Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2014 Jun 15.
Published in final edited form as: Mov Disord. 2013 Jun 15;28(7):990–1000. doi: 10.1002/mds.25583

Engineering animal models of dystonia

Janneth Oleas 1, Fumiaki Yokoi 1, Mark P DeAndrade 1, Antonio Pisani 2, Yuqing Li 1,*
PMCID: PMC3800691  NIHMSID: NIHMS513931  PMID: 23893455

Abstract

Dystonia is a neurological disorder characterized by abnormal involuntary movements that are prolonged and often cause twisting and turning. Several genetically modified worms, fruit flies, and rodents have been generated as models of genetic dystonias, and in particular DYT1, DYT11, and DYT12 dystonias. Although these models do not show overt dystonic symptoms, the rodent models exhibit pronounced motor deficits in specialized behavioral tasks, such as the rotarod and beam-walking tests. For example, in a rodent model of DYT12 dystonia, which is generally stress triggered, motor deficits are observed only after the animal is stressed. Moreover, in a rodent model of DYT1 dystonia, the motor and electrophysiological deficits can be rescued by trihexyphenidyl, a common anticholinergic medication used to treat dystonic symptoms in human patients. Biochemically, the DYT1 and DYT11 animal models also share some similarities to patients, such as a reduction in striatal D2 dopamine receptor and binding activities. Additionally, conditional knockout mouse models for DYT1 and DYT11 dystonia show that the loss of the causal dystonia related proteins in the striatum lead to motor deficits. Interestingly, loss of the DYT1 dystonia causal protein in Purkinje cells shows an improvement in motor performance, suggesting that gene therapy targeting of the cerebellum or intervention in its downstream pathways may be useful. Finally, recent studies using DYT1 dystonia worm and mouse models led to a potential novel therapeutic agent, which is currently undergoing clinical trials. These results indicate that genetic animal models are an extremely powerful tool to elucidate the pathophysiology and to further develop new therapeutics for dystonia.

1. Introduction

Dystonia is a neurological disorder characterized by sustained contractions of muscles, which cause abnormal movement, twisting, and postures.1 Several methods of classifying dystonia exist. The earliest method of classifying dystonia was to segregate it as either primary or secondary.2 In primary dystonia, the dystonic symptoms are not caused by another condition or environmental factor, and can be hereditary or nonhereditary. In secondary dystonia, the dystonic symptoms are usually the result of another condition (e.g. stroke, brain injury, or metabolic disease) or certain medications.3 A more recent classification system has been developed that segregates dystonia based on clinical characteristics and etiology.1 Current hypotheses suggest that dystonia is a neurodevelopmental disorder4 that is caused by an abnormal motor circuitry in the cerebral cortex, basal ganglia, thalamus, and cerebellum, which in turn leads to abnormal synaptic plasticity and altered neurotransmission5-10, and thereby likely causes the dystonic symptoms.

The animal models of dystonia can be classified into two fundamental groups: phenotypic and genotypic. The primary goal of the phenotypic animal models is to mimic the dystonic phenotype seen in patients. One common method has been to inject pharmacological compounds into specific brain regions, such as the cerebellum or striatum, or to systemically attempt to reproduce an overt dystonic phenotype.7,11, 12 The advantage of this approach is that it has the possibility of the potential identification of a brain region, cell receptor, or signaling pathway that may contribute to dystonia onset. These models are also particularly useful in analyzing secondary dystonia caused by brain injury or from the side effects of drugs.13 However, one must consider that the primary goal of genotypic dystonia animal models is to mimic the genetic mutations found in patients, which can be accomplished by targeted mutagenesis or through insertional transgenic techniques. This type of model is particularly useful to analyze the pathophysiology of genetic dystonias by examining the normal function of the causal proteins and to analyze the mutated effects. The primary aim of this review is to highlight the findings of genotypic models of DYT1, DYT11 and DYT12 dystonias using worms (Caenorhabditis elegans), fruit flies (Drosophila melanogaster), and rodents, and to illustrate their utility in elucidating dystonia pathophysiology, and to uncover novel therapeutics.

2. DYT1 Dystonia

2.1 Background

DYT1 dystonia (OMIM 128100) is a primary, generalized, early-onset torsion dystonia.14 Patients typically present with symptoms in childhood, but can present as late as 26 years of age.15, 16 Typically, symptoms first present in the limbs and then over several years in many patients these symptoms become generalized.17 DYT1 dystonia is an autosomal-dominant disorder with a reduced and incomplete penetrance of 30-40%. It is caused by a trinucleotide (GAG) deletion in the DYT1/TOR1A gene, which encodes the ubiquitously expressed torsinA protein.18 This mutation removes one of a pair of glutamic acid residues from the C-terminal region of the torsinA protein. This mutation is commonly referred to as either DYT1 ΔGAG or torsinAΔE. TorsinA is a member of the AAA+ superfamily (ATPases associated with diverse cellular activities), and has been implicated in a variety of cellular processes, including chaperone-mediated protein folding and vesiclular trafficking.19, 20

2.2 Invertebrate models of DYT1 dystonia

Worms (C. elegans) and fruit flies (D. melanogaster) have a short generation time and a well-defined nervous system and these properties make them very good model organisms to study neurological disorders.21-23 C. elegans have three torsin related genes: tor-1, tor-2, and ooc-5. All three genes encode ATPases that are putative proteins for the Clp/Hsp100 and AAA+ superfamilies.24, 25 Caldwell and colleagues conducted in vivo assays in C. elegans to investigate the effect of these torsin proteins on polyglutamine-induced protein aggregation.26 First, they observed that human torsinA expression, wild-type TOR-2 overexpression, or TOR-2 and OOC-5 co-expression these approaches were able to reduce protein aggregation in C. elegans.26 However, expression of a mutated TOR-2 had lost its ability to reduce the protein aggregation (Table 1).26 In another study, overexpression of human torsinA or TOR-2 in C. elegans, specifically in dopaminergic neurons, was able to protect dopaminergic neurons from 6-hydroxydopamine (6-OHDA) induced neuronal loss, possibly through down regulation of the dopamine transporter.27 However, this protection was decreased when either mutant torsinA or mutant TOR-2 proteins were expressed.27 Lastly, human torsinA was able to suppress endoplasmic reticulum (ER) stress response in C. elegans, both at baseline and in response to the protein glycosylation inhibitor tunicamycin, which induces ER stress.28 However, in the presence of mutant human torsinA, baseline ER stress was found to be increased.28

Table 1. Invertebrate models of DYT1 dystonia.

Model Important findings
C. elegans containing fluorescent polyglutamine-
induced protein aggregates.
↓ Polyglutamine-induced protein aggregation when overexpressing human torsinA,
TOR-2, and co-expressing of TOR-2 and OOC-5.
TOR-2 (Δ368) mutant (DYT1 associated mutation) was not able to reduce protein
aggregation.26
C. elegans (dopamine-specific neuronal promoter) ↑ Resistance of DA neurons to 6-OHDA by overexpression of human torsinA or worm
TOR-2.
↓ Resistance of DA neurons to 6-OHDA by expression of torsinAΔE, mutant TOR-2
proteins, or by co-expression of wild-type torsinA and torsinAΔE27
C. elegans (ges-1 intestinal promoter) ↓ ER stress response at basal and when stress was induced by tunicamycin with wild
type torsinA expression.
↑ ER stress response at basal conditions with torsinAΔE expression, and by co-
expression of wild-type torsinA and torsinAΔE, but no significant change in response to
tunicamycin.28
D. melanogaster Gal4/UAS system for overexpression
of torsinAΔE in the nervous system and muscle
Protein aggregation by torsinAΔE.
Impaired locomotion after exposure to 38°C.
Ultrastructure alterations at the neuromuscular junction in larvae.29
D. melanogaster Gal4/UAS system for overexpression
or down-regulation of dtorsin
Downregulation or overexpression of dtorsin caused subtle rough eye phenotypes
(neurotoxicity) and abnormal eye pigment granule configuration in young flies.
Downregulation of dtorsin caused retinal degeneration in aged flies.
Overexpression of dtorsin caused protection from neural degeneration.
dtorsin localizes to nuclear envelope and ER.33
D. melanogaster Gal4/UAS system for overexpression
of torsinAΔE, and ΔFY HtorA in neurons or muscles
Protein aggregation in neurons and muscles by torsinAΔE, but absent in ΔFY HtorA
mutants.
Impaired locomotion after exposure to 38°C in flies expressing torsinAΔE, and ΔFY
HtorA.
Ultrastructure alterations at the neuromuscular junction in larvae, and synapses.31
D. melanogaster null mutant for dtorsin Prepulpal semi-lethality, sterility and altered locomotion behavior
↓ Dopamine levels in brain.
↓GTP cyclohydrolase protein and activity34

Abbreviations: 6-OHDA (6-hydroxydopamine), ER (endoplasmic reticulum), GTP (Guanosine-5′-triphosphate), TOR-2(Δ368) (lacking a codon for amino acid 368), torsinAΔE (human mutant torsinA with a GAG deletion), ΔFY HtorA (human mutant torsinA with an 18 bp deletion).

When human torsinAΔE was expressed either in the muscle or neurons of fruit flies, temperature dependent locomotion deficits were observed.29 Additionally, protein aggregation of mutant torsinA was observed at normal temperature.29 Fruit flies expressing an 18 bp deletion in human DYT1/TOR1A, a mutation reported in a family with early onset dystonia,30, 31 showed similar motor deficits after exposure to 38°C. Additionally, abnormalities in synapses and larva neuromuscular junction were observed in these fruit flies.31 Fruit flies only have a single torsin gene, dtorsin.32 Down-regulation of dtorsin results in increased neuronal degeneration.33 Furthermore, a null mutation of dtorsin will result in semi-lethality, sterility, locomotion deficits, and decreased dopamine levels in the brains of the larva and adult heterozygotes.34

These invertebrate studies have been integral in identifying the function of torsinA, such as a role chaperone-mediated protein folding, protection of dopaminergic neurons from neurotoxicity, and ER stress, and the effects of modulating expression or introducing mutations on behavior. These findings taken together support the continued use of invertebrate animal models to study DYT1 dystonia, and in particular the role of cellular functions.

2.3 Rodent Models of DYT1 dystonia

In mammals, there are four genes in the torsin family: torsinA, torsinB, torsin2A, and torsin3A. The human DYT1/TOR1A gene and rodent Dyt1/Tor1a genes, in particular mice and rats, are highly homologous, and therefore they are excellent model organisms.

2.3.1 Dyt1 ΔGAG knock-in, knockout, and knockdown mouse models of DYT1 dystonia

Multiple DYT1 genotypic dystonia models have been generated using rodents (Table 2). The most significant model is the Dyt1 ΔGAG heterozygous knock-in (KI) line of mice, as this line of mice recapitulates the trinucleotide deletion in Dyt1/Tor1a that is most often seen in DYT1 dystonia patients. Similar to humans, this deletion results in a loss of one of a pair of glutamic acid residues in torsinA.35, 36

TABLE 2. ES cell based dystonia models.

Mouse Microdialysis Monoamines Locomotion Rotarod Beam Walking Gait Analysis Neuronal Phenotype Additional findings
dYT1 Rodent Models Dyt1 AGAG KI ↓ DA at baseline and
by AMPH in
striatum(~2.3 months
male/female)39
↓ HVA in striatum (11 months-
male)35

↑ HVA in striatum
(10 months)39

↓ DA in midbrain (6 and 10 months
male/female)39
↑ Activity
(6 months- male)35

↑ Daytime motility
(3 and 6 months
male/female)39
Not affected35
(6months
male/female)

↓ Motor skills
(3,6,10 months
male/female)39
↑ Slips
(6months- male)35

↑ Transversal time and
slips
(3 months)
(male/female)39
↑ Paw overlap
distance
(6months-male)35

Affected gait
(male/female)39
Abnormal nuclear
envelope in CNS
(Dyt1 ΔGAG
homozygous KI and KO
embryonic day 15,18)36

Abnormal nuclear
envelope of cerebral
cortical neurons
(Dyt1 ΔGAG homozygous
KI-newborn)56

Inclusions for ubiquitin
and torsinA in brain stem
(6months-male)35

Enlarged and reduced
TH positive cells in
substantia nigra (3-6
months male/female)39

Shorter primary dendrites
(8 months)and reduced
spine numbers in
Purkinje cells (3-
8months)62
Microstructural abnormalities in
cerebellothalamocortical pathway (4
months)42

↓ Corticostriatal LTD (3-4 weeks-male)40

↓ D2R binding40

↓ D2R in striatum (6 months- male)40

↑ Synaptic vesicle recycling46

↑ Glutamate mEPSC frequency47
( hippocampus cultured neurons from
heterozygous and homozygous Dyt1 ΔGAG
KI mice postnatal Days 0–1)

↓ torsinA in brain53

↓ torsinA in striatum48
Dyt1 KD mice NA ↓ DOPAC in striatum (8–9 months-
male)51
↑ Horizontal activity
(6 months-male)51
Not affected
(6 months-
male/female)51
↑ Slips
(6 months-male)51
↓ Hind base
(male)52
NA
Purkinje cell-
specific Dyt1
conditional KO
NA NA NA Not affected
(~5.2-7.7months-
male/female)53
↓ Slips (~5.6-8 months-
male/female)53

No difference in slip
numbers between CT
and Dyt1 ΔGAG KI/Dyt1
pKO double mutant mice
(~5.3-5.7months-
male/female)53
Normal gait
(~5.4-7.8 months-
male/female)53
Shorter primary dendrites
(8 months)and reduced
spine numbers (3-8
months) in Purkinje
cells62

Normal nuclear envelope
in cerebellar cells
~2-4 months
male/female)53
Cortex-specific
Dyt1 conditional
KO
NA Not affected in striatum (~8.7-11.3
months-male/female)52
↑ Activity (~3.8-6.5
months male/female)52
Not affected (~3.8-
6.5 months
male/female)52
↑ Slips
(~3.8-6.3 months
male/female)52
↓ Hind base (~5.4-
8 months male)52
Normal nuclear envelope
in cerebral cortical
neurons
(8 weeks)56
Striatum-
specific Dyt1
conditional KO
NA Not affected in striatum (~10-11.3
months- male/female)56
Not affected
(~4-5.5 months-
male/female)56
Not affected (~ 3.4-
4.8 months-
male/female)56
↑ Slips
(~4.6-6 months-
male/female)56
NA Normal nuclear
envelope in striatal
medium spiny neurons (6
weeks)56
↓ D2R binding in striatum
(~8.8-10.5 months)56
Cholinergic-
specific Dyt1
conditional KO
mice
Normal DA at
baseline and by
AMPH in striatum
(7 months )58
NA Not affected
(3-8 months-
male/female)58
↓ Latency to fall
(8-11 months-
male/female)58
Not affected
(8-11 months-
male/female)58
NA NA Normal striatal acetylcholine content,
uptake and release
abnormal D2R response and altered
muscarinic receptor function (8-10 weeks)58
DYT1 dystonia models Sgce KO mice Normal DA at
baseline

↑ DA release by
AMPH in striatum (5
months)96
↑ DA, DOPAC, HVA in striatum
(~11-11.7 months-male/female)87
↑ Horizontal activity
(female)

↑ Vertical movement
number, movement time,
activity (male/ female)87
Not affected87 ↑ Slips, impaired motor
learning (~6.5-7.5 months
male/female)87
Not affected87 Abnormal nuclear
envelope in striatum
(2-5.5 months)93

Abnormal nuclear
envelopes in cerebellar
Purkinje cells (2-4
months)94
Myoclonus ( ~7.2-8 months male/female)87

Anxiety-like behavior (~7.3-8.2 months
male)87

Depression-like behaviors (~7.8-8.6 months
female)87

↓ D2R,normal D1R and DAT in striatum (5
months)96
Striatum-
specific Sgce
conditional KO
mice
NA NA Not affected
(~5.2-8 months
male/female)93
↓ Latency to fall in
trials 3 and 5

↓ Total latency to
fall (~5.7-8.3
months
male/female)93
↑ Slips
(~5.5-8 months-
male/female)93
NA Normal nuclear envelope
in striatum93
No myoclonus(~6.3-9 months-
male/female).93
Purkinje cell-
specific Sgce
conditional KO
mice
NA NA ↓Stereotypic activity,
movement numbers,
movement time
(~6.2-7.5 months-
male/female)94
↓ Latency to fall on
4th trial
(~7-8 months
male/female)94
Impaired motor learning
(~6.6-8 months-
male/female)94
NA Normal nuclear envelope
in cerebellar Purkinje
cells94
No significant myoclonus (~8.7-10 months-
male/female)94
DYT1 and
DYT11 Dystonia
model
Dyt1 and Sgce
double mutant
 mice
NA NA ↓ Horizontal movement
numbers compared to
DytlΔGAG KI (~4.4-5.3
months-male/female)48
NA ↑ Slips (5.5 months-
male/female)48
NA Normal nuclear envelope
in cortical neurons(5.5-
months-male/female)48
↓ torsinA in striatum (5.5 months
male/female)48
DYT12 Dystonia
model
Heterozygous
Atp1a3 KO mice
NA Not affected, but individual
neurotransmitter levels correlated to
open field activity levels showed that
stressed mice had changes in the
dopaminergic and serotonergic
systems106
↓ Clockwise (CW)
circling (stressed
6months male-female)
Voluntary activity levels
by wheel running: one
group hyperactive, the
other group hypoactive
(6months- stressed
female)106
↓ Latency to fall
(stressed 6
months- female)106
↑ Slips (stressed 6
months- female)106
NA NA ↓ Sensitivity to warm stimuli (stressed
female)106

Abbreviations: AMPH (amphetamine), CNS (central nervous system), CT(control littermate), DA (dopamine), DAT (DA transporter), DOPAC (3, 4-dihydroxyphenylacetic acid), Dyt1ΔGAG KI (heterozygous mutant Dyt1 gene with a GAG deletion), D1R D2R (dopamine receptor type 1 and 2), HVA (homovanillic acid), KD (knockdown), KI (knock-in), KO (knockout), LTD (long term depression), NA (Not analyzed),TH (tyrosine hydroxylase).

The Dyt1 KI mouse exhibited motor deficits and abnormal gait, which while not overt dystonia, is thought to represent a dystonia-like phenotype.35 Dyt1 KI mice also displayed a subtle anxiety-like behavior and enhancement of cued fear memory.37 These results were similar to DYT1 dystonia mutation carriers, who exhibited increased anxiety, verbal memory retroactive interference, and higher semantic fluency performance.38

Additionally, Dyt1 KI mice showed reduced release of striatal dopamine at baseline and after amphetamine stimulation.39 Furthermore, a reduction in striatal D2 dopamine receptor (D2R) binding was observed in the Dyt1 KI mice,40 which was consistent with post-mortem and in vivo PET imaging studies that revealed a reduction in D2R binding in human DYT1 dystonia mutation carriers.41

Lastly, Dyt1 KI mice exhibited cerebellothalamocortical (CbTC) pathway abnormalities using PET and DTI imaging techniques.42 This was similar to the white-matter alterations in the sensorimotor cortex43 and the superior cerebellar peduncle44 in DYT1 dystonia and primary dystonia patients, which suggest alterations in CbTC tract integrity.44, 45 Furthermore, the KI mice have shown alterations in corticostriatal long-term depression (LTD).40 Neurotransmission deficits were also reported in cell culture of hippocampal neurons from Dyt1 KI mice.46, 47

Dyt1 KI mice had reduced torsinA protein levels in the brain,36, 48 which was consistent with the accelerated degradation of mutant torsinA found in cultured cells.49, 50 Therefore, to determine whether a loss-of-function mutation of torsinA leads to DYT1 dystonia, Dyt1 knockdown (KD) mice were developed. These mice had approximately 36% reduction of torsinA protein, and also Dyt1 knockout mice were similarly developed. Similar to Dyt1 KI mice, Dyt1 KD mice displayed motor deficits, increased locomotor activity, and alterations of striatal dopamine metabolism51. In contrast, Dyt1 homozygous knockout (KO),36, 52 Dyt1 homozygous KI,35, 36 and KO/KI double mutant mice53 resulted in neonatal lethality, suggesting that a gross loss of torsinA function impaired normal development in mice. These findings are to date, consistent with no report of human carriers with mutations in both DYT1 alleles.

2.3.2 Brain-region specific Dyt1 conditional knockout mice

Brain-region specific Dyt1 conditional knockout mice have been used to understand how specific brain region or cell types contribute to the pathophysiology of the disease (Table 2). When the phenotypes of conditional knockout mice match with those of Dyt1 KI mice, it is proposed that the corresponding regions or cells have relevance to the pathophysiology. The cerebral cortex-specific Dyt1 conditional knockout (Dyt1 cKO) mice, were generated by crossing Emx1-cre54 and Dyt1 loxP mice, and exhibited motor deficits and hyperactivity.52 The striatum-specific Dyt1 conditional knockout (Dyt1 sKO) mice, generated by crossing Rgs9L-cre55 and Dyt1 loxP mice, exhibited motor deficits and reduced striatal D2R binding activity.56 However, both cKO52 and sKO56 mice showed no significant alteration in striatal monoamine levels. Next, cholinergic neuron-specific Dyt1 conditional knockout mice, generated by crossing ChAT-cre57 and Dyt1 loxP mice, showed motor deficits. Furthermore, striatal cholinergic interneurons in these mice showed alterations in response to muscarinic receptor activation and D2R receptor activation, but no change in response to either GABAA or metabotropic glutamate receptor activation.58 The cerebral cortex and striatum, along with cholinergic innervation seemed to be important components of the basal ganglia circuitry. These studies suggested that loss of torsinA function in these areas resulted in motor deficits and these findings may prove important to better understanding the pathogenesis of dystonia.

Another aspect of this puzzle that has yet to be investigated is using conditional knockout techniques to show the contribution of the direct and indirect pathways of the basal ganglia. Striatal medium spiny neurons (MSNs) expressing D1 dopamine receptor mediate the direct pathway, which is thought to be involved in the initiation of movement, while striatal MSNs expressing D2 dopamine receptor mediate the indirect pathway, and this may possibly prevent unwanted movements. Therefore, the motor deficits in DYT1 dystonia may be mediated, at least in part, through changes in D2R function in the basal ganglia circuit. The relative contributions of presynaptic D2R on cholinergic interneurons and postsynaptic D2R on medium spiny neurons to the pathophysiology of DYT1 dystonia remain to be determined.

Several studies have suggested the role of the cerebellum in the pathophysiology of dystonia. For instance, cerebellectomies of either dystonic (dt) rats or tottering mutant mice were able to improve their dystonic-like symptoms.7, 59 Moreover, crossing the tottering mutant mice with pcd mutant mice, which have Purkinje cell-specific degeneration, were similarly able to improve in their dystonic-like symptoms.60 Therefore, Purkinje cell-specific Dyt1 conditional knockout (Dyt1 pKO) mice were generated by crossing Pcp2-cre61 with Dyt1 loxP mice. These mice had alterations in Purkinje cell dendritic morphology and showed an improvement in motor performance compared to wild-type mice.62,53 Since Purkinje cells integrate cerebellar signals and the cerebellum modulates the basal ganglia circuits,63 loss of torsinA function in Purkinje cells may balance the abnormal basal ganglia circuits and attenuate the dystonic symptoms.

2.3.3 Transgenic rodent models of DYT1 dystonia

A line of transgenic mice overexpressing human torsinAΔE driven by the neuron-specific enolase promoter displayed self-clasping of hind limbs, hyperkinesia, altered circling behavior, abnormal gait, brain stem pathology, and disrupted striatal dopamine levels.64, 65 A second line of transgenic mice were generated using the human cytomegalovirus (hCMV) immediate early promoter in order to drive the overexpression of human torsinAΔE, and these have been referred to as hMT mice. The hMT mice exhibited motor learning deficits,66 motor deficits,67 and abnormal gait.67 Furthermore, there was an increase in dopamine turnover,67 altered dopamine release, decreased basal locomotion induced by amphetamine,68 and decreased dopamine transporter function.69

The hMT mice have been extensively studied using electrophysiological techniques. The hMT mice exhibited defected LTD, an enhanced long-term potentiation (LTP), and a deficit of synaptic depotentiation (SD).70 A shortened pause response by thalamic stimulation in cholinergic interneurons was also found, due to an altered D2R function affecting the synaptic convergence between thalamostriatal and corticostriatal responses.71 Furthermore, the hMT mice exhibited an excitatory, instead of inhibitory, striatal cholinergic interneuron response after dopamine D2R activation,72, 73 and also had a decreased D2R receptor level.74 Additionally, electrophysiological studies revealed that there was a critical alteration of striatal dopamine D2R-mediated function both in cholinergic interneurons as well as in the control of GABAergic synaptic transmission in MSNs of the hMT mice.72, 75These observations are compatible with current theories on the pathogenesis of dystonia, suggesting that both an increased propensity to “potentiation” (LTP-like) and a failure of “depression” mechanisms (LTD- and SD-like) lead to a “loss of inhibition” in the motor system which might, at least in part, explain the pathogenesis of the excess of abnormal movements observed in dystonic patients, though this remains speculative.9

A third line of transgenic mice overexpressing human torsinAΔE in dopaminergic neurons of the midbrain resulted in motor deficits and altered dopamine release in response to cocaine. These findings suggested that disruption of torsinA activity by overexpressing the mutant torsinA in dopaminergic neurons affected dopamine transmission. 76

A fourth line of transgenic mice was generated using a murine prion protein promoter, and this line of mice revealed that overexpressing wild-type or torsinAΔE caused inclusion bodies predominantly in the brainstem, nuclear envelope abnormalities, altered monoamine levels, and motor deficits.77 However, both the wild-type and torsinAΔE were expressed as fusion proteins with a carboxyl terminal attachment of a V5-His tag. The functional consequences of these fusion proteins have not been well characterized. Realizing the deficiency of the approach, the same research group generated transgenic rats expressing human torsinAΔE from the human torsinA promoter. The mutant rats showed altered synaptic plasticity, motor deficits, and nuclear membrane alterations.78

Various transgenic rodent models overexpressing wild-type torsinA or mutant torsinAΔE using different promoters have been produced (Table 3). Most of the transgenic rodent models showed an impairment of motor behavior. However, investigators should be cautious when interpreting these results because the observed behavioral and cellular abnormalities could have been a result of non-physiological and ectopic protein expression. Furthermore, it was difficult to generate proper control animals for these transgenic mice due to differences in the transgene insertion site, copy number, expression level, and pattern of expression.

Table 3. Transgenic DYT1 dystonia rodent models.

Promoter Microdialysis Monoamines Locomotion Rotarod Beam Walking Gait Analysis Neuronal phenotype Additional findings
Rat neuron-specific
enolase
NA ↓ DA in striatum (mice with
affected phenotype) 64

↓ DOPAC/DA in striatum
(mice with unaffected and
affected phenotype)
(1.5months)64
Self-clasping,
hyperkinesia,
circling (mice with
affected phenotype
(1.5-months)64

↓ Rearing
(3 months)65
Not affected
(3,6,9,12
months)65
NA ↓ Forelimb and
hind limb stride
length
( 9 months)65
Perinuclear inclusions for
ubiquitin and TorsinA in PPN
and PAG64
Transgene expression lost after repeated breeding65
Human
cytomegalovirus (CMV)
Not affected DA at
baseline in
striatum68

↓ DA by AMPH in
striatum (6months-
male)68

↓ DAT activity in
striatum (6-
months-male)69
Not affectedin striatum (6
months-male)68

↑ DOPAC/DA and HVA/DA in
striatum (3-4months- male)67
Not affected
(9months-male)66

↓ Basal locomotion
activity and by
AMPH
(6months- male)69
↓ Motor learning
(9 months-
male)66

Not affected (3-4
months male)67
↑ Transversal time
and slips
(3-4months-male)67
↑ Hind base width
(3-4 months-
male)67
No torsinA or ubiquitin inclusions
in striatum (male)66

No torsinA or ubiquitin inclusions
No bleb formation in nuclear
envelope (male)67
Abnormal response to D2 dopamine receptor
stimulation of Chls (9 months)72

Normal DAT, VMAT2, striatal DA receptors density (6
months-female)68

Impaired LTD and SD in medium spiny neurons70

↑ Striatal acetylcholinesterase activity (~2months)70

Altered GABAergic function in striatum (~2 months)75

↓ D2R, ↓ RGS9 in striatum (~2 months)74

↓Pause response by thalamic stimulation in Chls
(male)71
Murine prion NA htorsinAΔE mice:
↑ DOPAC, 5-HT, and 5-HIAA
in brain stem

htorsinA mice:
↓DA, 5-HT
and 5-HIAA in striatum
(5 months- male/female)77
htorsinAΔE mice:
↑ Activity (6months-
male/female)77

htorsinA mice:

↓ Activity
htorsinAΔE:
↓ Latency to fall,
↓ Learning
(6months-
ale/female)77

htorsinA mice:
Not affected
htorsinAΔE:
Not affected
(6months-
male/female)77

htorsinA mice:
↑ Transversal time
htorsinAΔE:
Not affected
(6months-
male/female)77

htorsinA mice:
↓ Mean stride
length for forelimbs
and hind limbs
htorsinAΔE and htorsinA:
torsinA or ubiquitin inclusions
mainly in brainstem,
nuclear envelope disruption in
brainstem and striatum
(1.5months male)77
Altered motor circuits integrity by diffusion tensor
imaging in cortex, striatum, cerebellum (5months -male)77
Tor1A promoter
(rat model)
NA NA NA ↓ Latency to fall
at training
sessions
( 2 months -
male)78
↑ Transversal time
3,5 cm round beam
(2months-male)78
↑ Variation of step
length (hind limb)
(11months male- )78
Abnormal protein localization of
human torsinAΔE to nuclear
envelope (cortex, hippocampus,
striatum, olfactory bulb,
substantia nigra)(2 months-
male)78

Altered nuclear envelope
structure in cortex and striatum
( 3 months- male)78
Limb grasping and clasping (11 months-male)

Impaired LTD and SD in medium spiny neurons (8-12
weeks old-male)78
Human TH promoter
fragment (transgene
expression in
dopaminergic neurons
of midbrain)
TH- htorsinAΔE:
Not affected DA at
baseline and ↓ DA
by cocaine in
striatum (male)

TH-htorsinA mice:
↑ DA release by
cocaine in
striatum76
TH- htorsinAΔEmice:
Normal DA and DOPAC in
striatum

TH-hTorsinA mice:
↓ DA and ↓ DOPAC in
striatum
(2.5 months-female)76
TH- htorsinAΔE
mice:
Normal baseline
locomotion

TH-htorsinA mice:
↓ Baseline
locomotion
 (4-10months
male)76
Not affected
(2,4,6 months
male/female)76
TH- htorsinAΔE
mice:
↑ Transversal time
and slips
male/female)76
NA NA

Abbreviations: 5-HIAA (5-hydroxyindoleacetic acid), 5-HT (5-hydroxytryptophan), AMPH (amphetamine), Chls (cholinergic interneurons), DA (dopamine), DAT (DA transporter), DOPAC (3, 4-dihydroxyphenylacetic acid), D2R ( dopamine receptor type 2), HVA (homovanillic acid), htorsinAΔE (human wild-type torsinA), htorsinAAE (human mutant torsinA with a GAG deletion), LTD (long-term depression, NA (Not analyzed), PPN (pedunculopontine nucleus), PAG (periaqueductal gray), RGS9 (Regulator of G-protein signaling 9), SD (synaptic depotentiation),TH (tyrosine hydroxylase), VMAT2 (vesicular monoamine transporter).

3. DYT11 dystonia

3.1 Background

DYT11 dystonia (OMIM 159900) is the major subtype of myoclonus-dystonia (M-D), and is characterized by myoclonic jerks with dystonic symptoms.79, 80 Additionally, it is often accompanied by psychiatric symptoms, such as depression and anxiety disorders. DYT11 dystonia generally presents in childhood, but in some individuals can present in late adulthood.81 It is caused by mutations in SGCE, which encode the transmembrane glycoprotein ε-sarcoglycan.79 SGCE is maternally imprinted and paternally expressed.82 Furthermore, it has been demonstrated that there is exclusive paternal expression of ε-sarcoglycan in the brains of mice83 and humans.84

3.2 Rodent models of DYT11 dystonia

Epsilon-sarcoglycan was first identified in mice as a homolog of α-sarcoglycan.85, 86 Two lines of Sgce knockout mice have been reported. The first lacks exon 4, which results in a frame-shift mutation and has exhibited significant relevant phenotypes.83, 87 Another mouse line was generated and it lackedexons 6 through 9, which did not result in a frame-shift mutation, and did not exhibit an overt phenotype.88 We will focus on the findings from the first line of mice.

3.2.1 Sgce heterozygous KO mice

Sgce heterozygous knockout (KO) mice lacking exon 4 were generated by crossing Sgce loxP83 with CMV-cre89 mice (Table 2). Paternally-inherited Sgce KO mice showed myoclonus and deficits in fine motor coordination and balance, motor learning in the beam-walking test, and anxiety and depression-like behaviors.87 The motor learning deficit was recently demonstrated in DYT11 dystonia patients with impaired saccadic adaptation,90 and this was thought to be a form of cerebellar motor learning.91 Sgce KO mice also exhibited abnormal nuclear envelopes in striatal neurons and cerebellar Purkinje cells, suggesting that DYT11 dystonia belongs to a growing family of nuclear envelopathies.92-94 Additionally, loss of ε-sarcoglycan did not cause a reduction of other sarcoglycan isoforms (α, δ, and ζ), suggesting that ε-sarcoglycan did not make a sarcoglycan complex similar to other sarcoglycans.93 Furthermore, the levels of striatal dopamine and its metabolites in Sgce KO mice were significantly increased.87 The hyperdopaminergic state seemed to be potentially further strengthened by evidence that DYT11 dystonia patients had a reduction in striatal D2R binding.95 Similarly, Sgce KO mice exhibited reduced striatal D2R protein levels and an increased dopamine release after amphetamine administration.96 In conclusion, these studies suggested that DYT11 dystonia had functional alterations of the monoamine system in the striatum, and that the Sgce KO mice reasonably modeled DYT11 dystonia.

3.2.2 Brain-region specific Dyt1 conditional knockout mice

Brain region-specific Sgce knockout models have been generated to dissect brain circuits involved in the pathogenesis of DYT11 dystonia. Paternally-inherited striatum-specific Sgce conditional knockout (Sgce sKO) mice exhibited motor deficits, but no myoclonus and also had normal nuclear envelopes in the striatum.93 Paternally-inherited cerebellar Purkinje cell-specific Sgce conditional knockout (Sgce pKO) mice showed motor learning deficits, but no myoclonus and a normal nuclear envelope in the cerebellar Purkinje cells.94 The results suggested that loss of ε-sarcoglycan function in the striatum and in the cerebellar Purkinje cells contributed to motor deficits and motor learning deficits observed in the complete Sgce KO mice.87 In contrast, loss of ε-sarcoglycan function in these regions alone did not contribute to myoclonus or nuclear envelope abnormalities, suggesting that other brain regions or a combination of these brain regions may contribute to these phenotypes.

4. DYT12 dystonia

4.1 Background

DYT12 dystonia (OMIM 128235), or rapid-onset dystonia-parkinsonism (RDP), is characterized by symptoms of both dystonia and parkinsonism, which include resting tremor, akinesia, bradykinesia, and postural instability.97 In DYT12 dystonia, the human patients’ symptoms can appear within minutes to a few days, and do not remit.97-101 The symptoms can be triggered by a physiological stressor, such as a high fever or pregnancy.98, 102 DYT12 dystonia is caused by missense mutations in the ATP1A3 gene, which encodes the Na+, K+-ATPase α3 isoform. The α3 isoform is only expressed in neurons and cardiac cells.97, 103 It is hypothesized that the loss of the α3 isoform during stress may interfere with the normal response to physiological stressors. These responses have been referred to as stress induced channelopathies, and this type of finding has been observed in other diseases, such as myasthenia gravis and chronic fatigue syndrome.8, 104

4.2 Rodent model of DYT12 dystonia

Na+, K+-ATPase α3 isoform-deficient mice were generated as a DYT12 dystonia model from gene targeting that resulted in aberrant splicing of the gene.105 The heterozygous Atp1a3 KO mice exhibited an increase in locomotor activity at baseline and in response to methamphetamine in the open-field test. Additionally, the heterozygous Atp1a3 KO mice showed impaired learning in the Morris water maze and decreased hippocampal NR1 subunit of NMDA receptor. More importantly, the heterozygous Atp1a3 KO mice had no motor deficits prior to stress induction, but following an immobilized stress protocol, female heterozygous Atp1a3 KO mice exhibited motor deficits in the rotarod and beam-walking tests.106 The stressed Atp1a3 KO mice also exhibited alterations in their sensory response to warm stimuli, circling behavior, and the monoamine neurotransmitter system. Moreover, in a separate study stress increased the susceptibility to depression-like phenotypes in Atp1a3 KO mice.107

5. Dystonia animal models and preclinical drug discovery

Currently there is no known cure for dystonia. However, physical therapy, medications, and surgery aim to lessen symptoms. Positive symptomatic outcomes have resulted from botulinum toxin treatment for focal dystonia,108 medications such as the anticholinergic trihexyphenidyl, and from the use of gamma-aminobutyric acid (GABA) derivatives for primary dystonia.109 Deep brain stimulation for generalized and cervical dystonia have also been shown effective in select cases.109 However, not all patients respond favorably to these treatments likely as a result of the heterogeneity of dystonia, and the differences in underlying pathophysiology. Therefore, there is tremendous need for new and more effective treatments for dystonia.

Although none of the genetic dystonia rodent models exhibit overt dystonic symptoms, the beam-walking and rotarod tests have successfully detected motor deficits in a majority of genotypic rodent models.35, 39, 48, 51, 56, 58, 66, 67, 76-78, 87, 93, 94, 106 Two particular cases worth noting are the DYT12 and DYT1 dystonia mouse models. In DYT12 dystonia patients, symptoms are often triggered by a stressor, resulting in a rapid progression of dystonic symptoms and parkinsonism. In the mouse model of DYT12 dystonia, there are no beam-walking deficits but when a stress protocol is used, the deficits emerge.106 Furthermore, the beam-walking deficits in Dyt1 KI mice, could be rescued using the anticholinergic trihexyphenidyl, a common human dystonia medication.40 We believe these two studies provide some preliminary evidence that the beam-walking test may be an indirect test for dystonia in genotypic rodent models though this will need further validation.

We posit that potential therapeutics aimed at treating genetic dystonias could possibly rescue beam-walking deficits observed in respective genotypic models. For example, an in vivo screen using transgenic C. elegans resulted in the identification of the antibiotic ampicillin as being able to enhance wild-type torsinA activity and rescuing beam-walking deficits and torsinA protein levels in the Dyt1 KI mice.110 The effect of ampicillin on DYT1 dystonia patients is now currently being investigated in a clinical trial (NCT01433757). In another case, loss of the torsinA protein in the Purkinje cells of the cerebellum (Dyt1 pKO) in combination with the global Dyt1 ΔGAG mutation (Dyt1 KI) in mice exhibited lower numbers of slips compared to Dyt1 KI mice on the beam-walking test.53 This finding indicated the possibility that the molecular lesions of torsinA in cerebellar Purkinje cells by gene therapy or intervention in the signaling pathway downstream of the cerebellar Purkinje cells may rescue motor symptoms in DYT1 dystonia patients. Additionally, further refinement of drugs targeting the corticostriatal pathway and its modulatory pathways will hopefully provide new treatments for DYT1 dystonia and other related dystonias. It is likely that other candidate tests like beam-walking will emerge as we better understand the animal models of dystonia.

6. Improving existing models for dystonia and future directions

One approach for the development of a better animal model would be gene targeting mediated by zinc finger nucleases (ZFNs), which are engineered proteins that bind to DNA at specific sites and produce double strand breaks in DNA.111 The advantage of ZFNs lies in targeting efficiency and the possibility of creating gene modifications from different organisms,111 omitting the need for producing embryonic stem cell based chimeric animal models.112 Since ZFN technology can be applicable to various organisms,113, 114 it can be useful to generate novel Dyt1 KI or Dyt1 knockout rats. Rats have larger bodies and brain sizes compared to mice, and this may facilitate easier in vivo electrophysiological, behavioral, and imaging studies.

Additionally, the engineered rodent dystonia models developed to date suffer from the lack of overt motor deficits. Future efforts should focus on developing models with early onset and pronounced symptoms. The process of generating such models will also provide insights into novel factors that contribute to onset and severity of symptoms. Recent studies have shown that Dyt1 homozygous KO mice lethality varied from 12 hours to 3 weeks depending on the genetic background.115 Therefore, it is possible that motor symptoms might also be influenced by genetic background. Additionally, it has been shown that additional gene mutations can affect the temporal onset of motor deficits, such as is the case of mutant mice harboring both the Dyt1 KI and Sgce KO mutations.48 Finally, stress induced motor deficits in the DYT12 dystonia mouse model suggest that environmental factors could also influence the onset and severity of the phenotype. Combinations of these approaches will hopefully lead to better mammalian models and facilitate the development of novel therapeutics.

7. Conclusion

Although all engineered rodent models have not revealed overt dystonic symptoms, motor deficits in these models may correspond to dystonic symptoms in humans. In addition to behavioral similarities between genotypic models of dystonia and patients, biochemical similarities also exist. For example, reductions in striatal D2R and/or its binding activities are common findings in both DYT1 and DYT11 dystonia mouse models. Analysis of these rodent models also demonstrates the pathophysiological changes in the corticostriatal, thalamostriatal, and cerebellothalamocortical pathways and understanding these changes will be critical to therapy development. It is our hope that further investigation of these pathways will lead to novel therapeutics to treat dystonia. Using a combination of invertebrate and rodent models, ampicillin, a common antibiotic, was shown to rescue motor deficits in animals and is a candidate for human trials. Moreover, Dyt1 pKO mice indicated that molecular lesions of torsinA in cerebellar Purkinje cells placed by gene therapy or alternatively intervening in the signaling pathways downstream of cerebellar Purkinje cells may both be able to rescue motor symptoms. These results indicate that the genetic animal models developed can be useful to further study the pathophysiology of dystonia and to attempt to develop novel therapeutics.

Acknowledgments

We would like to thank Michael S. Okun, M.D. for reviewing and editing this manuscript.

Financial Disclosure

Y.L. was supported by grants from the Tyler’s Hope for a Dystonia Care, Inc., the National Institutes of Health (NS37409, NS47466, NS47692, NS54246, NS57098, NS65273, NS72872, and NS74423), and startup funds from the Departments of Neurology at UAB and UF.

References

  • 1.Albanese A, Bhatia K, Bressman SB, et al. Phenomenology and classification of dystonia: A consensus update. Mov Disord. 2013 doi: 10.1002/mds.25475. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Fahn S, Eldridge R. Definition of dystonia and classification of the dystonic states. Adv Neurol. 1976;14:1–5. [PubMed] [Google Scholar]
  • 3.Bertram KL, Williams DR. Diagnosis of dystonic syndromes--a new eight-question approach. Nat Rev Neurol. 2012;8(5):275–283. doi: 10.1038/nrneurol.2012.39. [DOI] [PubMed] [Google Scholar]
  • 4.Niethammer M, Carbon M, Argyelan M, Eidelberg D. Hereditary dystonia as a neurodevelopmental circuit disorder: Evidence from neuroimaging. Neurobiol Dis. 2011;42(2):202–209. doi: 10.1016/j.nbd.2010.10.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Tinazzi M, Rosso T, Fiaschi A. Role of the somatosensory system in primary dystonia. Mov Disord. 2003;18(6):605–622. doi: 10.1002/mds.10398. [DOI] [PubMed] [Google Scholar]
  • 6.Zhuang P, Li Y, Hallett M. Neuronal activity in the basal ganglia and thalamus in patients with dystonia. Clin Neurophysiol. 2004;115(11):2542–2557. doi: 10.1016/j.clinph.2004.06.006. [DOI] [PubMed] [Google Scholar]
  • 7.Neychev VK, Fan X, Mitev VI, Hess EJ, Jinnah HA. The basal ganglia and cerebellum interact in the expression of dystonic movement. Brain. 2008;131(Pt 9):2499–2509. doi: 10.1093/brain/awn168. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Breakefield XO, Blood AJ, Li Y, Hallett M, Hanson PI, Standaert DG. The pathophysiological basis of dystonias. Nat Rev Neurosci. 2008;9(3):222–234. doi: 10.1038/nrn2337. [DOI] [PubMed] [Google Scholar]
  • 9.Quartarone A, Pisani A. Abnormal plasticity in dystonia: Disruption of synaptic homeostasis. Neurobiol Dis. 2011;42(2):162–170. doi: 10.1016/j.nbd.2010.12.011. [DOI] [PubMed] [Google Scholar]
  • 10.Sadnicka A, Hoffland BS, Bhatia KP, van de Warrenburg BP, Edwards MJ. The cerebellum in dystonia - help or hindrance? Clin Neurophysiol. 2012;123(1):65–70. doi: 10.1016/j.clinph.2011.04.027. [DOI] [PubMed] [Google Scholar]
  • 11.Calderon DP, Fremont R, Kraenzlin F, Khodakhah K. The neural substrates of rapid-onset Dystonia-Parkinsonism. Nat Neurosci. 2011;14(3):357–365. doi: 10.1038/nn.2753. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Ikeda K, Satake S, Onaka T, et al. Enhanced inhibitory neurotransmission in the cerebellar cortex of the Atp1a3-deficient heterozygous mice. J Physiol. 2013 doi: 10.1113/jphysiol.2012.247817. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Raike RS, Jinnah HA, Hess EJ. Animal models of generalized dystonia. NeuroRx. 2005;2(3):504–512. doi: 10.1602/neurorx.2.3.504. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Ozelius LJ, Hewett JW, Page CE, et al. The early-onset torsion dystonia gene (DYT1) encodes an ATP-binding protein. Nat Genet. 1997;17(1):40–48. doi: 10.1038/ng0997-40. [DOI] [PubMed] [Google Scholar]
  • 15.Bressman SB, de Leon D, Brin MF, et al. Idiopathic dystonia among Ashkenazi Jews: evidence for autosomal dominant inheritance. Ann Neurol. 1989;26(5):612–620. doi: 10.1002/ana.410260505. [DOI] [PubMed] [Google Scholar]
  • 16.Fahn S. The genetics of idiopathic torsion dystonia. Int J Neurol. 1991;25-26:70–80. [PubMed] [Google Scholar]
  • 17.Muller U. The monogenic primary dystonias. Brain. 2009;132(Pt 8):2005–2025. doi: 10.1093/brain/awp172. [DOI] [PubMed] [Google Scholar]
  • 18.Ozelius LJ, Hewett JW, Page CE, et al. The gene (DYT1) for early-onset torsion dystonia encodes a novel protein related to the Clp protease/heat shock family. Adv Neurol. 1998;78:93–105. [PubMed] [Google Scholar]
  • 19.Vale RD. AAA proteins. Lords of the ring. J Cell Biol. 2000;150(1):F13–19. doi: 10.1083/jcb.150.1.f13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Ogura T, Wilkinson AJ. AAA+ superfamily ATPases: common structure--diverse function. Genes Cells. 2001;6(7):575–597. doi: 10.1046/j.1365-2443.2001.00447.x. [DOI] [PubMed] [Google Scholar]
  • 21.Bargmann CI. Neurobiology of the Caenorhabditis elegans genome. Science. 1998;282(5396):2028–2033. doi: 10.1126/science.282.5396.2028. [DOI] [PubMed] [Google Scholar]
  • 22.Shulman JM, Shulman LM, Weiner WJ, Feany MB. From fruit fly to bedside: translating lessons from Drosophila models of neurodegenerative disease. Curr Opin Neurol. 2003;16(4):443–449. doi: 10.1097/01.wco.0000084220.82329.60. [DOI] [PubMed] [Google Scholar]
  • 23.Bilen J, Bonini NM. Drosophila as a model for human neurodegenerative disease. Annu Rev Genet. 2005;39:153–171. doi: 10.1146/annurev.genet.39.110304.095804. [DOI] [PubMed] [Google Scholar]
  • 24.Consortium CeS Genome sequence of the nematode C. elegans: a platform for investigating biology. Science. 1998;282(5396):2012–2018. doi: 10.1126/science.282.5396.2012. [DOI] [PubMed] [Google Scholar]
  • 25.Basham SE, Rose LS. The Caenorhabditis elegans polarity gene ooc-5 encodes a Torsin-related protein of the AAA ATPase superfamily. Development. 2001;128(22):4645–4656. doi: 10.1242/dev.128.22.4645. [DOI] [PubMed] [Google Scholar]
  • 26.Caldwell GA, Cao S, Sexton EG, Gelwix CC, Bevel JP, Caldwell KA. Suppression of polyglutamine-induced protein aggregation in Caenorhabditis elegans by torsin proteins. Hum Mol Genet. 2003;12(3):307–319. doi: 10.1093/hmg/ddg027. [DOI] [PubMed] [Google Scholar]
  • 27.Cao S, Gelwix CC, Caldwell KA, Caldwell GA. Torsin-mediated protection from cellular stress in the dopaminergic neurons of Caenorhabditis elegans. J Neurosci. 2005;25(15):3801–3812. doi: 10.1523/JNEUROSCI.5157-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Chen P, Burdette AJ, Porter JC, et al. The early-onset torsion dystonia-associated protein, torsinA, is a homeostatic regulator of endoplasmic reticulum stress response. Hum Mol Genet. 2010;19(18):3502–3515. doi: 10.1093/hmg/ddq266. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Koh YH, Rehfeld K, Ganetzky B. A Drosophila model of early onset torsion dystonia suggests impairment in TGF-beta signaling. Hum Mol Genet. 2004;13(18):2019–2030. doi: 10.1093/hmg/ddh208. [DOI] [PubMed] [Google Scholar]
  • 30.Leung JC, Klein C, Friedman J, et al. Novel mutation in the TOR1A (DYT1) gene in atypical early onset dystonia and polymorphisms in dystonia and early onset parkinsonism. Neurogenetics. 2001;3(3):133–143. doi: 10.1007/s100480100111. [DOI] [PubMed] [Google Scholar]
  • 31.Lee DW, Seo JB, Ganetzky B, Koh YH. DeltaFY mutation in human torsin A [corrected] induces locomotor disability and abberant synaptic structures in Drosophila. Mol Cells. 2009;27(1):89–97. doi: 10.1007/s10059-009-0009-z. [DOI] [PubMed] [Google Scholar]
  • 32.Ozelius LJ, Page CE, Klein C, et al. The TOR1A (DYT1) gene family and its role in early onset torsion dystonia. Genomics. 1999;62(3):377–384. doi: 10.1006/geno.1999.6039. [DOI] [PubMed] [Google Scholar]
  • 33.Muraro NI, Moffat KG. Down-regulation of torp4a, encoding the Drosophila homologue of torsinA, results in increased neuronal degeneration. J Neurobiol. 2006;66(12):1338–1353. doi: 10.1002/neu.20313. [DOI] [PubMed] [Google Scholar]
  • 34.Wakabayashi-Ito N, Doherty OM, Moriyama H, et al. Dtorsin, the Drosophila ortholog of the early-onset dystonia TOR1A (DYT1), plays a novel role in dopamine metabolism. PLoS One. 2011;6(10):e26183. doi: 10.1371/journal.pone.0026183. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Dang MT, Yokoi F, McNaught KS, et al. Generation and characterization of Dyt1 DeltaGAG knock-in mouse as a model for early-onset dystonia. Exp Neurol. 2005;196(2):452–463. doi: 10.1016/j.expneurol.2005.08.025. [DOI] [PubMed] [Google Scholar]
  • 36.Goodchild RE, Kim CE, Dauer WT. Loss of the dystonia-associated protein torsinA selectively disrupts the neuronal nuclear envelope. Neuron. 2005;48(6):923–932. doi: 10.1016/j.neuron.2005.11.010. [DOI] [PubMed] [Google Scholar]
  • 37.Yokoi F, Dang MT, Miller CA, et al. Increased c-fos expression in the central nucleus of the amygdala and enhancement of cued fear memory in Dyt1 DeltaGAG knock-in mice. Neurosci Res. 2009;65(3):228–235. doi: 10.1016/j.neures.2009.07.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Balas M, Peretz C, Badarny S, Scott RB, Giladi N. Neuropsychological profile of DYT1 dystonia. Mov Disord. 2006;21(12):2073–2077. doi: 10.1002/mds.21070. [DOI] [PubMed] [Google Scholar]
  • 39.Song CH, Fan X, Exeter CJ, Hess EJ, Jinnah HA. Functional analysis of dopaminergic systems in a DYT1 knock-in mouse model of dystonia. Neurobiol Dis. 2012;48(1):66–78. doi: 10.1016/j.nbd.2012.05.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Dang MT, Yokoi F, Cheetham CC, et al. An anticholinergic reverses motor control and corticostriatal LTD deficits in Dyt1 ΔGAG knock-in mice. Behav Brain Res. 2012;226(2):465–472. doi: 10.1016/j.bbr.2011.10.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Asanuma K, Ma Y, Okulski J, et al. Decreased striatal D2 receptor binding in non-manifesting carriers of the DYT1 dystonia mutation. Neurology. 2005;64(2):347–349. doi: 10.1212/01.WNL.0000149764.34953.BF. [DOI] [PubMed] [Google Scholar]
  • 42.Uluğ AM, Vo A, Argyelan M, et al. Cerebellothalamocortical pathway abnormalities in torsinA DYT1 knock-in mice. Proc Natl Acad Sci U S A. 2011;108(16):6638–6643. doi: 10.1073/pnas.1016445108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Carbon M, Kingsley PB, Su S, et al. Microstructural white matter changes in carriers of the DYT1 gene mutation. Ann Neurol. 2004;56(2):283–286. doi: 10.1002/ana.20177. [DOI] [PubMed] [Google Scholar]
  • 44.Carbon M, Ghilardi MF, Argyelan M, Dhawan V, Bressman SB, Eidelberg D. Increased cerebellar activation during sequence learning in DYT1 carriers: an equiperformance study. Brain. 2008;131(Pt 1):146–154. doi: 10.1093/brain/awm243. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Argyelan M, Carbon M, Niethammer M, et al. Cerebellothalamocortical connectivity regulates penetrance in dystonia. J Neurosci. 2009;29(31):9740–9747. doi: 10.1523/JNEUROSCI.2300-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Kakazu Y, Koh JY, Ho KW, Gonzalez-Alegre P, Harata NC. Synaptic vesicle recycling is enhanced by torsinA that harbors the DYT1 dystonia mutation. Synapse. 2012;66(5):453–464. doi: 10.1002/syn.21534. [DOI] [PubMed] [Google Scholar]
  • 47.Kakazu Y, Koh JY, Iwabuchi S, Gonzalez-Alegre P, Harata NC. Miniature release events of glutamate from hippocampal neurons are influenced by the dystonia-associated protein torsinA. Synapse. 2012;66(9):807–822. doi: 10.1002/syn.21571. [DOI] [PubMed] [Google Scholar]
  • 48.Yokoi F, Yang G, Li J, DeAndrade MP, Zhou T, Li Y. Earlier onset of motor deficits in mice with double mutations in Dyt1 and Sgce. J Biochem. 2010;148(4):459–466. doi: 10.1093/jb/mvq078. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Giles LM, Chen J, Li L, Chin LS. Dystonia-associated mutations cause premature degradation of torsinA protein and cell-type-specific mislocalization to the nuclear envelope. Hum Mol Genet. 2008;17(17):2712–2722. doi: 10.1093/hmg/ddn173. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Gordon KL, Gonzalez-Alegre P. Consequences of the DYT1 mutation on torsinA oligomerization and degradation. Neuroscience. 2008;157(3):588–595. doi: 10.1016/j.neuroscience.2008.09.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Dang MT, Yokoi F, Pence MA, Li Y. Motor deficits and hyperactivity in Dyt1 knockdown mice. Neurosci Res. 2006;56(4):470–474. doi: 10.1016/j.neures.2006.09.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Yokoi F, Dang MT, Mitsui S, Li J, Li Y. Motor deficits and hyperactivity in cerebral cortex-specific Dyt1 conditional knockout mice. J Biochem. 2008;143(1):39–47. doi: 10.1093/jb/mvm191. [DOI] [PubMed] [Google Scholar]
  • 53.Yokoi F, Dang MT, Li Y. Improved motor performance in Dyt1 ΔGAG heterozygous knock-in mice by cerebellar Purkinje-cell specific Dyt1 conditional knocking-out. Behav Brain Res. 2012;230(2):389–398. doi: 10.1016/j.bbr.2012.02.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Jin XL, Guo H, Mao C, et al. Emx1-specific expression of foreign genes using “knock-in” approach. Biochem Biophys Res Commun. 2000;270(3):978–982. doi: 10.1006/bbrc.2000.2532. [DOI] [PubMed] [Google Scholar]
  • 55.Dang MT, Yokoi F, Yin HH, Lovinger DM, Wang Y, Li Y. Disrupted motor learning and long-term synaptic plasticity in mice lacking NMDAR1 in the striatum. Proc Natl Acad Sci U S A. 2006;103(41):15254–15259. doi: 10.1073/pnas.0601758103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Yokoi F, Dang MT, Li J, Standaert DG, Li Y. Motor deficits and decreased striatal dopamine receptor 2 binding activity in the striatum-specific Dyt1 conditional knockout mice. PLoS One. 2011;6(9):e24539. doi: 10.1371/journal.pone.0024539. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Rossi J, Balthasar N, Olson D, et al. Melanocortin-4 receptors expressed by cholinergic neurons regulate energy balance and glucose homeostasis. Cell Metab. 2011;13(2):195–204. doi: 10.1016/j.cmet.2011.01.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Sciamanna G, Hollis R, Ball C, et al. Cholinergic dysregulation produced by selective inactivation of the dystonia-associated protein torsinA. Neurobiol Dis. 2012;47(3):416–427. doi: 10.1016/j.nbd.2012.04.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.LeDoux MS, Lorden JF, Ervin JM. Cerebellectomy eliminates the motor syndrome of the genetically dystonic rat. Exp Neurol. 1993;120(2):302–310. doi: 10.1006/exnr.1993.1064. [DOI] [PubMed] [Google Scholar]
  • 60.Campbell DB, North JB, Hess EJ. Tottering mouse motor dysfunction is abolished on the Purkinje cell degeneration (pcd) mutant background. Exp Neurol. 1999;160(1):268–278. doi: 10.1006/exnr.1999.7171. [DOI] [PubMed] [Google Scholar]
  • 61.Barski JJ, Dethleffsen K, Meyer M. Cre recombinase expression in cerebellar Purkinje cells. Genesis. 2000;28(3-4):93–98. [PubMed] [Google Scholar]
  • 62.Zhang L, Yokoi F, Jin YH, et al. Altered Dendritic Morphology of Purkinje cells in Dyt1 DeltaGAG Knock-In and Purkinje Cell-Specific Dyt1 Conditional Knockout Mice. PLoS One. 2011;6(3):e18357. doi: 10.1371/journal.pone.0018357. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Hoshi E, Tremblay L, Féger J, Carras PL, Strick PL. The cerebellum communicates with the basal ganglia. Nat Neurosci. 2005;8(11):1491–1493. doi: 10.1038/nn1544. [DOI] [PubMed] [Google Scholar]
  • 64.Shashidharan P, Sandu D, Potla U, et al. Transgenic mouse model of early-onset DYT1 dystonia. Hum Mol Genet. 2005;14(1):125–133. doi: 10.1093/hmg/ddi012. [DOI] [PubMed] [Google Scholar]
  • 65.Lange N, Hamann M, Shashidharan P, Richter A. Behavioural and pharmacological examinations in a transgenic mouse model of early-onset torsion dystonia. Pharmacol Biochem Behav. 2011;97(4):647–655. doi: 10.1016/j.pbb.2010.11.005. [DOI] [PubMed] [Google Scholar]
  • 66.Sharma N, Baxter MG, Petravicz J, et al. Impaired motor learning in mice expressing torsinA with the DYT1 dystonia mutation. J Neurosci. 2005;25(22):5351–5355. doi: 10.1523/JNEUROSCI.0855-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Zhao Y, DeCuypere M, LeDoux MS. Abnormal motor function and dopamine neurotransmission in DYT1 DeltaGAG transgenic mice. Exp Neurol. 2008;210(2):719–730. doi: 10.1016/j.expneurol.2007.12.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Balcioglu A, Kim MO, Sharma N, Cha JH, Breakefield XO, Standaert DG. Dopamine release is impaired in a mouse model of DYT1 dystonia. J Neurochem. 2007;102(3):783–788. doi: 10.1111/j.1471-4159.2007.04590.x. [DOI] [PubMed] [Google Scholar]
  • 69.Hewett J, Johanson P, Sharma N, Standaert D, Balcioglu A. Function of dopamine transporter is compromised in DYT1 transgenic animal model in vivo. J Neurochem. 2010;113(1):228–235. doi: 10.1111/j.1471-4159.2010.06590.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Martella G, Tassone A, Sciamanna G, et al. Impairment of bidirectional synaptic plasticity in the striatum of a mouse model of DYT1 dystonia: role of endogenous acetylcholine. Brain. 2009;132(Pt 9):2336–2349. doi: 10.1093/brain/awp194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Sciamanna G, Tassone A, Mandolesi G, et al. Cholinergic Dysfunction Alters Synaptic Integration between Thalamostriatal and Corticostriatal Inputs in DYT1 Dystonia. J Neurosci. 2012;32(35):11991–12004. doi: 10.1523/JNEUROSCI.0041-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Pisani A, Martella G, Tscherter A, et al. Altered responses to dopaminergic D2 receptor activation and N-type calcium currents in striatal cholinergic interneurons in a mouse model of DYT1 dystonia. Neurobiol Dis. 2006;24(2):318–325. doi: 10.1016/j.nbd.2006.07.006. [DOI] [PubMed] [Google Scholar]
  • 73.Sciamanna G, Tassone A, Martella G, et al. Developmental profile of the aberrant dopamine D2 receptor response in striatal cholinergic interneurons in DYT1 dystonia. PLoS One. 2011;6(9):e24261. doi: 10.1371/journal.pone.0024261. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Napolitano F, Pasqualetti M, Usiello A, et al. Dopamine D2 receptor dysfunction is rescued by adenosine A2A receptor antagonism in a model of DYT1 dystonia. Neurobiol Dis. 2010;38(3):434–445. doi: 10.1016/j.nbd.2010.03.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Sciamanna G, Bonsi P, Tassone A, et al. Impaired striatal D2 receptor function leads to enhanced GABA transmission in a mouse model of DYT1 dystonia. Neurobiol Dis. 2009;34(1):133–145. doi: 10.1016/j.nbd.2009.01.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Page ME, Bao L, Andre P, et al. Cell-autonomous alteration of dopaminergic transmission by wild type and mutant (DeltaE) TorsinA in transgenic mice. Neurobiol Dis. 2010;39(3):318–326. doi: 10.1016/j.nbd.2010.04.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Grundmann K, Reischmann B, Vanhoutte G, et al. Overexpression of human wildtype torsinA and human DeltaGAG torsinA in a transgenic mouse model causes phenotypic abnormalities. Neurobiol Dis. 2007;27(2):190–206. doi: 10.1016/j.nbd.2007.04.015. [DOI] [PubMed] [Google Scholar]
  • 78.Grundmann K, Glockle N, Martella G, et al. Generation of a novel rodent model for DYT1 dystonia. Neurobiol Dis. 2012;47(1):61–74. doi: 10.1016/j.nbd.2012.03.024. [DOI] [PubMed] [Google Scholar]
  • 79.Zimprich A, Grabowski M, Asmus F, et al. Mutations in the gene encoding epsilon-sarcoglycan cause myoclonus-dystonia syndrome. Nat Genet. 2001;29(1):66–69. doi: 10.1038/ng709. [DOI] [PubMed] [Google Scholar]
  • 80.Asmus F, Zimprich A, Tezenas Du Montcel S, et al. Myoclonus-dystonia syndrome: epsilon-sarcoglycan mutations and phenotype. Ann Neurol. 2002;52(4):489–492. doi: 10.1002/ana.10325. [DOI] [PubMed] [Google Scholar]
  • 81.Foncke EM, Gerrits MC, van Ruissen F, et al. Distal myoclonus and late onset in a large Dutch family with myoclonus-dystonia. Neurology. 2006;67(9):1677–1680. doi: 10.1212/01.wnl.0000242880.49051.1f. [DOI] [PubMed] [Google Scholar]
  • 82.Piras G, El Kharroubi A, Kozlov S, et al. Zac1 (Lot1), a potential tumor suppressor gene, and the gene for epsilon-sarcoglycan are maternally imprinted genes: identification by a subtractive screen of novel uniparental fibroblast lines. Mol Cell Biol. 2000;20(9):3308–3315. doi: 10.1128/mcb.20.9.3308-3315.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Yokoi F, Dang MT, Mitsui S, Li Y. Exclusive paternal expression and novel alternatively spliced variants of epsilon-sarcoglycan mRNA in mouse brain. FEBS Lett. 2005;579(21):4822–4828. doi: 10.1016/j.febslet.2005.07.065. [DOI] [PubMed] [Google Scholar]
  • 84.Ritz K, van Schaik BD, Jakobs ME, et al. SGCE isoform characterization and expression in human brain: implications for myoclonus-dystonia pathogenesis? Eur J Hum Genet. 2011;19(4):438–444. doi: 10.1038/ejhg.2010.206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Ettinger AJ, Feng G, Sanes JR. epsilon-Sarcoglycan, a broadly expressed homologue of the gene mutated in limb-girdle muscular dystrophy 2D. J Biol Chem. 1997;272(51):32534–32538. doi: 10.1074/jbc.272.51.32534. [DOI] [PubMed] [Google Scholar]
  • 86.Ettinger AJ, Feng G, Sanes JR. Additions and Corrections to epsilon-Sarcoglycan, a broadly expressed homologue of the gene mutated in limb-girdle muscular dystrophy 2D. J Biol Chem. 1998;273(31):19922. doi: 10.1074/jbc.272.51.32534. [DOI] [PubMed] [Google Scholar]
  • 87.Yokoi F, Dang MT, Li J, Li Y. Myoclonus, motor deficits, alterations in emotional responses and monoamine metabolism in epsilon-sarcoglycan deficient mice. J Biochem. 2006;140(1):141–146. doi: 10.1093/jb/mvj138. [DOI] [PubMed] [Google Scholar]
  • 88.Lancioni A, Luisa Rotundo I, Monique Kobayashi Y, et al. Combined deficiency of alpha and epsilon sarcoglycan disrupts the cardiac dystrophin complex. Hum Mol Genet. 2011 doi: 10.1093/hmg/ddr398. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Schwenk F, Baron U, Rajewsky K. A cre-transgenic mouse strain for the ubiquitous deletion of loxP-flanked gene segments including deletion in germ cells. Nucleic Acids Res. 1995;23(24):5080–5081. doi: 10.1093/nar/23.24.5080. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Hubsch C, Vidailhet M, Rivaud-Péchoux S, et al. Impaired saccadic adaptation in DYT11 dystonia. J Neurol Neurosurg Psychiatry. 2011;82(10):1103–1106. doi: 10.1136/jnnp.2010.232793. [DOI] [PubMed] [Google Scholar]
  • 91.Prsa M, Thier P. The role of the cerebellum in saccadic adaptation as a window into neural mechanisms of motor learning. Eur J Neurosci. 2011;33(11):2114–2128. doi: 10.1111/j.1460-9568.2011.07693.x. [DOI] [PubMed] [Google Scholar]
  • 92.Nagano A, Arahata K. Nuclear envelope proteins and associated diseases. Curr Opin Neurol. 2000;13(5):533–539. doi: 10.1097/00019052-200010000-00005. [DOI] [PubMed] [Google Scholar]
  • 93.Yokoi F, Dang MT, Zhou T, Li Y. Abnormal nuclear envelopes in the striatum and motor deficits in DYT11 myoclonus-dystonia mouse models. Hum Mol Genet. 2012;21(4):916–925. doi: 10.1093/hmg/ddr528. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Yokoi F, Dang MT, Yang G, et al. Abnormal nuclear envelope in the cerebellar Purkinje cells and impaired motor learning in DYT11 myoclonus-dystonia mouse models. Behav Brain Res. 2012;227(1):12–20. doi: 10.1016/j.bbr.2011.10.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Beukers RJ, Booij J, Weisscher N, Zijlstra F, van Amelsvoort TA, Tijssen MA. Reduced striatal D2 receptor binding in myoclonus-dystonia. Eur J Nucl Med Mol Imaging. 2009;36(2):269–274. doi: 10.1007/s00259-008-0924-9. [DOI] [PubMed] [Google Scholar]
  • 96.Zhang L, Yokoi F, Parsons DS, Standaert DG, Li Y. Alteration of striatal dopaminergic neurotransmission in a mouse model of DYT11 myoclonus-dystonia. PLoS One. 2012;7(3):e33669. doi: 10.1371/journal.pone.0033669. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Brashear A, Dobyns WB, de Carvalho Aguiar P, et al. The phenotypic spectrum of rapid-onset dystonia-parkinsonism (RDP) and mutations in the ATP1A3 gene. Brain. 2007;130(Pt 3):828–835. doi: 10.1093/brain/awl340. [DOI] [PubMed] [Google Scholar]
  • 98.Dobyns WB, Ozelius LJ, Kramer PL, et al. Rapid-onset dystonia-parkinsonism. Neurology. 1993;43(12):2596–2602. doi: 10.1212/wnl.43.12.2596. [DOI] [PubMed] [Google Scholar]
  • 99.Linazasoro G, Indakoetxea B, Ruiz J, Van Blercom N, Lasa A. Possible sporadic rapid-onset dystonia-parkinsonism. Mov Disord. 2002;17(3):608–609. doi: 10.1002/mds.10103. [DOI] [PubMed] [Google Scholar]
  • 100.Pittock SJ, Joyce C, O’Keane V, et al. Rapid-onset dystonia-parkinsonism: a clinical and genetic analysis of a new kindred. Neurology. 2000;55(7):991–995. doi: 10.1212/wnl.55.7.991. [DOI] [PubMed] [Google Scholar]
  • 101.Zaremba J, Mierzewska H, Lysiak Z, Kramer P, Ozelius LJ, Brashear A. Rapid-onset dystoniaparkinsonism: a fourth family consistent with linkage to chromosome 19q13. Mov Disord. 2004;19(12):1506–1510. doi: 10.1002/mds.20258. [DOI] [PubMed] [Google Scholar]
  • 102.Brashear A, DeLeon D, Bressman SB, Thyagarajan D, Farlow MR, Dobyns WB. Rapid-onset dystonia-parkinsonism in a second family. Neurology. 1997;48(4):1066–1069. doi: 10.1212/wnl.48.4.1066. [DOI] [PubMed] [Google Scholar]
  • 103.de Carvalho Aguiar P, Sweadner KJ, Penniston JT, et al. Mutations in the Na+/K+ -ATPase alpha3 gene ATP1A3 are associated with rapid-onset dystonia parkinsonism. Neuron. 2004;43(2):169–175. doi: 10.1016/j.neuron.2004.06.028. [DOI] [PubMed] [Google Scholar]
  • 104.Cannon SC. Pathomechanisms in channelopathies of skeletal muscle and brain. Annu Rev Neurosci. 2006;29:387–415. doi: 10.1146/annurev.neuro.29.051605.112815. [DOI] [PubMed] [Google Scholar]
  • 105.Moseley AE, Williams MT, Schaefer TL, et al. Deficiency in Na,K-ATPase alpha isoform genes alters spatial learning, motor activity, and anxiety in mice. J Neurosci. 2007;27(3):616–626. doi: 10.1523/JNEUROSCI.4464-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.DeAndrade MP, Yokoi F, van Groen T, Lingrel JB, Li Y. Characterization of Atp1a3 mutant mice as a model of rapid-onset dystonia with parkinsonism. Behav Brain Res. 2011;216(2):659–665. doi: 10.1016/j.bbr.2010.09.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Kirshenbaum GS, Saltzman K, Rose B, Petersen J, Vilsen B, Roder JC. Decreased neuronal Na+, K+ -ATPase activity in Atp1a3 heterozygous mice increases susceptibility to depression-like endophenotypes by chronic variable stress. Genes Brain Behav. 2011;10(5):542–550. doi: 10.1111/j.1601-183X.2011.00691.x. [DOI] [PubMed] [Google Scholar]
  • 108.Quartarone A, Rizzo V, Morgante F. Clinical features of dystonia: a pathophysiological revisitation. Curr Opin Neurol. 2008;21(4):484–490. doi: 10.1097/WCO.0b013e328307bf07. [DOI] [PubMed] [Google Scholar]
  • 109.Yu H, Neimat JS. The treatment of movement disorders by deep brain stimulation. Neurotherapeutics. 2008;5(1):26–36. doi: 10.1016/j.nurt.2007.10.072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Cao S, Hewett JW, Yokoi F, et al. Chemical enhancement of torsinA function in cell and animal models of torsion dystonia. Dis Model Mech. 2010;3(5-6):386–396. doi: 10.1242/dmm.003715. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Carroll D. Genome engineering with zinc-finger nucleases. Genetics. 2011;188(4):773–782. doi: 10.1534/genetics.111.131433. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Yan Z, Sun X, Engelhardt JF. Progress and prospects: techniques for site-directed mutagenesis in animal models. Gene Ther. 2009;16(5):581–588. doi: 10.1038/gt.2009.16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Maeder ML, Thibodeau-Beganny S, Osiak A, et al. Rapid “open-source” engineering of customized zinc-finger nucleases for highly efficient gene modification. Mol Cell. 2008;31(2):294–301. doi: 10.1016/j.molcel.2008.06.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Reyon D, Kirkpatrick JR, Sander JD, et al. ZFNGenome: a comprehensive resource for locating zinc finger nuclease target sites in model organisms. BMC Genomics. 2011;12:83. doi: 10.1186/1471-2164-12-83. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Tanabe LM, Martin C, Dauer WT. Genetic background modulates the phenotype of a mouse model of DYT1 dystonia. PLoS One. 2012;7(2):e32245. doi: 10.1371/journal.pone.0032245. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES