinguishing feature of BACHD mice from the other full-length huntingtin mouse models is the lack of early nuclear localization of EM48-positive mhtt in the striatum and cortex. Instead, at 12 months of age, EM48 staining reveals large and predominantly neuropil mhtt aggregates in the deep cortical layers and very few small aggregates in the striatum. This distribution pattern of mhtt aggregates is reminiscent of that in adult-onset HD (DiFiglia et al., 1997; Gutekunst et al., 1999), albeit the abundance of the aggregates in BACHD mice appears to be less than that in the patients.

Survival and Body Weight

BACHD mice have a normal lifespan compared with their wild-type littermates. Similar to YAC128 mice, BACHD mice also exhibit significant weight gain between 2 and 6 months and maintain the weight differential without further weight gain until 12 months. As discussed above, this phenotype may be related to the overexpression of human huntingtin. Importantly, rotarod performance and body weight at 6 months demonstrate that body weight in BACHD mice is not correlated with their poor rotarod performance; furthermore, a subset of BACHD mice within the normal weight range still exhibit robust rotarod deficits (Gray et al., 2008; Menalled et al., 2009). These results suggest that the neuronal dysfunction and degeneration in this model are related to the toxicity of mhtt in the brain and are unrelated to the body weight changes.

Summary of Preclinical HD Mouse Models

In this section, we have described the three major types of preclinical HD mouse models. These models are used in various preclinical studies because of their particular strengths in genetic and/or phenotypic characteristics. The mhtt fragment transgenic models exhibit early-onset and rapidly progressing behavioral and neuropathological phenotypes associated with significant weight loss and premature death. Furthermore, these models also exhibit distinct molecular changes (i.e., gene expression changes and 8-OHDG level) that are also replicated in the patients. The advantage of the full-length Hdh-KI and mhtt transgenic models is that they may better resemble the pathogenic mechanisms that occur in the patients (Gusella and MacDonald, 2006; Van Raamsdonk et al., 2007). The shared disadvantage of these models, compared with the fragment models, is that they are slowly progressive models, and quantifiable motor and pathological deficits usually do not occur until 6–12 months of age. Hdk-KI mice are the most precise genetic model of HD and hence are valuable to study molecular mechanisms and therapeutic interventions requiring such precise level of mutant Hdh expression in the endogenous genomic context. The motor and pathological phenotypes of the Hdh-KI mice are slowly progressing and mild, but recent studies of CAG140 (Menalled et al., 2003) and Hdh(CAG)150 (Heng et al., 2007; Woodman et al., 2007) mice demonstrate that these models may exhibit motor and striatal atrophy phenotypes that are relatively robust and could be used as preclinical outcomes. Finally, the human fl-mhtt Tg mice, BACHD and YAC128, both exhibit progressive rotarod and open field deficits, late onset, and selective neuropathology reminiscent of HD. In these latter two models, rotarod deficits and striatal or cortical atrophy may be used as outcome measures in a preclinical study. In the next section, we will discuss how these HD mouse models can be used for preclinical studies in HD.

PRECLINICAL TRIALS WITH MOUSE MODELS OF HD

From a genetic perspective, HD is a relatively simple single gene disorder with full penetrance; hence the study of HD from the very beginning has been heralded as a model genetic brain disorder to uncover a rational route from etiology to treatment (Wexler et al., 1991). Our current understanding of HD pathogenic mechanisms remains incomplete, and a large number of pathogenic mechanisms are implicated in HD pathogenesis. These include proteolysis of mhtt to generate toxic mhtt fragments, transcriptional dysregulation, impairment of the proteasome, mitochondrial dysfunction and energetic deficit, excitotoxicity, deficits in vesicular trafficking/transport, and finally, mhtt toxic conformation change (i.e., aggregation) (Gatchel and Zoghbi, 2005; Li and Li, 2006). In addition to targeting to these specific mechanisms, many HD neuroprotective therapeutics are designed to target broadly the processes of cell death (i.e., apoptosis) or increase cellular “healthiness” (Beal and Ferrante, 2004). Based on the broad set of potential HD pathogenic mechanisms and even broader set of therapeutic strategies, a large number of preclinical candidates are being developed in the HD therapeutic pipeline, particularly from the high-throughput screening using cellular and small organism models (i.e., C. elegans or Drosophila models) (Hughes and Olson, 2001). Because only a small subset of these candidates can possibly be moved into human clinical trials, preclinical studies in the HD mouse models will play a central role in prioritizing the lead candidates for the very expensive next stage of drug discovery (i.e., chemical lead optimization, toxicity studies, and clinical trials). In the remainder of this chapter, we will focus on addressing several important issues related to the use of preclinical HD mouse models in the current large-scale HD drug discovery process. Because of space limitations, we will not be able to discuss in detail the specific preclinical compounds that have already been tested in HD mice. Excellent recent reviews on the subject can be found elsewhere (Beal and Ferrante, 2004; Butler and Bates, 2006; Di Prospero and Fischbeck, 2005; Hersch and Ferrante, 2004).

A Dual-Model Approach to Test HD Preclinical Candidates

Because of the rich repertoire of HD mouse models available, a critical first question is which mouse model or models should be used in a preclinical study of a therapeutic candidate. An ideal HD mouse model should have (1) a genetic construct that is similar if not identical to the patients; (2) robust behavioral deficits and selective neuropathology mimicking the patients; and (3) these phenotypes are early onset, rapidly progressing, easily quantifiable, and have limited variability between mice. From our description of the preclinical HD mouse models, it is apparent that each of the current models only partially satisfies such requirements. The fragment models (i.e., R6/2) have very rapid onset and progression of disease, as well as low phenotypic variability (10–20 per treatment group in preclinical trials); weight loss and early lethality in these models can be used as surrogate markers of disease. Thus, these fragment models are particularly suitable for screening a relatively large number of candidates and have already made significant contributions in identifying promising preclinical leads in HD (Beal and Ferrante, 2004). However, there are also some concerns with the use of mhtt fragment models to determine the preclinical efficacy of therapeutic candidates (Beal and Ferrante, 2004; Hersch and Ferrante, 2004). First, the neuropathology in the fragment models is more widespread and relatively nonselective. Second, because the disease progression in patients is usually very slow, another concern is that a potential efficacious therapeutic mechanism or candidate in human HD (i.e., with a more slowly progressing disease process) may not be effective in the fragment model. Third, the fragment models only express a small portion of mhtt regulatory elements and proteins; thus, potential pathogenic interventions requiring the intact mhtt genomic context (i.e., RNA interference based on the full-length human htt mRNA) or the protein context (i.e., proteolysis inhibition) could not be tested in such a model.

The availability of several full-length mhtt mouse models (i.e., full-length models; Table 7.1) provides new opportunities to test HD preclinical candidates in model systems that (1) have genetic, genomic, and protein context more similar to human HD; (2) have a slowly progressing disease process that may be more closely related to the process in the human disease; and (3) have neuropathology that is selective to the striatum and cortex. However, the main concern in the use of the full-length models in preclinical trials is cost, which is relatively high because of the length of the trial (i.e., up to 12 months in YAC128 and BACHD mice and up to 1–2 years in Hdh-KI mice), as well as because of the potential variability of their outcomes, which may require a larger number of mice for each study (see below). However, recent studies using YAC128 mice in preclinical studies provide some examples of full-length models being effectively used to identify promising preclinical candidates in HD (Van Raamsdonk et al., 2005b, 2005d).

It is clear that both the mhtt fragment models and the full-length models have their potential strengths and weaknesses in the preclinical studies. Without unequivocal proof that any of these HD models can predict the clinical outcome in patients (i.e., predictive validity), and faced with the daunting task of prioritizing a potentially large number of preclinical candidates for very costly and lengthy clinical studies, we and many others in the HD research community favor the dual-model approach of using both a fragment model and a full-length huntingtin mouse model for the preclinical study (Bates and Hockly, 2003; Beal and Ferrante, 2004). In this scheme (see Figure 7.1), a large number of preclinical candidates coming out of the screening assays will be first tested, if the therapeutic mechanism permits, in a fragment mouse model of HD. R6/2 mice may be a preferred model for this purpose because of the model’s rapid disease course and superior statistical power to detect efficacy (Hersch and Ferrante, 2004; Hockly et al., 2003). Such a therapeutic screen, with primary outcomes focusing on behavioral improvements (i.e., rotarod and grip strength) and neuroprotection (i.e., striatal atrophy), can provide relatively rapid information on the potential therapeutic efficacy of a large number of preclinical candidates at a reasonable cost. After demonstrating therapeutic efficacy in the fragment model (defined as both behavioral and pathological improvements), the next decision point is whether the compound should go directly to the clinical study or whether a second mouse trial in a full-length model is warranted. At this point, if the compound has proven safety in humans and/or is already in clinical use, direct clinical study in HD patients may be warranted. If a compound would require more extensive medicinal chemistry studies and toxicity and tolerability studies in animals and humans, which are time consuming and costly, we prefer the strategy of performing a second preclinical study in a full-length mouse model of HD before these compounds enter the clinical phase (Figure 7.1). Any compounds demonstrating therapeutic efficacy in such dual-model trials will be the best clinical candidates for further therapeutic development toward clinical trials. Of course, preclinical studies in a full-length mhtt model could also be initiated without a fragment model trial or despite a negative result from such trial if (1) the mechanism of therapy requires the genomic or protein context of the full-length mhtt (i.e., small interfering RNA against human full-length mRNA, or proteolysis inhibitor targeting mhtt regions beyond those in the fragment models); or (2) the therapeutic modality favors a more slowly progressive and/or selective disease process (i.e., aging-related therapeutic targets). In summary, in the bottleneck process of selecting preclinical candidates for further clinical studies, we believe the dual-model approach in which first screening with a fragment model of HD and then following up with a second study using a full-length model is the most rational and prudent approach to identify promising clinical candidates in HD.

FIGURE 7.1. A dual-model scheme for preclinical studies using both the fragment models and full-length models of HD.

FIGURE 7.1

A dual-model scheme for preclinical studies using both the fragment models and full-length models of HD.

Designing and Conducting Mouse Preclinical Trials

Preclinical trials in HD mice are an expensive and, in the case of a full-length model, relatively lengthy process. Therefore, before the initiation of such a study, careful design of the testing paradigms and rigorous standardization of the testing conditions and data analyses within a laboratory and across different laboratories are critical to produce accurate and reproducible results. The importance of mouse trial design and standardization in HD is highlighted in a Hereditary Disease Foundation-sponsored workshop at Cardiff in 2002 (see the workshop report at http://www.hdfoundation.org/workshops/200207Report.php) and reviewed by others (Bates and Hockly, 2003; Hersch and Ferrante, 2004; Hockly et al., 2003). In this section, we will discuss some of the salient points related to the design and execution of a mouse trial for HD.

Power Analyses to Determine the Sample Size for the Study

A very important consideration in the design of preclinical trials is to perform power analyses for a given preclinical outcome to determine the number of HD mice needed for a given set of outcomes in a preclinical trial (Bates and Hockly, 2003). Similar to the human clinical trials, insufficiently powered studies can be costly and may not provide the necessary information to assess the efficacy of a compound. Performing power analyses before the mouse trial will help one to determine how many animals will be needed in a study to have a predetermined reasonable chance (i.e., 80% or 90%) of detecting a significant improvement at a predetermined magnitude of improvement (i.e., 30% improvement). The basic formula for sample size estimation is the following (adapted from Hockly et al., 2003):

Image ch7_eq01.jpg

In this formula, n represents the number of mice in each treatment group; μ1 and μ2 are the means of the two groups; and SD1 and SD2 are the standard deviations in the two groups. Furthermore, y is 1.96 for P = 0.05 (two-sided normal distribution); and z is 1.28 for power of 90% and 0.84 for power of 80%.

Because of cost concerns, power analyses are crucial for minimizing cost by selecting outcomes and/or HD mouse models that may require a relatively smaller number of mice per treatment group to detect a significant therapeutic benefit. For example, the use of R6/2 mice normally requires only about 10 mice per group for most of the phenotypic outcomes, whereas the use of N171-82Q mice may require 20 mice per group (Hersch and Ferrante, 2004). For the full-length models, most of the potential preclinical outcomes in these models should be analyzed by power analyses to determine whether they are suitable for preclinical studies. In the case of the YAC128 model, striatal atrophy is a particularly good readout because only eight animals per group are needed to provide 80% power to detect a significant (P < 0.05) treatment benefit that is 33% or greater (Slow et al., 2003). In BACHD mice, both our analyses (Gray et al., 2008) and an independent analysis at PsychoGenics (Tarrytown, NY) (L. Menalled and D. Howland, personal communications) reveal that the rotarod deficit at 6 months is a particularly robust phenotype outcome: only 22 BACHD mice are needed per group to provide 80% power to detect significant (P < 0.05) improvement of 33% or better. Thus, in both fragment and full-length models of HD, power analyses should be done for all the major outcomes so that one can begin the study with sufficient power to detect a therapeutic benefit.

Rigorous Standardization of Preclinical Mouse Trials

A key to reducing the variability in a mouse trial is the rigorous standardization of the testing condition (see HDF Cardiff Workshop Report; also Bates and Hockly, 2003; Hockly et al., 2003). There are several areas to which the investigator should pay particular attention. First, the genetics of a given HD mouse model has to be carefully controlled. If possible, an inbred mouse background should be used for a study. We do not recommend that all the mouse trials for all the different types of HD models should be done in the same genetic background, although it is well known that certain behavioral readouts may be more robust in one inbred background than another (Crawley et al., 1997). We do recommend, however, within the same type of model, a consistent and relatively pure genetic background (i.e., inbred or recombinant F1) should be used, and preferably a genetic background in which a given model exhibits the most robust phenotypes that will be used as preclinical outcomes (i.e., YAC128 model in FvB background) (Van Raamsdonk et al., 2007). For the fragment models such as R6/2, which can only be maintained in a mixed F1 background, one should consistently split transgenic littermates between the treatment group and control group, and use litters from different breeding pairs to avoid potential breeder effects (Hockly et al., 2003). Because CAG repeat size can vary in certain mouse models, such as R6/2 and Hdh-KI mice, the mice used in the study should be routinely genotyped to rule out any mice with dramatic expansion or contraction of the repeats, which may inf luence the phenotypes. Second, mouse housing and diet conditions should also be standardized. There are several reports that demonstrate environmental enrichment and dietary inf luences, which may significantly improve outcomes in HD mice (Hockly et al., 2002; Schilling et al., 2004a; Spires et al., 2004). For R6/2 mice, a moderately enriched housing condition is recommended (Hockly et al., 2003). For other HD mouse models, such as the full-length models, the housing conditions should be maintained constant in different trials and comparable across different laboratories. Finally, the outcome testing conditions and statistical analyses should also be the same for a given model in different trials and different laboratories. For example, the equipment and methods for behavioral testing and neuropathological studies (i.e., stereological counting and dark neuron quantization) may be different from laboratory to laboratory. Therefore, it is critical that those laboratories using the same preclinical mouse model reach a consensus on a set of standardized protocols for various outcome studies and make such standardized protocols accessible to others in the community. For the phenotypic studies, the investigators conducting the tests should be blind to the genotypes. Finally, statistical analyses should be performed in the same way for a given outcome in a given model, so one can readily compare the preclinical results of different compounds and their ability to be replicated by different laboratories.

Improving the Outcome Measures in Preclinical Mouse Studies

In the majority of preclinical trials with HD mice, the primary phenotypic outcomes are rotarod for motor behaviors and stereological measurement of the striatal atrophy. Secondary outcomes, such as weight loss and lethality, are only possible for the fragment models, and their relevance to HD remains unclear. However, in HD patients, the clinical features are far more complex and consist of motor, psychiatric, and cognitive deficits. Because we are currently unclear whether this clinical triad is caused by neurodegeneration or neuronal dysfunction in HD, and whether different brain regions may mediate these distinct deficits, the exclusive use of motor and striatal pathology outcomes may be insufficient, particularly in evaluating candidates that may selectively improve cognitiveand/or psychiatric-related phenotypes (which may be highly relevant to HD). For example, a recent preclinical study with R6/2 mice showed that the imposition of sleep with alprazolam selectively improves the cognitive deficits in this model but not motor deficits (Pallier et al., 2007). Currently, there is a concerted effort to develop more sensitive potentially automatable cognitive behavioral outcomes for the preclinical models of HD (Lione et al., 1999; Morton et al., 2006; Van Raamsdonk, et al., 2005e). In our view, the use of robust cognitive measures, especially those affecting the same corticostriatal neural circuit as in patients, should be evaluated carefully and should be included when possible as primary outcome measures in HD preclinical studies.

Finally, another concerted effort in the HD therapeutic field is the attempt to discover biomarkers that may reflect the disease process in patients and may be used in future clinical studies. Some of these potential biomarkers, such as gene expression alterations and/or specific molecular changes in the serum such as increase in 8-OHDG (a marker for oxidized damage to DNA), have already been elucidated in patients (Dalrymple et al., 2007; Hersch et al., 2006; Runne et al., 2007). Importantly, in a recent clinical study with creatine, 8-OHDG was shown to dramatically normalize with the treatment, thus indicating it may be a useful peripheral biomarker that can predict therapeutic benefit (Hersch et al., 2006). If similar biomarkers also are altered in preclinical mouse models, they could be incorporated in the mouse trial to provide the opportunity to noninvasively monitor preclinical outcomes in mice and clinical outcomes in patients.

TRANSLATING PRECLINICAL STUDIES TO THE CLINIC

The ultimate goal for the preclinical studies with HD mice is to develop clinical candidates that can eventually succeed in human clinical trials. This goal will not only validate a specific mouse model and/or other nonmammalian models of HD but also validate the entire rational approach we are undertaking to develop an effective treatment or a cure for HD. Currently, with the predominant use of fragment models of HD, a large set of compounds have already been demonstrated to have preclinical efficacy (Beal and Ferrante, 2004). Several of these compounds, including creatine, coenzyme Q, α-lipoic acid, cystamine, and histone deacetylase inhibitors, have been consistently beneficial in multiple HD mouse models and in trials done by different laboratories (Beal and Ferrante, 2004). Furthermore, cystamine remains the only compound that satisfies the dual-model testing paradigm, demonstrating some efficacy in both fragment models (Beal and Ferrante, 2004) and in the full-length YAC128 model (Van Raamsdonk et al., 2005b).

If the role of a genetic mouse model in a rational development of therapy for HD or other brain disorders remains unproven, we may learn some lessons from another field in which 30 years of intensive research has brought some visible successes in the clinic. In the cancer field, mouse genetic models have provided critical mechanistic insights into the disease pathogenic process (Frese and Tuveson, 2007; Van Dyke and Jacks, 2002); some lessons about the difference in the biology between the mouse and human, which allowed the building of better mouse models (Rangarajan and Weinberg, 2003); and eventually a few success stories—in the case for acute promyelocytic leukemia (APL), a mouse model of APL provided promising preclinical results (Lallemand-Breitenbach et al., 1999), which subsequently translated into effective new treatment for patients (Soignet and Maslak, 2004). In the HD field, a relatively encouraging Phase III clinical trial demonstrated coenzyme Q10, which has reproducible efficacy in the preclinical models of HD (i.e., R6/2 and N171-82Q mice), also reveals a nonsignificant trend (14%) of improvement in HD patients. With the cancer field in mind, we are optimistic that the preclinical HD mouse models will not only teach us about the intricate pathogenic mechanisms in HD but will also play a critical role in identifying promising therapeutic candidates that can be translated into the clinic.

ACKNOWLEDGMENTS

We apologize for our omission, in part because of space limitations, of citations or topics. We thank the National Institute of Neurodegenerative Disorders and Stroke/National Institutes of Health (Grants 5R01NS049501 and 1R21NS047391), Hereditary Disease Foundation, High Q Foundation, and CHDI Foundation, Inc. (to X. W. Yang) for their support. We also thank Nancy Wexler, Carl Johnson, Allan Tobin, and Ethan Signer for their support. We thank Bernard Balleine at UCLA, Liliana Menalled and Daniela Brunner at PsychoGenics, and David Howland at High Q Foundation for sharing unpublished data. Finally, we thank Xiaohong Lu, Xiaofeng Gu, Binh Vuong, and other members of the Yang laboratory for discussions and for help in preparing the chapter manuscript.

REFERENCES

  • Andreassen O. A., Dedeoglu A., Ferrante R. J., Jenkins B. G., Ferrante K. L., Thomas M., Friedlich A., Browne S. E., Schilling G., Borchelt D. R., et al. Creatine increase survival and delays motor symptoms in a transgenic animal model of Huntington’s disease. Neurobiol Dis. 2001;8(3):479–91. [PubMed: 11447996]
  • Arguello P. A., Gogos J. A. Modeling madness in mice: one piece at a time. Neuron. 2006;52(1):179–96. [PubMed: 17015235]
  • Aron A. R., Sahakian B. J., Robbins T. W. Distractibility during selection-for-action: differential deficits in Huntington’s disease and following frontal lobe damage. Neuropsychologia. 2003;41(9):1137–47. [PubMed: 12753954]
  • Arrasate M., Mitra S., Schweitzer E. S., Segal M. R., Finkbeiner S. Inclusion body formation reduces levels of mutant huntingtin and the risk of neuronal death. Nature. 2004;431(7010):805–10. [PubMed: 15483602]
  • Aylward E. H., Sparks B. F., Field K. M., Yallapragada V., Shpritz B. D., Rosenblatt A., Brandt J., Gourley L. M., Liang K., Zhou H., et al. Onset and rate of striatal atrophy in preclinical Huntington disease. Neurology. 2004;63(1):66–72. [PubMed: 15249612]
  • Balleine B. W. Neural bases of food-seeking: affect, arousal and reward in corticostriatolimbic circuits. Physiol Behav. 2005;86(5):717–30. [PubMed: 16257019]
  • Bates G. Huntingtin aggregation and toxicity in Huntington’s disease. Lancet. 2003;361(9369):1642–44. [PubMed: 12747895]
  • Bates G. P., Hockly E. Experimental therapeutics in Huntington’s disease: are models useful for therapeutic trials? Curr Opin Neurol. 2003;16(4):465–70. [PubMed: 12869804]
  • Bates G. P., Harper P. S., Jones A. L., editors. Huntington’s Disease. Oxford; Oxford University Press: 2002.
  • Beal M. F., Ferrante R. J. Experimental therapeutics in transgenic mouse models of Huntington’s disease. Nat Rev Neurosci. 2004;5(5):373–84. [PubMed: 15100720]
  • Bogdanov M. B., Andreassen O. A., Dedeoglu A., Ferrante R. J., Beal M. F. Increased oxidative damage to DNA in a transgenic mouse model of Huntington’s disease. J Neurochem. 2001;79(6):1246–49. [PubMed: 11752065]
  • Bolivar V. J., Manley K., Messer A. Exploratory activity and fear conditioning abnormalities develop early in R6/2 Huntington’s disease transgenic mice. Behav Neurosci. 2003;117(6):1233–42. [PubMed: 14674843]
  • Branda C. S., Dymecki S. M. Talking about a revolution: the impact of site-specific recombinases on genetic analyses in mice. Dev Cell. 2004;6(1):7–28. [PubMed: 14723844]
  • Butler R., Bates G. P. Histone deacetylase inhibitors as therapeutics for polyglutamine disorders. Nat Rev Neurosci. 2006;7(10):784–96. [PubMed: 16988654]
  • Capecchi M. R. Gene targeting in mice: functional analysis of the mammalian genome for the twenty-first century. Nat Rev Genet. 2005;6(6):507–12. [PubMed: 15931173]
  • Carter R. J., Lione L. A., Humby T., Mangiarini L., Mahal A., Bates G. P., Dunnett S. B., Morton A. J. Characterization of progressive motor deficits in mice transgenic for the human Huntington’s disease mutation. J Neurosci. 1999;19(8):3248–57. [PMC free article: PMC6782264] [PubMed: 10191337]
  • Crawley J. N., Belknap J. K., Collins A., Crabbe J. C., Frankel W., Henderson N., Hitzemann R. J., Maxson S. C., Miner L. L., Silva A. J., et al. Behavioral phenotypes of inbred mouse strains: implications and recommendations for molecular studies. Psychopharmacology (Berl). 1997;132(2):107–24. [PubMed: 9266608]
  • Dalrymple A., Wild E. J., Joubert R., Sathasivam K., Bjorkqvist M., Petersen A., Jackson G. S., Isaacs J. D., Kristiansen M., Bates G. P., et al. Proteomic profiling of plasma in Huntington’s disease reveals neuroinflammatory activation and biomarker candidates. J Proteome Res. 2007;6(7):2833–40. [PubMed: 17552550]
  • Davies S. W., Turmaine M., Cozens B. A., DiFiglia M., Sharp A. H., Ross C. A., Scherzinger E., Wanker E. E., Mangiarini L., Bates G. P. Formation of neuronal intranuclear inclusions underlies the neurological dysfunction in mice transgenic for the HD mutation. Cell. 1997;90(3):537–48. [PubMed: 9267033]
  • Di Prospero N. A., Fischbeck K. H. Therapeutics development for triplet repeat expansion diseases. Nat Rev Genet. 2005;6(10):756–65. [PubMed: 16205715]
  • DiFiglia M., Sapp E., Chase K. O., Davies S. W., Bates G. P., Vonsattel J. P., Aronin N. Aggregation of huntingtin in neuronal intranuclear inclusions and dystrophic neurites in brain. Science. 1997;277(5334):1990–93. [PubMed: 9302293]
  • Fleming S. M., Fernagut P. O., Chesselet M. F. Genetic mouse models of parkinsonism: strengths and limitations. NeuroRx. 2005;2(3):495–503. [PMC free article: PMC1144493] [PubMed: 16389313]
  • Fossale E., Wheeler V. C., Vrbanac V., Lebel L. A., Teed A., Mysore J. S., Gusella J. F., MacDonald M. E., Persichetti F. Identification of a presymptomatic molecular phenotype in Hdh CAG knock-in mice. Hum Mol Genet. 2002;11(19):2233–41. [PubMed: 12217951]
  • Frese K. K., Tuveson D. A. Maximizing mouse cancer models. Nat Rev Cancer. 2007;7(9):654–58. [PubMed: 17687385]
  • Gardian G., Browne S. E., Choi D. K., Klivenyi P., Gregorio J., Kubilus J. K., Ryu H., Langley B., Ratan R. R., Ferrante R. J., et al. Neuroprotective effects of phenylbutyrate in the N171-82Q transgenic mouse model of Huntington’s disease. J Biol Chem. 2005;280(1):556–63. [PubMed: 15494404]
  • Gatchel J. R., Zoghbi H. Y. Diseases of unstable repeat expansion: mechanisms and common principles. Nat Rev Genet. 2005;6(10):743–55. [PubMed: 16205714]
  • Graham R. K., Slow E. J., Deng Y., Bissada N., Lu G., Pearson J., Shehadeh J., Leavitt B. R., Raymond L. A., Hayden M. R. Levels of mutant huntingtin influence the phenotypic severity of Huntington disease in YAC128 mouse models. Neurobiol Dis. 2006;21(2):444–55. [PubMed: 16230019]
  • Gray M., Shirasaki D. I., Cepeda C., Andre V. M., Wilburn B., Lu X.-H., Tao J., Yamazaki I., Li S.-H., Sun Y. E., et al. Full-length human mutant huntingtin with a stable polyglutamine repeat can elicit progressive and selective neuropathogenesis in BACHD mice. J Neurosci. 2008;28(24):6182–95. [PMC free article: PMC2630800] [PubMed: 18550760]
  • Gu X., Li C., Wei W., Lo V., Gong S., Li S. H., Iwasato T., Itohara S., Li X. J., Mody I., et al. Pathological cell-cell interactions elicited by a neuropathogenic form of mutant Huntingtin contribute to cortical pathogenesis in HD mice. Neuron. 3. 2005;46:433–44. [PubMed: 15882643]
  • Gusella J. F., Macdonald M. E. Huntington’s disease: seeing the pathogenic pro-cess through a genetic lens. Trends Biochem Sci. 2006;31(9):533–40. [PubMed: 16829072]
  • Gutekunst C. A., Li S. H., Yi H., Mulroy J. S., Kuemmerle S., Jones R., Rye D., Ferrante R. J., Hersch S. M., Li X. J. Nuclear and neuropil aggregates in Huntington’s disease: relationship to neuropathology. J Neurosci. 1999;19(7):2522–34. [PMC free article: PMC6786077] [PubMed: 10087066]
  • Hedreen J. C., Peyser C. E., Folstein S. E., Ross C. A. Neuronal loss in layers V and VI of cerebral cortex in Huntington’s disease. Neurosci Lett. 1991;133(2):257–61. [PubMed: 1840078]
  • Heinsen H., Strik M., Bauer M., Luther K., Ulmar G., Gangnus D., Jungkunz G., Eisenmenger W., Gotz M. Cortical and striatal neurone number in Huntington’s disease. Acta Neuropathol (Berl). 1994;88(4):320–33. [PubMed: 7839825]
  • Heintz N. BAC to the future: the use of bac transgenic mice for neuroscience research. Nat Rev Neurosci. 2001;2(12):861–70. [PubMed: 11733793]
  • Heng M. Y., Tallaksen-Greene, J S., Detloff P. J., Albin R. L. Longitudinal evaluation of the Hdh(CAG)150 knock-in murine model of Huntington’s disease. J Neurosci. 2007;27(34):8989–98. [PMC free article: PMC6672210] [PubMed: 17715336]
  • Hersch S. M., Ferrante R. J. Translating therapies for Huntington’s disease from genetic animal models to clinical trials. NeuroRx. 2004;1(3):298–306. [PMC free article: PMC534928] [PubMed: 15717031]
  • Hersch S. M., Gevorkian S., Marder K., Moskowitz C., Feigin A., Cox M., Como P., Zimmerman C., Lin M., Zhang L., et al. Creatine in Huntington disease is safe, tolerable, bioavailable in brain and reduces serum 8OH2′dG. Neurology. 2. 2006;66:250–52. [PubMed: 16434666]
  • Hickey M. A., Gallant K., Gross G. G., Levine M. S., Chesselet M. F. Early behavioral deficits in R6/2 mice suitable for use in preclinical drug testing. Neurobiol Dis. 2005;20(1):1–11. [PubMed: 16137562]
  • Hockly E., Cordery P. M., Woodman B., Mahal A., van Dellen A., Blakemore C., Lewis C. M., Hannan A. J., Bates G. P. Environmental enrichment slows disease progression in R6/2 Huntington’s disease mice. Ann Neurol. 2002;51(2):235–42. [PubMed: 11835380]
  • Hockly E., Woodman B., Mahal A., Lewis C. M., Bates G. Standardization and statistical approaches to therapeutic trials in the R6/2 mouse. Brain Res Bull. 2003;61(5):469–79. [PubMed: 13679245]
  • Hodgson J. G., Agopyan N., Gutekunst C. A., Leavitt B. R., LePiane F., Singaraja R., Smith D. J., Bissada N., McCutcheon K., Nasir J., et al. A YAC mouse model for Huntington’s disease with full-length mutant huntingtin, cytoplasmic toxicity, and selective striatal neurodegeneration. Neuron. 1999;23(1):181–92. [PubMed: 10402204]
  • Holmes A., Wrenn C. C., Harris A. P., Thayer K. E., Crawley J. N. Behavioral profiles of inbred strains on novel olfactory, spatial and emotional tests for reference memory in mice. Genes Brain Behav. 2002;1(1):55–69. [PubMed: 12886950]
  • Hughes R. E., Olson J. M. Therapeutic opportunities in polyglutamine disease. Nat Med. 2001;7(4):419–23. [PubMed: 11283667]
  • The Huntington’s Disease Collaborative Research Group. A novel gene containing a trinucleotide repeat that is expanded and unstable on Huntington’s disease chromosomes. Cell. 1993;72(6):971–83. [PubMed: 8458085]
  • Jenkins B. G., Klivenyi P., Kustermann E., Andreassen O. A., Ferrante R. J., Rosen B. R., Beal M. F. Nonlinear decrease over time in N-acetyl aspartate levels in the absence of neuronal loss and increases in glutamine and glucose in transgenic Huntington’s disease mice. J Neurochem. 2000;74(5):2108–19. [PubMed: 10800956]
  • Kazantsev A., Preisinger E., Dranovsky A., Goldgaber D., Housman D. Insoluble detergent-resistant aggregates form between pathological and nonpathological lengths of polyglutamine in mammalian cells. Proc Natl Acad Sci U S A. 1999;96(20):11404–09. [PMC free article: PMC18046] [PubMed: 10500189]
  • Kovtun I. V., Liu Y., Bjoras M., Klungland A., Wilson S. H., McMurray C. T. OGG1 initiates age-dependent CAG trinucleotide expansion in somatic cells. Nature. 2007;447(7143):447–52. [PMC free article: PMC2681094] [PubMed: 17450122]
  • Kuhn A., Goldstein D. R., Hodges A., Strand A. D., Sengstag T., Kooperberg C., Becanovic K., Pouladi M. A., Sathasivam K., Cha J. H., et al. Mutant huntingtin’s effects on striatal gene expression in mice recapitulate changes observed in human Huntington’s disease brain and do not differ with mutant huntingtin length or wild-type huntingtin dosage. Hum Mol Genet. 2007;16(15):1845–61. [PubMed: 17519223]
  • Lallemand-Breitenbach V., Guillemin M. C., Janin A., Daniel M. T., Degos L., Kogan S. C., Bishop J. M., de The H. Retinoic acid and arsenic synergize to eradicate leukemic cells in a mouse model of acute promyelocytic leukemia. J Exp Med. 1999;189(7):1043–52. [PMC free article: PMC2193002] [PubMed: 10190895]
  • Landles C., Bates G. P. Huntingtin and the molecular pathogenesis of Huntington’s disease. Fourth in molecular medicine review series. EMBO Rep. 2004;5(10):958–63. [PMC free article: PMC1299150] [PubMed: 15459747]
  • Levine M. S., Cepeda C., Hickey M. A., Fleming S. M., Chesselet M. F. Genetic mouse models of Huntington’s and Parkinson’s diseases: illuminating but imperfect. Trends Neurosci. 2004;27(11):691–97. [PubMed: 15474170]
  • Li H., Li S. H., Cheng A. L., Mangiarini L., Bates G. P., Li X. J. Ultrastructural localization and progressive formation of neuropil aggregates in Huntington’s disease transgenic mice. Hum Mol Genet. 1999;8(7):1227–36. [PubMed: 10369868]
  • Li J. Y., Popovic N., Brundin P. The use of the R6 transgenic mouse models of Huntington’s disease in attempts to develop novel therapeutic strategies. NeuroRx. 2005;2(3):447–64. [PMC free article: PMC1144488] [PubMed: 16389308]
  • Li S., Li X. J. Multiple pathways contribute to the pathogenesis of Huntington disease. Mol Neurodegener. 2006;1 [PMC free article: PMC1764744] [PubMed: 17173700]
  • Lin C. H., Tallaksen-Greene S., Chien W. M., Cearley J. A., Jackson W. S., Crouse A. B., Ren S., Li X. J., Albin R. L., Detloff P. J. Neurological abnormalities in a knock-in mouse model of Huntington’s disease. Hum Mol Genet. 2001;10(2):137–44. [PubMed: 11152661]
  • Lione L. A., Carter R. J., Hunt M. J., Bates G. P., Morton A. J., Dunnett S. B. Selective discrimination learning impairments in mice expressing the human Huntington’s disease mutation. J Neurosci. 1999;19(23):10428–37. [PMC free article: PMC6782405] [PubMed: 10575040]
  • Lloret A., Dragileva E., Teed A., Espinola J., Fossale E., Gillis T., Lopez E., Myers R. H., MacDonald M. E., Wheeler V. C. Genetic background modifies nuclear mutant huntingtin accumulation and HD CAG repeat instability in Huntington’s disease knock-in mice. Hum Mol Genet. 2006;15(12):2015–24. [PubMed: 16687439]
  • Mangiarini L., Sathasivam K., Seller M., Cozens B., Harper A., Hetherington C., Lawton M., Trottier Y., Lehrach H., Davies S. W., et al. Exon 1 of the HD gene with an expanded CAG repeat is sufficient to cause a progressive neurological phenotype in transgenic mice. Cell. 1996;87(3):493–506. [PubMed: 8898202]
  • Marsh J. L., Thompson L. M. Drosophila in the study of neurodegenerative disease. Neuron. 2006;52(1):169–78. [PubMed: 17015234]
  • Menalled L. B. Knock-in mouse models of Huntington’s disease. NeuroRx. 2005;2(3):465–70. [PMC free article: PMC1144489] [PubMed: 16389309]
  • Menalled L. B., Chesselet M. F. Mouse models of Huntington’s disease. Trends Pharmacol Sci. 2002;23(1):32–39. [PubMed: 11804649]
  • Menalled L. B., El-Khodor F., Patry M., Suarez-Farinas M., Orenstein S.J., Zahasky B., Leahy C., Wheeler V., Yang X. W., MacDonald M., et al. Systematic behavioral evaluation of Huntington’s disease transgenic and knock-in mouse models. Neurobiol Dis. 2009;35(3):319–36. [PMC free article: PMC2728344] [PubMed: 19464370]
  • Menalled L. B., Sison J. D., Dragatsis I., Zeitlin S., Chesselet M. F. Time course of early motor and neuropathological anomalies in a knock-in mouse model of Huntington’s disease with 140 CAG repeats. J Comp Neurol. 2003;465(1):11–26. [PubMed: 12926013]
  • Menalled L. B., Sison J. D., Wu Y., Olivieri M., Li X. J., Li H., Zeitlin S., Chesselet M. F. Early motor dysfunction and striosomal distribution of huntingtin microaggregates in Huntington’s disease knock-in mice. J Neurosci. 2002;22(18):8266–76. [PMC free article: PMC6758087] [PubMed: 12223581]
  • Morton A. J., Skillings E., Bussey T. J., Saksida L. M. Measuring cognitive deficits in disabled mice using an automated interactive touchscreen system. Nat Methods. 2006;3(10) [PubMed: 16990806]
  • Morton A. J., Wood N. I., Hastings M. H., Hurelbrink C., Barker R. A., Maywood E. S. Disintegration of the sleep-wake cycle and circadian timing in Huntington’s disease. J Neurosci. 2005;25(1):157–63. [PMC free article: PMC6725210] [PubMed: 15634777]
  • Mouse Genome Sequencing Consortium.2002Initial sequencing and comparative analysis of the mouse genome Nature 420(6915):520–62. [PubMed: 12466850]
  • Murphy K. P., Carter R. J., Lione L. A., Mangiarini L., Mahal A., Bates G. P., Dunnett S. B., Morton A. J. Abnormal synaptic plasticity and impaired spatial cognition in mice transgenic for exon 1 of the human Huntington’s disease mutation. J Neurosci. 2000;20(13):5115–23. [PMC free article: PMC6772265] [PubMed: 10864968]
  • Orr H. T., Zoghbi H. Y. Trinucleotide repeat disorders. Annu Rev Neurosci. 2007;30:575–621. [PubMed: 17417937]
  • Pallier P. N., Maywood E. S., Zheng Z., Chesham J. E., Inyushkin A. N., Dyball R., Hastings M. H., Morton A. J. Pharmacological imposition of sleep slows cognitive decline and reverses dysregulation of circadian gene expression in a transgenic mouse model of Huntington’s disease. J Neurosci. 2007;27(29):7869–78. [PMC free article: PMC6672877] [PubMed: 17634381]
  • Rangarajan A., Weinberg R. A. Opinion: Comparative biology of mouse versus human cells: modelling human cancer in mice. Nat Rev Cancer. 2003;3(12):952–59. [PubMed: 14737125]
  • Reddy P. H., Williams M., Charles V., Garrett L., Pike-Buchanan L., Whetsell W. O. Jr, Miller G., Tagle D. A. Behavioural abnormalities and selective neuronal loss in HD transgenic mice expressing mutated full-length HD cDNA. Nat Genet. 1998;20(2):198–202. [PubMed: 9771716]
  • Rogers D. C., Fisher E. M., Brown S. D., Peters J., Hunter A. J., Martin J. E. Behavioral and functional analysis of mouse phenotype: SHIRPA, a proposed protocol for comprehensive phenotype assessment. Mamm Genome. 1997;8(10):711–13. [PubMed: 9321461]
  • Rosas H. D., Hevelone N. D., Zaleta A. K., Greve D. N., Salat D. H., Fischl B. Regional cortical thinning in preclinical Huntington disease and its relationship to cognition. Neurology. 2005;65(5):745–47. [PubMed: 16157910]
  • Rosas H. D., Tuch D. S., Hevelone N. D., Zaleta A. K., Vangel M., Hersch S. M., Salat D. H. Diffusion tensor imaging in presymptomatic and early Huntington’s disease: selective white matter pathology and its relationship to clinical measures. Mov Disord. 2006;21(9):1317–25. [PubMed: 16755582]
  • Runne H., Kuhn A., Wild E. J., Pratyaksha W., Kristiansen M., Isaacs J. D., Regulier E., Delorenzi M., Tabrizi S. J., Luthi-Carter R. Analysis of potential transcriptomic biomarkers for Huntington’s disease in peripheral blood. Proc Natl Acad Sci U S A. 2007;104(36):14424–29. [PMC free article: PMC1964868] [PubMed: 17724341]
  • Schilling G., Becher M. W., Sharp A. H., Jinnah H. A., Duan K., Kotzuk J. A., Slunt H. H., Ratovitski T., Cooper J. K., Jenkins N. A., et al. Intranuclear inclusions and neuritic aggregates in transgenic mice expressing a mutant N-terminal fragment of huntingtin. Hum Mol Genet. 1999;8(3):397–407. [PubMed: 9949199]
  • Schilling G., Savonenko A. V., Coonfield M. L., Morton J. L., Vorovich E., Gale A., Neslon C., Chan N., Eaton M., Fromholt D., Ross C. A., et al. Environmental, pharmacological, and genetic modulation of the HD phenotype in transgenic mice. Exp Neurol. 2004a;187(1):137–49. [PubMed: 15081595]
  • Schilling G., Savonenko A. V., Klevytska A., Morton J. L., Tucker S. M., Poirier M., Gale A., Chan N., Gonzales V., Slunt H. H., et al. Nuclear-targeting of mutant huntingtin fragments produces Huntington’s disease-like phenotypes in transgenic mice. Hum Mol Genet. 2004b;13(15):1599–610. [PubMed: 15190011]
  • Schilling G., Sharp A. H., Loev S. J., Wagster M. V., Li S. H., Stine O. C., Ross C. A. Expression of the Huntington’s disease (IT15) protein product in HD patients. Hum Mol Genet. 1995;4(8):1365–71. [PubMed: 7581375]
  • Seong I. S., Ivanova E., Lee J. M., Choo Y. S., Fossale E., Anderson M., Gusella J. F., Laramie J. M., Myers R. H., Lesort M., et al. HD CAG repeat implicates a dominant property of huntingtin in mitochondrial energy metabolism. Hum Mol Genet. 2005;14(19):2871–80. [PubMed: 16115812]
  • Sharp A. H., Loev S. J., Schilling G., Li S. H., Li X. J., Bao J., Wagster M. V., Kotzuk J. A., Steiner J. P., Lo A., et al. Widespread expression of Huntington’s disease gene (IT15) protein product. Neuron. 1995;14(5):1065–74. [PubMed: 7748554]
  • Silva E., Simpson, Takahashi J., Lipp H., Nakanishi S., Wehner J., Giese K., Tully T., Abel T., Chapman P. Mutant mice and neuroscience: recommendations concerning genetic background. Neuron. 1997;19(4):755–59. [PubMed: 9354323]
  • Simmons D. A., Casale M., Alcon B., Pham N., Narayan N., Lynch G. Ferritin accumulation in dystrophic microglia is an early event in the development of Huntington’s disease. Glia. 2007;55(10):1074–84. [PubMed: 17551926]
  • Sipione S., Cattaneo E. Modeling Huntington’s disease in cells, flies, and mice. Mol Neurobiol. 2001;23(1):21–51. [PubMed: 11642542]
  • Slow E. J., van Raamsdonk J., Rogers D., Coleman S. H., Graham R. K., Deng Y., Oh R., Bissada N., Hossain S. M., Yang Y. Z., et al. Selective striatal neuronal loss in a YAC128 mouse model of Huntington disease. Hum Mol Genet. 2003;12(13):1555–67. [PubMed: 12812983]
  • Soignet S., Maslak P. Therapy of acute promyelocytic leukemia. Adv Pharmacol. 2004;51:35–58. [PubMed: 15464904]
  • Spires T. L., Grote H. E., Garry S., Cordery P. M., Van Dellen A., Blakemore C., Hannan A. J. Dendritic spine pathology and deficits in experience-dependent dendritic plasticity in R6/1 Huntington’s disease transgenic mice. Eur J Neurosci. 2004;19(10):2799–807. [PubMed: 15147313]
  • Swerdlow N. R., Braff D. L., Taaid N., Geyer M. A. Assessing the validity of an animal model of deficient sensorimotor gating in schizophrenic patients. Arch Gen Psychiatry. 1994;51(2):139–54. [PubMed: 8297213]
  • Swerdlow N. R., Paulsen J., Braff D. L., Butters N., Geyer M. A., Swenson M. R. Impaired prepulse inhibition of acoustic and tactile startle response in patients with Huntington’s disease. J Neurol Neurosurg Psychiatry. 1995;58(2):192–200. [PMC free article: PMC1073317] [PubMed: 7876851]
  • Tallaksen-Greene S. J., Crouse A. B., Hunter J. M., Detloff P. J., Albin R. L. Neuronal intranuclear inclusions and neuropil aggregates in HdhCAG(150) knockin mice. Neuroscience. 2005;131(4):843–52. [PubMed: 15749339]
  • Turmaine M., Raza A., Mahal A., Mangiarini L., Bates G. P., Davies S. W. Nonapoptotic neurodegeneration in a transgenic mouse model of Huntington’s disease. Proc Natl Acad Sci U S A. 2000;97(14):8093–97. [PMC free article: PMC16675] [PubMed: 10869421]
  • Van Dyke T., Jacks T. Cancer modeling in the modern era: progress and challenges. Cell. 2002;108(2):135–44. [PubMed: 11832204]
  • Van Raamsdonk J. M., Gibson W. T., Pearson J., Murphy Z., Lu G., Leavitt B. R., Hayden M. R. Body weight is modulated by levels of full-length huntingtin. Hum Mol Genet. 2006;15(9):1513–23. [PubMed: 16571604]
  • Van Raamsdonk J. M., Murphy Z., Slow E. J., Leavitt B. R., Hayden M. R. Selective degeneration and nuclear localization of mutant huntingtin in the YAC128 mouse model of Huntington disease. Hum Mol Genet. 2005a;14(24):3823–35. [PubMed: 16278236]
  • Van Raamsdonk J. M., Pearson J., Bailey C. D., Rogers D. A., Johnson G. V., Hayden M. R., Leavitt B. R. Cystamine treatment is neuroprotective in the YAC128 mouse model of Huntington disease. J Neurochem. 2005b;95(1):210–20. [PubMed: 16181425]
  • Van Raamsdonk J. M., Pearson J., Rogers D. A., Bissada N., Vogl A. W., Hayden M. R., Leavitt B. R. Loss of wild-type huntingtin influences motor dysfunction and survival in the YAC128 mouse model of Huntington disease. Hum Mol Genet. 2005c;14(10):1379–92. [PubMed: 15829505]
  • Van Raamsdonk J. M., Pearson J., Rogers D. A., Lu G., Barakauskas V. E., Barr A. M., Honer W. G., Hayden M. R., Leavitt B. R. Ethyl-EPA treatment improves motor dysfunction, but not neurodegeneration in the YAC128 mouse model of Huntington disease. Exp Neurol. 2005d;196(2):266–72. [PubMed: 16129433]
  • Van Raamsdonk J. M., Pearson J., Slow E. J., Hossain S. M., Leavitt B. R., Hayden M. R. 2005e.Cognitive dysfunction precedes neuropathology and motor abnormalities in the YAC128 mouse model of Huntington’s disease J Neurosci 25(16):4169–80. [PMC free article: PMC6724951] [PubMed: 15843620]
  • Van Raamsdonk J. M., Warby S. C., Hayden M. R. Selective degeneration in YAC mouse models of Huntington disease. Brain Res Bull. 2007;72(2–3):124–31. [PubMed: 17352936]
  • Vonsattel J. P., DiFiglia M. Huntington disease. J Neuropathol Exp Neurol. 1998;57(5):369–84. [PubMed: 9596408]
  • Watase K., Zoghbi H. Y. Modelling brain diseases in mice: the challenges of design and analysis. Nat Rev Genet. 2003;4(4):296–307. [PubMed: 12671660]
  • Wexler N. S., Rose E. A., Housman D. E. Molecular approaches to hereditary diseases of the nervous system: Huntington’s disease as a paradigm. Annu Rev Neurosci. 1991;14:503–29. [PubMed: 1827708]
  • Wheeler V. C., Auerbach W., White J. K., Srinidhi J., Auerbach A., Ryan A., Duyao M. P., Vrbanac V., Weaver M., Gusella J. F., et al. Length-dependent gametic CAG repeat instability in the Huntington’s disease knock-in mouse. Hum Mol Genet. 1999;8(1):115–22. [PubMed: 9887339]
  • Wheeler V. C., Gutekunst C. A., Vrbanac V., Lebel L. A., Schilling G., Hersch S., Friedlander R. M., Gusella J. F., Vonsattel J. P., Borchelt D. R., et al. Early phenotypes that presage late-onset neurodegenerative disease allow testing of modifiers in Hdh CAG knock-in mice. Hum Mol Genet. 2002;11(6):633–40. [PubMed: 11912178]
  • Wheeler V. C., Lebel L. A., Vrbanac V., Teed A., te Riele H., MacDonald M. E. Mismatch repair gene Msh2 modifies the timing of early disease in Hdh(Q111) striatum. Hum Mol Genet. 2003;12(3):273–81. [PubMed: 12554681]
  • Wheeler V. C., White J. K., Gutekunst C. A., Vrbanac V., Weaver M., Li X. J., Li S. H., Yi H., Vonsattel J. P., Gusella J. F., et al. Long glutamine tracts cause nuclear localization of a novel form of huntingtin in medium spiny striatal neurons in HdhQ92 and HdhQ111 knock-in mice. Hum Mol Genet. 2000;9(4):503–13. [PubMed: 10699173]
  • Woodman B., Butler R., Landles C., Lupton M. K., Tse J., Hockly E., Moffitt H., Sathasivam K., Bates G. P. The Hdh(Q150/Q150) knock-in mouse model of HD and the R6/2 exon 1 model develop comparable and widespread molecular phenotypes. Brain Res Bull. 2007;72(2–3):83–97. [PubMed: 17352931]
  • Yamamoto A., Lucas J. J., Hen R. Reversal of neuropathology and motor dysfunction in a conditional model of Huntington’s disease. Cell. 2000;101(1):57–66. [PubMed: 10778856]
  • Yang X. W., Model P., Heintz N. Homologous recombination based modification in Escherichia coli and germline transmission in transgenic mice of a bacterial artificial chromosome. Nat Biotechnol. 1997;15(9):859–65. [PubMed: 9306400]
  • Yu Z. X., Li S. H., Evans J., Pillarisetti A., Li H., Li X. J. Mutant huntingtin causes context-dependent neurodegeneration in mice with Huntington’s disease. J Neurosci. 2003;23(6):2193–202. [PMC free article: PMC6742008] [PubMed: 12657678]
  • Zeitlin S., Liu J. P., Chapman D. L., Papaioannou V. E., Efstratiadis A. Increased apoptosis and early embryonic lethality in mice nullizygous for the Huntington’s disease gene homologue. Nat Genet. 1995;11(2):155–63. [PubMed: 7550343]
  • Zoghbi H. Y., Orr H. T. Glutamine repeats and neurodegeneration. Annu Rev Neurosci. 2000;23:217–47. [PubMed: 10845064]
Copyright © 2011 by Taylor and Francis Group, LLC.
Bookshelf ID: NBK56001PMID: 21882419

Views

  • PubReader
  • Print View
  • Cite this Page

Other titles in this collection

Related information

  • PMC
    PubMed Central citations
  • PubMed
    Links to PubMed

Similar articles in PubMed

See reviews...See all...

Recent Activity

Your browsing activity is empty.

Activity recording is turned off.

Turn recording back on

See more...
-