CHAPTER EIGHT

Animal Models Recapitulating the Multifactorial Origin of Tourette Syndrome Simone Macrì*,1, Martina Proietti Onori*,1, Veit Roessner†, Giovanni Laviola*,2

*Section of Behavioural Neuroscience, Department Cell Biology and Neuroscience, Istituto Superiore di Sanita`, Roma, Italy † Department of Child and Adolescent Psychiatry, Technical University Dresden, Dresden, Germany 1 S. M. and M. P.O. equally contributed to the chapter. 2 Corresponding author: e-mail address: [email protected]

Contents 1. Tic Disorders and Tourette Syndrome 2. Tourette Syndrome: Etiological Factors 2.1 Genetic factors 2.2 Environmental factors 3. Animal Models of TS 3.1 Transgenic animal models 3.2 Immune-mediated animal models 3.3 Stress paradigms mimicking psychosocial stress in mice 4. Future Perspectives Acknowledgments References

212 213 214 215 217 218 221 224 227 228 228

Abstract Tourette Syndrome (TS) is a neurological disorder characterized by motor and phonic tics affecting approximately 1% of the pediatric population. Behavioral comorbidities often include obsessive–compulsive behavior and impaired attention. The neurobiological substrates associated with TS generally entail abnormalities in neurotransmitter circuitry regulating basal ganglia activity. The neurotransmitters most often associated with TS are dopamine, serotonin, and GABA. TS origin roots in genetic predisposing factors, and environmental variables favoring tic onset and exacerbation. Among the latter, repeated infections with group A beta-hemolytic Streptococcus and psychosocial stressors encountered during development have been proposed to constitute likely susceptibility factors. In this chapter, we describe how this clinical/epidemiological knowledge has been translated into animal models of TS. Specifically, we review several studies attempting to reproduce TS-like symptoms (tics and behavioral stereotypies) and comorbidities (impaired attention, increased locomotion, and perseverative

International Review of Neurobiology, Volume 112 ISSN 0074-7742 http://dx.doi.org/10.1016/B978-0-12-411546-0.00008-1

#

2013 Elsevier Inc. All rights reserved.

211

212

Simone Macrì et al.

responding) in laboratory rodents. Additionally, we discuss studies in which the genetic and environmental predisposing factors have been modeled in experimental subjects. Ultimately, we propose a unifying perspective recapitulating dependent and independent variables in the preclinical study of TS and discuss its potential theoretical and heuristic implications.

1. TIC DISORDERS AND TOURETTE SYNDROME Tics are defined as nonrhythmic, involuntary, and rapid movements or sounds (American Psychiatric Association, 2000). Most children, approximately 15%, experience during their development a transient tic condition, not requiring medical treatment, with one or few tics that disappear within less than 1 year (Robertson, 2008). The severity and course of the expression of tics differentiate the transient tic condition from the chronic tic disorder in which the motor or vocal tics are present for more than 1 year (Kurlan et al., 2001). Tourette Syndrome (TS) is diagnosed when both chronic motor and phonic tics are expressed for more than 1 year (Swain, Scahill, Lombroso, King, & Leckman, 2007). TS is a neuropsychiatric disorder with a childhood onset that is estimated to mainly affect prepubertal boys (Leckman, Bloch, Smith, Larabi, & Hampson, 2010). Its prevalence in the pediatric population ranges between 0.4% and 1%. TS symptoms fluctuate over time in frequency and intensity, reach a peak in severity between 10 and 14 years of age and tend to dissipate during late adolescence or early adulthood (Leckman, 2012). Additionally, affected children may experience comorbid symptoms that characterize other neurobehavioral manifestations, including attentiondeficit hyperactivity disorder (ADHD) and obsessive–compulsive disorder (OCD), such as aggressiveness, emotional liability, affective disorders, and compulsive behaviors (Cohen, Leckman, & Bloch, 2013; Lombroso & Scahill, 2008). Different environmental variables may influence the waxing and waning course of tics. Specifically, psychosocial stressors, anxiety, and emotional tension may exacerbate the expression of tics (Hoekstra, Dietrich, Edwards, Elamin, & Martino, 2013). In most cases, affected individuals are able to exert a voluntary cognitive control over the expression of tics, but after a period of suppression tics may even become more uncontrollable and disabling (Jung, Jackson, Parkinson, & Jackson, 2012). Furthermore, shortly before the expression of tics, patients describe a restless feeling of discomfort, a “premonitory urge” experienced in the body’s area where the tic is about to occur (Bliss, 1980; Jankovic & Kurlan, 2011; Leckman et al., 2010).

Animal Models of Tourette Syndrome

213

2. TOURETTE SYNDROME: ETIOLOGICAL FACTORS Several studies (neuroimaging, neurophysiological, and postmortem analyses on brain tissues) identified functional and anatomical abnormalities in brains of TS patients at the level of the circuitry connecting the basal ganglia to the cerebral cortex (Berardelli, Curra`, Fabbrini, Gilio, & Manfredi, 2003; Singer & Minzer, 2003). The basal ganglia, which include the striatum, the subthalamic nucleus, and the globus pallidus, play a significant role in the regulation of the cortex excitability and in the selection of movements through the involvement of different neurotransmitter systems. In particular, the inhibitory neurotransmitter gamma-aminobutyric acid (GABA) has a prominent role in the functioning of the striatum and globus pallidus, in conjunction with a modulatory role played by dopamine projections from the substantia nigra, located in the midbrain, to the striatum (Ganos, Roessner, & Mu¨nchau, 2013). The efficacy of specific medical treatments for the reduction of tics and the use of pharmacological animal models (see Section 3) focused the attention on the involvement of specific neuromodulatory systems on the emergence of tics (Bronfeld, Israelashvili, & Bar-Gad, 2012). Evidence for a causal relationship between dopamine dysfunction and TS pathophysiology came from the finding that the use of antipsychotic drugs (pimozide and haloperidol) acting as antagonists on D2 receptors reduces the expression of tics. Furthermore, an excess of nigrostriatal activity has been found in TS patients (Albin et al., 2003) with increased levels of D2 receptors and DA transporter (DAT) levels (Cheon et al., 2004; Wolf et al., 1996). The dopaminergic pathway and other neurochemical systems are modulated by the noradrenergic system located in the brain stem, mainly in the locus coeruleus, and innervating all central nervous system regions (Foote, Bloom, & Aston-Jones, 1983). Alpha 2-adrenergic agonists have been found to exert positive effects on tic suppression, suggesting that a hyperadrenergic system may be associated with TS (Arnsten, 2001; Arnsten & Pliszka, 2011). However, a systematic metaanalysis revealed that the efficacy of Alpha 2-adrenergic agonists in mitigating TS symptoms may depend on the presence of comorbid symptoms. Thus, Alpha 2 adrenergic agonists were particularly beneficial only in the presence of comorbid ADHD symptoms (Weisman, Qureshi, Leckman, Scahill, & Bloch, 2013). Most treatments affecting serotonin functions have not shown significant results (Scahill et al., 1997) even though patients have been found to have lower plasma levels of tryptophan (the precursor of serotonin in the

214

Simone Macrì et al.

biosynthetic pathway) than normal and some postmortem studies have shown reduced brain tryptophan concentrations (Comings, 1990).

2.1. Genetic factors Several studies supported the hypothesis that TS is an inherited developmental disorder of neurotransmission. Prevalence of TS in first-degree relatives ranges between 5% and 15% (O’Rourke, Scharf, Yu, & Pauls, 2009). Furthermore, genetic studies show that the ratio of concordance in monozygotic versus dizygotic twin pairs is approximately 5:1 (Price, Kidd, Cohen, Pauls, & Leckman, 1985). Previous data supported the hypothesis of a single major autosomal dominant gene with pleiotropic expression (i.e., chronic motor tics, TS, or OCD) and incomplete penetrance (about 70% in women, 99% in men; Pauls, 1992). Recent linkage methods and analyses of families with visible chromosomal abnormalities are currently used to decipher the genetic contribution to TS (Deng, Gao, & Jankovic, 2012). Several candidate susceptibility genes are emerging from clinical studies of chromosomal aberrations. For example, the disruption of the IMMP2L (inner mitochondrial membrane peptidase 2 like) gene, due to a translocation breakpoint on the chromosome 7, has been identified in one isolated and one familial case of TS, providing evidence for a role of this gene in the pathogenesis of TS (Patel et al., 2011; Petek et al., 2001). Particular attention has been dedicated to members of cell-adhesion molecules at nerve cell synapses. In this framework, a mutation in the X-linked gene neuroligin4 (NLGN4X) has been discovered in a family with various developmental disorders (autism, ADHD, and TS) (Lawson-Yuen, Saldivar, Sommer, & Picker, 2008) and a truncation in the region of the CNTNAP2 gene has been found in a family with TS (Verkerk et al., 2003). The CNTNAP2 is a gene encoding Caspr2 (contactin-associated protein-like 2), a member of the neurexins superfamily localized to the juxtaparanodal regions of myelinated axons and involved in the correct positioning of the Shaker-type voltage-activated Kþ channels (Poliak et al., 1999). The expression of tics in TS patients may be related to the influence of this mutation in the conduction and repolarization of action potentials in specific brain regions, specifically in striatal circuits and in the frontal cortex of the adult human brain, where the messenger of CNTNAP2 is particularly enriched (Abrahams et al., 2007). Nevertheless, disruption of CNTNAP2 does not necessarily lead to the symptoms of TS (Belloso et al., 2007), giving more validity to the likely polygenic character of the syndrome.

Animal Models of Tourette Syndrome

215

Two different mutations have been found in a member of the SLIT and TRK family of proteins, SLITRK1, in TS patients: a deletion that led to a truncated form of the protein and a missense mutation in the 30 -untranslated region of the gene (Abelson et al., 2005; Proenca, Gao, Shmelkov, Rafii, & Lee, 2011). This protein is involved in neurite outgrowth and branching (Kajiwara, Buxbaum, & Grice, 2009; Linhoff et al., 2009) and is expressed in the cortex, thalamus, and basal ganglia, reflecting the circuit most commonly implicated in TS (Stillman et al., 2009). Interestingly, the finding of a single rare coding mutation in the gene encoding the rate-limiting enzyme in the biosynthetic pathway of histamine, the L-histidine decarboxylase (HDC) gene, led to the hypothesis of a possible involvement of histaminergic pathway in the pathogenesis of TS (Ercan-Sencicek et al., 2010). The G protein-coupled receptor for histamine, H3R, is highly enriched in the striatum of rodents and humans and regulates both dopamine and serotonin actions (Haas, Sergeeva, & Selbach, 2008). A recent study by Karagiannidis et al. (2013) partly supported the histaminergic hypothesis through the analysis of variations across the HDC gene in a large sample of familiar cases of TS. In this study, the authors observed an over-transmission of two single nucleotide polymorphisms in the HDC gene associated with TS (Karagiannidis et al., 2013).

2.2. Environmental factors Recent observations suggest that TS may have a multifactorial nature in which environmental factors in interaction with the polygenic background (Seuchter et al., 2000) may contribute to the onset of the pathology (State, 2011). Prenatal environmental risk factors (such as perinatal hypoxic/ischemic events, prenatal maternal smoking, low birth weight, and maternal stress) may exert an organizational role in the development and functioning of brain pathways thought to be relevant for the emergence of tics (Hoekstra et al., 2013). Studies involving affected monozygotic twins revealed that the twin that suffered perinatal complications had a higher intensity in tic severity (Randolph, Hyde, Gold, Goldberg, & Weinberger, 1993). Beside prenatal factors, postnatal environmental factors may affect the activity of neurons in crucial brain regions, possibly relating to the fluctuation in tic severity. More recently, streptococcal infections have been proposed as an additional environmental factor potentially favoring the production of tics through immune-mediated mechanisms (Cardona & Orefici, 2001). Martino et al. (2011) systematically addressed this hypothesis in a cohort study. Although the authors observed that infections did not

216

Simone Macrì et al.

predict exacerbations in tic severity, they observed an increased susceptibility to streptococcal infections in TS patients. In several cases, tics began suddenly after a streptococcal infection, thus leading to the proposal of a working hypothesis explaining pediatric autoimmune neuropsychiatric disorders associated with streptococcal infection (PANDAS; Swedo et al., 1998). The proposed definition criteria for PANDAS are: (1) prepubertal onset, (2) the presence of chronic tic disorders or OCD, (3) a relapsing–remitting course, (4) clinical evidence of Group A Streptococcus (GAS) infection associated with onset or exacerbation of tics, (5) association with neurological abnormalities such as a reduced motor coordination or motor hyperactivity (but not chorea). It is important to notice that defining criteria for PANDAS are still highly debated and under constant reappraisal. Despite the potential role attributed to streptococcal infections in modulating TS symptoms, the precise relationship between such infections, antineuronal antibodies, and TS remains elusive. Such working hypothesis is currently under investigation in several initiatives including a pan-European research project (http://www.emtics.eu). GASinduced antibodies directed against different epitopes in the basal ganglia have been detected in some cases of TS (Kiessling et al., 1993, 1994; Martino & Giovannoni, 2004; Singer et al., 1998), thereby strengthening the hypothesis of an immune or autoantibody-mediated mechanism. In other cases, the link between GAS infections and TS has not been detected (Kawikova et al., 2010; Martino et al., 2011; Singer, Gause, Morris, & Lopez, 2008) and the levels of autoantibodies in patients’ sera have been found equivalent to the levels of controls (Kawikova et al., 2010; Morris, Pardo-Villamizar, Gause, & Singer, 2009; Singer, Hong, Yoon, & Williams, 2005). Mimicry models have been proposed to explain the autoantibody-mediated mechanism, one involving a cross-reactivity between a carbohydrate domain of the bacterial cell wall and brain lysoganglioside GM1 and another involving cross-reactivity between streptococcal and neuronal isoforms of glycolytic enzymes (Dale et al., 2006). Some observations support a connection between GAS infection and tics. Children with TS have altered immunoregulatory mechanisms (such as reduced numbers of regulatory T cells and abnormal cytokine secretion) which may predispose to autoimmune responses to GAS and/or may predispose to a higher number of GAS infections (Hornig & Lipkin, 2013). In a large case–control study, children with OCD or a chronic tic disorder were twice as likely as controls to have had a documented GAS infection in the 3 months prior to the neuropsychiatric diagnosis, and children with multiple GAS infections in

Animal Models of Tourette Syndrome

217

the preceding year were more likely to be diagnosed with TS than children who did not encounter multiple infections (Mell, Davis, & Owens, 2005). However, independent investigators failed to replicate these findings, whereby a clear association between streptococcal infections and TS symptoms was not observed (Schrag et al., 2009). Yet, the limited statistical power of this study does not allow unequivocal conclusions. Furthermore, several patients with either tics or OCD have been found to have high levels of antistreptococcal circulating antibodies directed against neuronal components (Kiessling et al., 1993; Wendlandt, Grus, Hansen, & Singer, 2001). The complexity of TS etiology is increased by the clinical observation that TS patients seem to be more sensitive to psychosocial stress than normal subjects (Chappell et al., 1994; Conelea & Woods, 2008; Corbett, Mendoza, Baym, Bunge, & Levine, 2008). Children with TS show a higher reactivity of the hypothalamic–pituitary–adrenal (HPA) axis resulting in a significant elevation of corticotropin-releasing factor and cortisol in response to stressful events (Conelea, Woods, & Brandt, 2011; Corbett et al., 2008). GAS infections and psychosocial stress have been proposed to interact with immune and endocrine systems in creating a neurobiological vulnerability to tics (Lin et al., 2010).

3. ANIMAL MODELS OF TS In line with the aforementioned genetic and environmental hypotheses, several scholars attempted to develop suitable animal models recapitulating TS phenotypic abnormalities and etiological factors. While hardly ever do laboratory rodents exhibit tics, several abnormal spontaneous or pharmacologically induced movements expressed by animals have been considered indicators of TS-like symptoms. The theoretical rationale for the analysis of these manifestations can be traced back to the concepts of face and construct validities. While the former relates to the phenomenological similarity between human symptoms and animal model, the latter is an indicator of the correspondence between the etiological factors precipitating the disease in humans and in the animal model. Ultimately, the behavioral analysis of rodent models of TS attempted to address symptoms that (i) are highly similar to tics, (ii) depend on a dysfunction of the neurotransmitter systems potentially involved in the etiology of TS, (iii) closely resemble symptoms that characterize frequent comorbidities. Within this framework, several authors identified specific behavioral manifestations analogous to tics. These behavioral patterns generally manifest as jerk-like movements or brief and

218

Simone Macrì et al.

0.0 s

0.1 s

0.2 s

0.3 s

0.4 s

0.5 s

Figure 8.1 Still-frame sequence of a body-jerk response induced by the administration of the selective 5HT2A receptor agonist DOI (2,5-dimethoxy-4-iodoamphetamine).

repetitive twitching performed at low frequency during grooming behavior (Tse & Wei, 1986). A rapid increase in head twitch and body-jerk (Fig. 8.1) responses, highly isomorphic to tics, has been described in response to the administration of drugs that increase serotonergic activity (Canal & Morgan, 2012). An analogous pharmacological modulation has been leveraged through the use of substances affecting neurotransmitter systems within the cortico-striatal circuitry, as GABAa (McKenzie, Gordon, & Viik, 1972; Patel & Slater, 1987) and dopamine (Dantzer, 1986). These substances have been shown to induce abnormal behaviors similar to those induced by serotonergic drugs (Dursun & Handley, 1992). Behavioral stereotypies, apparently purposeless repetitive movements, are also generally ascribed to a dysfunction of the basal ganglia. Therefore, they often constitute a relevant parameter in the analysis of putative mouse models of TS. Ultimately, several studies attempted to investigate behavioral and cognitive domains frequently affected in comorbid disturbances as ADHD and OCD. Within this framework, attentional set-shifting abilities and hyperlocomotion may constitute relevant outcome measures. The validity of these behavioral patterns as indicators of TS-like abnormalities has already been detailed elsewhere (Macrı`, Proietti Onori, & Laviola, 2013).

3.1. Transgenic animal models Beside the pharmacological induction of behaviors resembling the symptomatology, several attempts have been made to reproduce the genetic

Animal Models of Tourette Syndrome

219

predispositions that may favor the onset and progression of the pathology. Mutations in genes known to be involved in the disorder, null mutations as well as mutations which alter but do not eliminate gene function, can be introduced in the mouse genome by means of gene targeting procedures. The aim is to replicate, as far as possible, the symptoms according to the genetic etiological factors proposed for the disease (construct validity). Susceptibility genes related to neurotransmitter systems involved in the cortico-striatal circuit have been studied through different engineered rodent models (Paschou, 2013). Specifically, various dopamine receptors (Comings et al., 1991; Lee et al., 2005) and the dopamine transporter gene (DAT1) (Dı´az-Anzaldu´a et al., 2004) have been investigated for association with TS. According to the “dopamine hypothesis” for TS (Buse, Schoenefeld, Mu¨nchau, & Roessner, 2013), based on the idea of an excess nigrostriatal dopaminergic activity (Singer, Butler, Tune, Seifert, & Coyle, 1982), a mutant mouse model of hyperdopaminergic activation (DAT knockdown) has been developed (Zhuang et al., 2001). These mice exhibit elevated levels of extracellular DA in the neostriatum and tend to be hyperactive and show excessive body grooming (an activity in which the animal, usually in sitting position, licks its fur and grooms with the forepaws). Interestingly, the excessive locomotor activity is alleviated by the administration of psychostimulants (Robertson & Feng, 2011). Though not equivalent to tics, abnormal repetitive grooming constitutes a motor stereotypy associated with dysfunctions at the level of the basal ganglia (Garner & Mason, 2002; Gross, Richter, Engel, & Wu¨rbel, 2012). Ultimately, abnormalities in grooming may provide relevant information regarding the neurobiology of tics. Greer and Capecchi (2002) described an excessive grooming behavior in Hoxb8 homozygous mutant mice resulting in remarkable hair pulling (Greer & Capecchi, 2002). The Hoxb8 gene encodes a nuclear protein which binds DNA sequences, expressed in multiple brain regions including the orbitofrontal cortex and the limbic system (Huber, Ferdin, Holzmann, Stubbusch, & Rohrer, 2012). In the light of the compulsive grooming and hair removal, these mice have been proposed as a potential model for the study of OC-spectrum disorders, which represent a common comorbidity of TS (Chen et al., 2010). Alzghoul et al. (2012) recently developed a genetically engineered mouse model exhibiting a reduction in the levels of functional MAO-A enzyme, involved in the degradation of monoamine neurotransmitters. These engineered mice show high levels of serotonin (5-HT) and noradrenaline

220

Simone Macrì et al.

(NA) in the hippocampus, in motor and prefrontal cortex, and in the striatum (Alzghoul et al., 2012; Bortolato et al., 2011). They exhibit reduced motor coordination and balance, reflecting the morphological abnormalities found in the cerebellum. The disruption of the monoamine systems could alter the circuitry connecting the cerebellum with the fronto-striatal basal ganglia region, resulting in motor deficits. These mice also express perseverative behaviors, potential endophenotypes for the study of some neurodevelopmental disorders, including autism-spectrum disorder and TS (Bortolato, Chen, & Shih, 2008). To gain further insights regarding the potential role of the histaminergic transmission, HDC-null mice have been recently developed and constitute a promising research avenue. Even though only partially do HDC-null models recapitulate TS symptomatology, they show decreased histamine levels and a significant increase of stereotypic behaviors upon administration of DA agonists, compared to wild-type mice (Kubota et al., 2002). A high degree of face and construct validity (Van Der Staay, Arndt, & Nordquist, 2009) has been reached with the transgenic mouse model (D1CT-7) developed by Nordstrom and Burton (2002). It expresses two hyperactive groups of neuronal populations in the adult CNS, a glutamatergic neuronal population and a population of GABAergic interneurons (Campbell, McGrath, & Burton, 1999). The D1CT-7 mouse model exhibits TS-like symptoms, including very brief (0.05–0.1 s) isolated head and body jerk or shake resembling human tics. The use of the Alpha 2-noradrenergic agonist clonidine, a commonly used treatment for the reduction of tics in humans (Nordstrom & Burton, 2002), supports the predictive validity of the model, defined as the degree of correspondence between the clinical efficacy of the treatment in human and animals (Korff & Harvey, 2006). Based on the clinical evidence of sequence variants in Slitrk1 gene associated with TS (Proenca et al., 2011), Katayama et al. (2010) developed a Slitrk1-knockout model and analyzed the behavioral and neurochemical abnormalities associated with the loss of this gene. The behavioral results show that these transgenic mice exhibit elevated anxiety-like and depressive-like behaviors. Specifically, in the elevated plus-maze, an approach-avoidance test in which rodents are faced with the possibility of exploring two protected and unprotected arms, they spend less time in the open arms, compared to control mice (the latter being considered an index of anxiety) (Pellow, Chopin, File, & Briley, 1985); in the forced swimming test (the mouse is placed in a tall cylinder filled with water where

Animal Models of Tourette Syndrome

221

it swims seeking an escape route) and in the tail suspension test (the mouse is suspended by its tail and struggles to reach a solid surface), the percentage of time spent in immobility, considered as an index of a depressive-like state, is increased (Katayama et al., 2010). The authors show that the administration of clonidine, such as for the D1CT-7 mouse model, attenuates the increased anxiety-like behavior. The efficacy of clonidine suggests an involvement of the noradrenergic pathways in TS (Kang, Zhang, Jiao, Guo, & Gao, 2009; Leckman et al., 1991). Even though slitrk1-deficient mice do not exhibit any motor stereotypies or tic-like symptoms, neurochemical analyses unveil an alteration in the noradrenergic system with high levels of NA and its metabolite (MHPG) in the prefrontal cortex, striatum, and nucleus accumbens of transgenic mice. These observations, together with a clear genetic hypothesis derived from clinical evidence, beget this model an elevated degree of validity.

3.2. Immune-mediated animal models The heterogeneity of TS pathology makes it likely that environmental risk factors may contribute to the emergence and/or progression of the disorder acting on genetically vulnerable individuals. Within this context, animal models provide the opportunity to clarify the contribution of streptococcal infections to the progression and severity of the pathology. Different immune-mediated animal models of TS have been produced leveraging a variety of strategies. Some of them involve the exposure to immune mediators, such as proinflammatory cytokines (IL-1b and TNF-a), that alter the function of neural pathways relevant to TS; other approaches are based on the immunization with immunogenic microbial components that are thought to cross-react with endogenous targets. Furthermore, sera derived from actively immunized animals or from affected patients can be transfected into naı¨ve animals with the aim of inducing a disruption in the CNS signaling and studying its behavioral consequences (Hornig & Lipkin, 2013). The first line of studies infused directly the serum of patients with antineuronal antibodies into rodent striatum (Hallett, Harling-Berg, Knopf, Stopa, & Kiessling, 2000; Taylor et al., 2002). These studies demonstrated that the infusion of sera from TS patients induced stereotypic movements and the presence of IgG deposits in rats. Furthermore, they were useful to detect whether distinct isotypes of Ig were capable of inducing striatal dysfunctions in vivo, although they did not allow the identification of the etiologic factors responsible for the generation of autoantibodies. A promising approach is constituted by the use of preclinical animal

222

Simone Macrì et al.

models exposed to passive transfer immunization protocols. These studies entail the active immunization of subjects through streptococcal antigens. Such immunization readily results in the production of antibodies, which are then passively transferred to naı¨ve mice (e.g., Yaddanapudi et al., 2010). This approach is functional to understanding the role of antibodies in mediating the effects of Streptococcus regardless of a general unspecific activation of the immune system. 3.2.1 Experimentally induced increment in brain immune mediators Different studies have shown that the peripheral administration of immune mediators such as cytokines and soluble cytokine receptors can alter the development of the CNS leading to persistent behavioral changes, such as motor stereotypies and abnormal repetitive behaviors (Zalcman, Murray, Dyck, Greenberg, & Nance, 1998; Zalcman, Patel, Mohla, Zhu, & Siegel, 2012). Specifically, different cytokines affect the function of distinct brain monoaminergic systems. For example, the administration of IL-2 to male BALB/c mice influences the release of DA in mesocorticolimbic structures while IL-6 also affects the activity of 5-HT (Zalcman et al., 1994, 1998). A series of behavioral domains appear to be affected. Prenatal treatments that involved the exposure of mice to IL-2 and IL-6 cytokines led to long-lasting immune and behavioral dysfunctions in the offspring of mice (Ponzio, Servatius, Beck, Marzouk, & Kreider, 2007; Smith, Li, Garbett, Mirnics, & Patterson, 2007). These studies demonstrated that the exposure to increased levels of cytokines during plastic stages of brain development might cause immunological alterations and behavioral disturbances (increased grooming and rearing) that represent a core characteristic of TS. It has been proposed that increased cytokine exposure during key periods of brain development as a consequence of early life infections may act as a “sensitizing” factor altering the way the brain responds to later-life immune challenges (Bilbo & Schwarz, 2009). Microglial cells have been proposed as the suitable target for the long-term changes occurring within the brain in consequence of neonatal infections (Bilbo et al., 2005). Microglia represent the immunocompetent cells of the brain that, in response to immune stimulation, become active and produce cytokines and chemokines to recruit immune cells into the brain. Early infections during critical periods of brain development may induce a change in microglia cells (glial priming). Primed microglial cells will overproduce cytokines in response to a future immune challenge compared to normal microglial cells that were not sensitized (Perry, Newman, & Cunningham, 2003).

Animal Models of Tourette Syndrome

223

3.2.2 Animal models based on immunization with microbial immunogens The peripheral exposure of rodents to microbial antigens allows the study of the mechanisms involved in the induction of autoantibodies putatively associated to the etiology of TS. The same experimental approach may provide useful information regarding the process through which immune molecules access the brain by crossing the blood–brain barrier (BBB). Peripherally generated antibodies and immune molecules usually do not cross the BBB (Willis, 2012). A chronic exposure to cytokines, stressful events, and infection can compromise the BBB integrity, thereby facilitating the immune molecules entry into the brain (Mas´li nska, 2001). It has been also hypothesized that they can use an active transport mechanism to cross the BBB. Hoffman, Hornig, Yaddanapudi, Jabado, and Lipkin (2004) observed that female SJL/J mice (a strain with a high propensity to show autoimmune responses), actively immunized with GAS homogenate, exhibited some behavioral abnormalities reminiscent of TS that satisfy some of the criteria for PANDAS (Hoffman et al., 2004). Specifically, these mice show a spectrum of behavioral abnormalities (increased rearing behavior, compulsive grooming, and decreased motor coordination) that correlate with an elevation in titers of antibodies cross-reacting with cell bodies of neurons within the mouse cerebellum, globus pallidus, and thalamus. A subsequent study by the same research group reproduced the key aspects of the previous model (Yaddanapudi et al., 2010). In this case, naı¨ve mice, passively immunized with GAS sera derived from actively immunized mice, exhibited behavioral dysfunctions similar to those exhibited by donor mice (increased rearing and passive social behaviors). Yet, these mice failed to show any alteration in motor coordination. Furthermore, the observation that IgG-depleted GAS sera did not produce the same effect indicates that IgG is the active, key component in the induction of behavioral abnormalities. The principal difference between donor and recipient mice was found in the localization of the IgG deposits in brain. In fact, while IgG deposits in the brains of donor mice were found within the cerebellum, globus pallidus, and thalamus, IgG deposits in recipient mice were confined to neurons in the hippocampus and periventricular area. This could be correlated with the use of different substances used to disrupt the integrity of the BBB (Freund’s adjuvant in the first case and LPS in the passive transfer) (Kerr, Krishnan, Pucak, & Carmen, 2005). Similar findings were obtained with a rat model of PANDAS in which experimental subjects were immunized with a cell-wall antigen preparation

224

Simone Macrì et al.

of GAS M protein type 18 bacteria as immunogen (in contrast with M protein type 6 used by Hoffman and colleagues). Male Lewis rats exposed to GAS exhibited a behavioral profile and an immunological response phenotype that was similar to that shown by mice exposed to active immunization with GAS (Hoffman et al., 2004). Specifically, motor disturbances (impaired manipulation of food and inability to traverse a narrow beam) were observed along with an increase in induced-grooming behavior that might represent a form of compulsive activity, consistent with the compulsive and obsessive symptoms found in the clinic for PANDAS cases (Brimberg et al., 2012). These symptoms were alleviated by the administration of haloperidol (DR2 antagonist) and paroxetine (SSRI), respectively, used to alleviate motor and compulsive symptoms. As previously demonstrated by Hoffman et al. (2004), this study further supports a causal relationship between the exposure to GAS and the development of behavioral disorders underlined by the detection of IgG deposits in the striatum, thalamus, and frontal cortex. In addition, Brimberg et al. (2012) revealed that exposure to GAS resulted in alteration in the glutamatergic and dopaminergic systems, with lower glutamate content and increased levels of DA in the frontal cortex and entopeduncular nucleus. Furthermore, this new model demonstrates that antistreptococcal rat sera (IgG) from GAS-exposed rats react significantly with human dopamine receptor D1 and D2, as detected in sera from children with PANDAS and Sydenham’s chorea (Brimberg et al., 2012). These findings further suggest a role for a dysfunctional cortico-striatal pathway in the induction of behavioral abnormalities, as hypothesized from clinical evidence (Felling & Singer, 2011).

3.3. Stress paradigms mimicking psychosocial stress in mice To investigate the interaction among TS predisposing factors and psychosocial stressors, it is tenable to combine the aforementioned disease models with paradigms imposing variable degrees of stress to laboratory rodents. This aim has been achieved through a plethora of ad hoc developed stimuli and paradigms. A preliminary distinction between these paradigms may rest on the dissociation between chronic and acute stressors. Chronic stressors entail repeated exposures to mild challenges, such as electric shock, water immersion, or restraint. Yet, these models have been criticized based on the following ground: first of all, the perceived relative intensity of a given stressor is both species- and strain-dependent; furthermore, the timing and definition of “chronic” varies across studies, ranging from 2 to 16 h per day

Animal Models of Tourette Syndrome

225

applied over a period of 7–28 days (Schmidt et al., 2007); therefore, some authors questioned the chronic nature of these models as the stressors are only applied for a relatively short time per day or for only a few consecutive days. Finally, laboratory rodents may display habituation to predictable and repeated stressful stimuli, thereby showing reduced psychophysiological responses to the same challenge. To address this potential shortcoming, Willner (2005) devised the chronic mild stress procedure. The latter is characterized by different types of mild stressors presented in a randomized order: this strategy is aimed at minimizing habituation and predictability. A potential limitation for the exploitation of these models to the study of TS is that the latter is favored by social stress. Thus, one major drawback of these rodent stress paradigms is their weak consideration of social stress (Heinrichs & Gaab, 2007). To overcome this limitation, and mimic constant social stress, it is possible to leverage the highly social nature of laboratory rodents altering their housing conditions. Specifically, it is possible to expose laboratory rodents to the visual, olfactory, and nontactile presence of a dominant individual in the home cage (Bartolomucci et al., 2005). The latter has been proposed to constitute one of the most pervasive stressors (Bartolomucci et al., 2005; Fuchs, 2005). There is a large body of evidence indicating that long exposure to social stressors may remarkably influence the activity of the HPA axis and relate to a high “mortality rate” in animals (Fuchs, 2005). Closely related, there is also a large body of evidence from rodent studies indicating a link between chronic or repeated stress and immunological dysfunctions (Nyuyki, Beiderbeck, Lukas, Neumann, & Reber, 2012; Schmidt et al., 2010). As also discussed earlier, the interaction between an altered regulation of the HPA axis and immune system may constitute predisposing factors for TS. Additionally, to further confirm the relevance of chronic stress paradigms within the field of TS, several studies demonstrated that experimental procedures interfering with the regulation of the HPA axis also altered dopaminergic activity. We recently discussed the interaction between TS and dopaminergic system in an independent paper (Buse et al., 2013). Alternative social stress paradigms in rodents have been developed: for example, the sensory contact model (Kudryavtseva, 1991), the chronic subordinate colony housing (Reber et al., 2007) and the social defeat paradigm in mice (Ader, 1969), or the visible burrow system in rats (Blanchard et al., 1995). The feasibility of these paradigms may be limited by the extensive need of man power, space, and time. All these aspects limit the number of animals per group that can be studied at a time. Reduced group size may hamper experimental validity due to the following aspects: (1) elevated

226

Simone Macrì et al.

interindividual variability in susceptibility to chronic stress and stressinduced symptoms: this may be due to genetic or epigenetic mechanisms and is remarkably elevated both within (Holmes, 2008) and between strains (Zhang-James, Middleton, & Faraone, 2013); (2) specific timing of stress procedures: the age at which the subject is exposed to chronic stress (e.g., in utero, infantile stage, adolescence, adulthood) is of high importance as there are strong indications that individual vulnerability to stress varies across development (Blanchard, McKittrick, & Blanchard, 2001); (3) also individual differences in circadian rhythmicity have to be considered because the consequences of repeated psychosocial stress are dependent on the time of day of stress exposure (Bartlang et al., 2012); (4) recent studies have indicated that some of the effects evoked by social stressors persist even if the stress is discontinued (Sterlemann et al., 2008) and that there is a large dynamic range in the adaptive plasticity of the brain, allowing the animals to adapt behaviorally and physiologically to the previously occurred stressful situation with the progression of time (Buwalda et al., 2005); (5) specific housing conditions may influence the consequences of chronic psychosocial stress in mice. For example, chronic psychosocial stress resulted in persistent deficits in sensorimotor gating (measured by prepulse inhibition) in individually but not in socially housed mice (Adamcio, Havemann-Reinecke, & Ehrenreich, 2009). Both topics are also a matter of debate in the pathophysiology of TS (Buse et al., 2013). Although the aforementioned protocols have been developed with different purposes, some of them directly addressed some of the dependent variables described in the previous sections. For example, we demonstrated that neonatal stress, in the form of corticosterone administration, may persistently modify individual responses in an attentional set-shifting task (Macrı` et al., 2009). Additionally, several authors demonstrated that both prenatal and postnatal stressors may persistently alter behavioral, hormonal, and immune parameters (Kleinhaus et al., 2010; Maccari & Morley-Fletcher, 2007; Morley-Fletcher, Rea, Maccari, & Laviola, 2003). For example, exposure to early life stressors has been shown to relate to an increase in locomotion (Hohmann, Hodges, Beard, & Aneni, 2013). Ultimately, different studies demonstrated that the exposure to stressful environmental conditions may favor the exhibition of behavioral stereotypies (Garner & Mason, 2002; Gross et al., 2012). The expression of cage stereotypies may be the result of the disinhibition of the response selection system represented by the basal ganglia, further supporting the study of abnormal behaviors as potential models of TS-like symptoms (Garner, Meehan, & Mench, 2003).

Animal Models of Tourette Syndrome

227

4. FUTURE PERSPECTIVES In the previous sections, we described two of the main approaches adopted to reproduce TS etiology in laboratory rodents: one that incorporates mutations in specific genes implicated in the development of TS and the other that focuses on specific environmental factors that can trigger and/or exacerbate the pathology. TS is a complex neurodevelopmental disorder and its heterogeneity emphasizes the importance of combining the genetic and environmental components of this pediatric disease in a unique animal model. Specifically, precocious environmental factors, including obstetric complications, have been suggested to contribute to the onset of TS, affecting the structural organization of the brain networks underlying TS symptomatology (Laviola, Adriani, Rea, Aloe, & Alleva, 2004; Leckman & Peterson, 1993; Leckman et al., 1987, 1990; Marco, Macrı`, & Laviola, 2011). Furthermore, clinical evidence supports the hypothesis that emotional variables (psychosocial stressors, anxiety, and frustration) may exert a short-term influence on the periodical fluctuation of tic expression (Conelea & Woods, 2008; Wood et al., 2003), involving the mediation of the HPA axis and the noradrenergic pathway (Chappell et al., 1994; Corbett et al., 2008). Genetic predisposition in association with environmental contribution is thought to play a primary role in the pathogenesis of the disorder during critical periods of brain development. Rodent models challenged with different stressors and/or insults during highly plastic developmental stages may favor the analysis of the environmental factors leading to neuropsychiatric symptoms (Macrı` et al., 2013; Marco et al., 2011). We offer that future studies on animal models of TS shall combine genetic and/or immune aspects with perinatal exposure to stress. For example, the exposure to gestational stressors in rodents could be used to test the role of maternal stress as a risk factor for Tourette. A number of different manipulations in early life have been demonstrated to affect permanently the course of central nervous system development, including neuroendocrine systems such as the HPA axis. Examples of models of prenatal manipulations include maternal stress, exposure to synthetic glucocorticoids, and nutrient restriction (Liu, Li, & Matthews, 2001; Maccari & Morley-Fletcher, 2007; Weinstock, 2001). Stress during gestation has also been associated with disturbances in the immune system. Adult rats exposed to prenatal stress show a significant reduction in CD4þ T and CD8þ T lymphocytes and circulating anti-inflammatory IL-10 cytokines and a significant increase in proinflammatory cytokine IL-1b,

228

Simone Macrì et al.

suggesting a role for stress hormones in the outcome of various immunemediated diseases (Laviola et al., 2004). The environment can exert its influence during different critical developmental stages, including postnatal period, early infancy, and adolescence. In animal models, some of the postnatal manipulations that can be used to permanently modify HPA axis function are constituted by neonatal handling, maternal deprivation, exposure to glucocorticoids, and infections (Bakker, Van Bel, & Heijnen, 2001; Macrı´, Mason, & Wu¨rbel, 2004; Macrı`, Zoratto, & Laviola, 2011; Meaney et al., 2000; Nilsson et al., 2002). For example, we identified a change in the level of natural antibodies against SERT and DAT proteins (5-HT and DATs, respectively) in serum samples of adult mice exposed to different doses of corticosterone during lactation (Macrı` et al., 2009). Psychosocial stress remains one of the most important contextual factors influencing tic severity and potentially modulating the effect of bacterial infections. The mechanism of this interaction needs further investigation and animal models combining autoimmune hypothesis with exposure to developmental stress may constitute a promising research avenue. So far, most of the TS-like symptoms have been modeled in late adolescent/fully adult rodents. We believe that a particular emphasis to the timing of the onset of the symptoms relevant for TS would be necessary. Future studies with a specific focus on the peripubertal stage are needed to clarify the role of the environment in shaping the neurobiological trajectories underlying TS.

ACKNOWLEDGMENTS This study was supported by the EU-FP7 framework project EMTICS under grant agreement n 278367.

REFERENCES Abelson, J. F., Kwan, K. Y., O’Roak, B. J., Baek, D. Y., Stillman, A. A., Morgan, T. M., et al. (2005). Sequence variants in SLITRK1 are associated with Tourette’s syndrome. Science, 310(5746), 317–320. Abrahams, B. S., Tentler, D., Perederiy, J. V., Oldham, M. C., Coppola, G., & Geschwind, D. H. (2007). Genome-wide analyses of human perisylvian cerebral cortical patterning. Proceedings of the National Academy of Sciences of the United States of America, 104(45), 17849–17854. Adamcio, B., Havemann-Reinecke, U., & Ehrenreich, H. (2009). Chronic psychosocial stress in the absence of social support induces pathological pre-pulse inhibition in mice. Behavioural Brain Research, 204(1), 246–249. Ader, R. (1969). Early experiences accelerate maturation of the 24-hour adrenocortical rhythm. Science, 163(3872), 1225–1226.

Animal Models of Tourette Syndrome

229

Albin, R. L., Koeppe, R. A., Bohnen, N. I., Nichols, T. E., Meyer, P., Wernette, K., et al. (2003). Increased ventral striatal monoaminergic innervation in Tourette syndrome. Neurology, 61(3), 310–315. Alzghoul, L., Bortolato, M., Delis, F., Thanos, P. K., Darling, R. D., Godar, S. C., et al. (2012). Altered cerebellar organization and function in monoamine oxidase A hypomorphic mice. Neuropharmacology, 63(7), 1208–1217. American Psychiatric Association (2000). Diagnostic and statistical manual of mental disorders: DSM-IV-TR® (4th ed.). American Psychiatric Publishing. Arnsten, A. F. (2001). Modulation of prefrontal cortical-striatal circuits: Relevance to therapeutic treatments for Tourette syndrome and attention-deficit hyperactivity disorder. Advances in Neurology, 85, 333–341. Arnsten, A. F. T., & Pliszka, S. R. (2011). Catecholamine influences on prefrontal cortical function: Relevance to treatment of attention deficit/hyperactivity disorder and related disorders. Pharmacology, Biochemistry, and Behavior, 99(2), 211–216. Bakker, J. M., Van Bel, F., & Heijnen, C. J. (2001). Neonatal glucocorticoids and the developing brain: Short-term treatment with life-long consequences? Trends in Neurosciences, 24(11), 649–653. Bartlang, M. S., Neumann, I. D., Slattery, D. A., Uschold-Schmidt, N., Kraus, D., HelfrichFo¨rster, C., et al. (2012). Time matters: Pathological effects of repeated psychosocial stress during the active, but not inactive, phase of male mice. The Journal of Endocrinology, 215(3), 425–437. Bartolomucci, A., Palanza, P., Sacerdote, P., Panerai, A. E., Sgoifo, A., Dantzer, R., et al. (2005). Social factors and individual vulnerability to chronic stress exposure. Neuroscience and Biobehavioral Reviews, 29(1), 67–81. Belloso, J. M., Bache, I., Guitart, M., Caballin, M. R., Halgren, C., Kirchhoff, M., et al. (2007). Disruption of the CNTNAP2 gene in a t(7;15) translocation family without symptoms of Gilles de la Tourette syndrome. European Journal of Human Genetics, 15(6), 711–713. Berardelli, A., Curra`, A., Fabbrini, G., Gilio, F., & Manfredi, M. (2003). Pathophysiology of tics and Tourette syndrome. Journal of Neurology, 250(7), 781–787. Bilbo, S. D., Biedenkapp, J. C., Der-Avakian, A., Watkins, L. R., Rudy, J. W., & Maier, S. F. (2005). Neonatal infection-induced memory impairment after lipopolysaccharide in adulthood is prevented via caspase-1 inhibition. Journal of Neuroscience, 25(35), 8000–8009. Bilbo, S. D., & Schwarz, J. M. (2009). Early-life programming of later-life brain and behavior: A critical role for the immune system. Frontiers in Behavioral Neuroscience, 3, 14. Blanchard, R. J., McKittrick, C. R., & Blanchard, D. C. (2001). Animal models of social stress: Effects on behavior and brain neurochemical systems. Physiology and Behavior, 73(3), 261–271. Blanchard, D. C., Spencer, R. L., Weiss, S. M., Blanchard, R. J., McEwen, B., & Sakai, R. R. (1995). Visible burrow system as a model of chronic social stress: Behavioral and neuroendocrine correlates. Psychoneuroendocrinology, 20(2), 117–134. Bliss, J. (1980). Sensory experiences of Gilles de la Tourette syndrome. Archives of General Psychiatry, 37(12), 1343–1347. Bortolato, M., Chen, K., Godar, S. C., Chen, G., Wu, W., Rebrin, I., et al. (2011). Social deficits and perseverative behaviors, but not overt aggression, in MAO-A hypomorphic mice. Neuropsychopharmacology, 36(13), 2674–2688. Bortolato, M., Chen, K., & Shih, J. C. (2008). Monoamine oxidase inactivation: From pathophysiology to therapeutics. Advanced Drug Delivery Reviews, 60(13–14), 1527–1533. Brimberg, L., Benhar, I., Mascaro-Blanco, A., Alvarez, K., Lotan, D., Winter, C., et al. (2012). Behavioral, pharmacological, and immunological abnormalities after streptococcal exposure: A novel rat model of Sydenham chorea and related neuropsychiatric disorders. Neuropsychopharmacology, 37(9), 2076–2087.

230

Simone Macrì et al.

Bronfeld, M., Israelashvili, M., & Bar-Gad, I. (2012). Pharmacological animal models of Tourette syndrome. Neuroscience and Biobehavioral Reviews, 37(6), 1101–1119. Buse, J., Schoenefeld, K., Mu¨nchau, A., & Roessner, V. (2013). Neuromodulation in Tourette syndrome: Dopamine and beyond. Neuroscience and Biobehavioral Reviews, 37(6), 1069–1084. Buwalda, B., Kole, M. H. P., Veenema, A. H., Huininga, M., De Boer, S. F., Korte, S. M., et al. (2005). Long-term effects of social stress on brain and behavior: A focus on hippocampal functioning. Neuroscience and Biobehavioral Reviews, 29(1), 83–97. Campbell, K. M., McGrath, M. J., & Burton, F. H. (1999). Behavioral effects of cocaine on a transgenic mouse model of cortical-limbic compulsion. Brain Research, 833(2), 216–224. Canal, C. E., & Morgan, D. (2012). Head-twitch response in rodents induced by the hallucinogen 2,5-dimethoxy-4-iodoamphetamine: A comprehensive history, a re-evaluation of mechanisms, and its utility as a model. Drug Testing and Analysis, 4(7–8), 556–576. Cardona, F., & Orefici, G. (2001). Group A streptococcal infections and tic disorders in an Italian pediatric population. The Journal of Pediatrics, 138(1), 71–75. Chappell, P., Riddle, M., Anderson, G., Scahill, L., Hardin, M., Walker, D., et al. (1994). Enhanced stress responsivity of Tourette syndrome patients undergoing lumbar puncture. Biological Psychiatry, 36(1), 35–43. Chen, S.-K., Tvrdik, P., Peden, E., Cho, S., Wu, S., Spangrude, G., et al. (2010). Hematopoietic origin of pathological grooming in Hoxb8 mutant mice. Cell, 141(5), 775–785. Cheon, K.-A., Ryu, Y.-H., Namkoong, K., Kim, C.-H., Kim, J.-J., & Lee, J. D. (2004). Dopamine transporter density of the basal ganglia assessed with [123I]IPT SPECT in drug-naive children with Tourette’s disorder. Psychiatry Research, 130(1), 85–95. Cohen, S. C., Leckman, J. F., & Bloch, M. H. (2013). Clinical assessment of Tourette syndrome and tic disorders. Neuroscience and Biobehavioral Reviews, 37(6), 997–1007. Comings, D. E. (1990). Blood serotonin and tryptophan in Tourette syndrome. American Journal of Medical Genetics, 36(4), 418–430. Comings, D. E., Comings, B. G., Muhleman, D., Dietz, G., Shahbahrami, B., Tast, D., et al. (1991). The dopamine D2 receptor locus as a modifying gene in neuropsychiatric disorders. The Journal of the American Medical Association, 266(13), 1793–1800. Conelea, C. A., & Woods, D. W. (2008). The influence of contextual factors on tic expression in Tourette’s syndrome: A review. Journal of Psychosomatic Research, 65(5), 487–496. Conelea, C. A., Woods, D. W., & Brandt, B. C. (2011). The impact of a stress induction task on tic frequencies in youth with Tourette Syndrome. Behaviour Research and Therapy, 49(8), 492–497. Corbett, B. A., Mendoza, S. P., Baym, C. L., Bunge, S. A., & Levine, S. (2008). Examining cortisol rhythmicity and responsivity to stress in children with Tourette syndrome. Psychoneuroendocrinology, 33(6), 810–820. Dale, R. C., Candler, P. M., Church, A. J., Wait, R., Pocock, J. M., & Giovannoni, G. (2006). Neuronal surface glycolytic enzymes are autoantigen targets in post-streptococcal autoimmune CNS disease. Journal of Neuroimmunology, 172(1–2), 187–197. Dantzer, R. (1986). Behavioral, physiological and functional aspects of stereotyped behavior: A review and a re-interpretation. Journal of Animal Science, 62(6), 1776–1786. Deng, H., Gao, K., & Jankovic, J. (2012). The genetics of Tourette syndrome. Nature Reviews Neurology, 8(4), 203–213. Dı´az-Anzaldu´a, A., Joober, R., Rivie`re, J.-B., Dion, Y., Lespe´rance, P., Richer, F., et al. (2004). Tourette syndrome and dopaminergic genes: A family-based association study in the French Canadian founder population. Molecular Psychiatry, 9(3), 272–277. Dursun, S. M., & Handley, S. L. (1992). Serotonin, eating disorders, and HIV infection. The British Journal of Psychiatry, 160, 866–867.

Animal Models of Tourette Syndrome

231

Ercan-Sencicek, A. G., Stillman, A. A., Ghosh, A. K., Bilguvar, K., O’Roak, B. J., Mason, C. E., et al. (2010). L-histidine decarboxylase and Tourette’s syndrome. The New England Journal of Medicine, 362(20), 1901–1908. Felling, R. J., & Singer, H. S. (2011). Neurobiology of tourette syndrome: Current status and need for further investigation. Journal of Neuroscience, 31(35), 12387–12395. Foote, S. L., Bloom, F. E., & Aston-Jones, G. (1983). Nucleus locus ceruleus: New evidence of anatomical and physiological specificity. Physiological Reviews, 63(3), 844–914. Fuchs, E. (2005). Social stress in tree shrews as an animal model of depression: An example of a behavioral model of a CNS disorder. CNS Spectrums, 10(3), 182–190. Ganos, C., Roessner, V., & Mu¨nchau, A. (2013). The functional anatomy of Gilles de la Tourette syndrome. Neuroscience and Biobehavioral Reviews, 37(6), 1050–1062. Garner, J. P., & Mason, G. J. (2002). Evidence for a relationship between cage stereotypies and behavioural disinhibition in laboratory rodents. Behavioural Brain Research, 136(1), 83–92. Garner, J. P., Meehan, C. L., & Mench, J. A. (2003). Stereotypies in caged parrots, schizophrenia and autism: Evidence for a common mechanism. Behavioural Brain Research, 145(1–2), 125–134. Greer, J. M., & Capecchi, M. R. (2002). Hoxb8 is required for normal grooming behavior in mice. Neuron, 33(1), 23–34. Gross, A. N., Richter, S. H., Engel, A. K. J., & Wu¨rbel, H. (2012). Cage-induced stereotypies, perseveration and the effects of environmental enrichment in laboratory mice. Behavioural Brain Research, 234(1), 61–68. Haas, H. L., Sergeeva, O. A., & Selbach, O. (2008). Histamine in the nervous system. Physiological Reviews, 88(3), 1183–1241. Hallett, J. J., Harling-Berg, C. J., Knopf, P. M., Stopa, E. G., & Kiessling, L. S. (2000). Anti-striatal antibodies in Tourette syndrome cause neuronal dysfunction. Journal of Neuroimmunology, 111(1–2), 195–202. Heinrichs, M., & Gaab, J. (2007). Neuroendocrine mechanisms of stress and social interaction: Implications for mental disorders. Current Opinion in Psychiatry, 20(2), 158–162. Hoekstra, P. J., Dietrich, A., Edwards, M. J., Elamin, I., & Martino, D. (2013). Environmental factors in Tourette syndrome. Neuroscience and Biobehavioral Reviews, 37(6), 1040–1049. Hoffman, K. L., Hornig, M., Yaddanapudi, K., Jabado, O., & Lipkin, W. I. (2004). A murine model for neuropsychiatric disorders associated with group A beta-hemolytic streptococcal infection. Journal of Neuroscience, 24(7), 1780–1791. Hohmann, C. F., Hodges, A., Beard, N., & Aneni, J. (2013). Effects of brief stress exposure during early postnatal development in balb/CByJ mice: I. Behavioral characterization. Developmental Psychobiology, 55(3), 283–293. Holmes, A. (2008). Genetic variation in cortico-amygdala serotonin function and risk for stress-related disease. Neuroscience and Biobehavioral Reviews, 32(7), 1293–1314. Hornig, M., & Lipkin, W. I. (2013). Immune-mediated animal models of Tourette syndrome. Neuroscience and Biobehavioral Reviews, 37(6), 1120–1138. Huber, L., Ferdin, M., Holzmann, J., Stubbusch, J., & Rohrer, H. (2012). HoxB8 in noradrenergic specification and differentiation of the autonomic nervous system. Developmental Biology, 363(1), 219–233. Jankovic, J., & Kurlan, R. (2011). Tourette syndrome: Evolving concepts. Movement Disorders, 26(6), 1149–1156. Jung, J., Jackson, S. R., Parkinson, A., & Jackson, G. M. (2012). Cognitive control over motor output in Tourette syndrome. Neuroscience and Biobehavioral Reviews, 37(6), 1016–1025. Kajiwara, Y., Buxbaum, J. D., & Grice, D. E. (2009). SLITRK1 binds 14-3-3 and regulates neurite outgrowth in a phosphorylation-dependent manner. Biological Psychiatry, 66(10), 918–925.

232

Simone Macrì et al.

Kang, H., Zhang, Y.-F., Jiao, F.-Y., Guo, X.-Y., & Gao, X.-M. (2009). Efficacy of clonidine transdermal patch for treatment of Tourette’s syndrome in children. Chinese Journal of Contemporary Pediatrics, 11(7), 537–539. Karagiannidis, I., Dehning, S., Sandor, P., Tarnok, Z., Rizzo, R., Wolanczyk, T., et al. (2013). Support of the histaminergic hypothesis in Tourette Syndrome: Association of the histamine decarboxylase gene in a large sample of families. Journal of Medical Genetics. Katayama, K., Yamada, K., Ornthanalai, V. G., Inoue, T., Ota, M., Murphy, N. P., et al. (2010). Slitrk1-deficient mice display elevated anxiety-like behavior and noradrenergic abnormalities. Molecular Psychiatry, 15(2), 177–184. Kawikova, I., Grady, B. P. X., Tobiasova, Z., Zhang, Y., Vojdani, A., Katsovich, L., et al. (2010). Children with Tourette’s syndrome may suffer immunoglobulin A dysgammaglobulinemia: Preliminary report. Biological Psychiatry, 67(7), 679–683. Kerr, D., Krishnan, C., Pucak, M. L., & Carmen, J. (2005). The immune system and neuropsychiatric diseases. International Review of Psychiatry, 17(6), 443–449. Kiessling, L. S., Marcotte, A. C., & Culpepper, L. (1993). Antineuronal antibodies in movement disorders. Pediatrics, 92(1), 39–43. Kiessling, L. S., Marcotte, A. C., & Culpepper, L. (1994). Antineuronal antibodies: Tics and obsessive-compulsive symptoms. Journal of Developmental and Behavioral Pediatrics, 15(6), 421–425. Kleinhaus, K., Steinfeld, S., Balaban, J., Goodman, L., Craft, T. S., Malaspina, D., et al. (2010). Effects of excessive glucocorticoid receptor stimulation during early gestation on psychomotor and social behavior in the rat. Developmental Psychobiology, 52(2), 121–132. Korff, S., & Harvey, B. H. (2006). Animal models of obsessive-compulsive disorder: Rationale to understanding psychobiology and pharmacology. The Psychiatric Clinics of North America, 29(2), 371–390. Kubota, Y., Ito, C., Sakurai, E., Sakurai, E., Watanabe, T., & Ohtsu, H. (2002). Increased methamphetamine-induced locomotor activity and behavioral sensitization in histamine-deficient mice. Journal of Neurochemistry, 83(4), 837–845. Kudryavtseva, N. N. (1991). A sensory contact model for the study of aggressive and submissive behavior in male mice. Aggressive Behavior, 17(5), 285–291. Kurlan, R., McDermott, M. P., Deeley, C., Como, P. G., Brower, C., Eapen, S., et al. (2001). Prevalence of tics in schoolchildren and association with placement in special education. Neurology, 57(8), 1383–1388. Laviola, G., Adriani, W., Rea, M., Aloe, L., & Alleva, E. (2004). Social withdrawal, neophobia, and stereotyped behavior in developing rats exposed to neonatal asphyxia. Psychopharmacology, 175(2), 196–205. Lawson-Yuen, A., Saldivar, J.-S., Sommer, S., & Picker, J. (2008). Familial deletion within NLGN4 associated with autism and Tourette syndrome. European Journal of Human Genetics, 16(5), 614–618. Leckman, James F. (2012). Tic disorders. British Medical Journal, 344, d7659. Leckman, James F., Bloch, M. H., Smith, M. E., Larabi, D., & Hampson, M. (2010). Neurobiological substrates of Tourette’s disorder. Journal of Child and Adolescent Psychopharmacology, 20(4), 237–247. Leckman, J. F., Dolnansky, E. S., Hardin, M. T., Clubb, M., Walkup, J. T., Stevenson, J., et al. (1990). Perinatal factors in the expression of Tourette’s syndrome: An exploratory study. Journal of the American Academy of Child and Adolescent Psychiatry, 29(2), 220–226. Leckman, J. F., Hardin, M. T., Riddle, M. A., Stevenson, J., Ort, S. I., & Cohen, D. J. (1991). Clonidine treatment of Gilles de la Tourette’s syndrome. Archives of General Psychiatry, 48(4), 324–328. Leckman, J. F., & Peterson, B. S. (1993). The pathogenesis of Tourette’s syndrome: Epigenetic factors active in early CNS development. Biological Psychiatry, 34(7), 425–427.

Animal Models of Tourette Syndrome

233

Leckman, J. F., Price, R. A., Walkup, J. T., Ort, S., Pauls, D. L., & Cohen, D. J. (1987). Nongenetic factors in Gilles de la Tourette’s syndrome. Archives of General Psychiatry, 44(1), 100. Lee, C.-C., Chou, I.-C., Tsai, C.-H., Wang, T.-R., Li, T.-C., & Tsai, F.-J. (2005). Dopamine receptor D2 gene polymorphisms are associated in Taiwanese children with Tourette syndrome. Pediatric Neurology, 33(4), 272–276. Lin, H., Williams, K. A., Katsovich, L., Findley, D. B., Grantz, H., Lombroso, P. J., et al. (2010). Streptococcal upper respiratory tract infections and psychosocial stress predict future tic and obsessive-compulsive symptom severity in children and adolescents with Tourette syndrome and obsessive-compulsive disorder. Biological Psychiatry, 67(7), 684–691. Linhoff, M. W., Laure´n, J., Cassidy, R. M., Dobie, F. A., Takahashi, H., Nygaard, H. B., et al. (2009). An unbiased expression screen for synaptogenic proteins identifies the LRRTM protein family as synaptic organizers. Neuron, 61(5), 734–749. Liu, L., Li, A., & Matthews, S. G. (2001). Maternal glucocorticoid treatment programs HPA regulation in adult offspring: Sex-specific effects. American Journal of Physiology, 280(5), 729–739. Lombroso, P. J., & Scahill, L. (2008). Tourette syndrome and obsessive-compulsive disorder. Brain and Development, 30(4), 231–237. Maccari, S., & Morley-Fletcher, S. (2007). Effects of prenatal restraint stress on the hypothalamus-pituitary-adrenal axis and related behavioural and neurobiological alterations. Psychoneuroendocrinology, 32(Suppl. 1), S10–S15. Macrı`, S., Granstrem, O., Shumilina, M., Gomes, Antunes, dos Santos, F. J., Berry, A., et al. (2009). Resilience and vulnerability are dose-dependently related to neonatal stressors in mice. Hormones and Behavior, 56(4), 391–398. Macrı´, S., Mason, G. J., & Wu¨rbel, H. (2004). Dissociation in the effects of neonatal maternal separations on maternal care and the offspring’s HPA and fear responses in rats. The European Journal of Neuroscience, 20(4), 1017–1024. Macrı`, S., Proietti Onori, M., & Laviola, G. (2013). Theoretical and practical considerations behind the use of laboratory animals for the study of Tourette syndrome. Neuroscience and Biobehavioral Reviews, 37(6), 1085–1100. Macrı`, S., Zoratto, F., & Laviola, G. (2011). Early-stress regulates resilience, vulnerability and experimental validity in laboratory rodents through mother-offspring hormonal transfer. Neuroscience and Biobehavioral Reviews, 35(7), 1534–1543. Marco, E. M., Macrı`, S., & Laviola, G. (2011). Critical age windows for neurodevelopmental psychiatric disorders: Evidence from animal models. Neurotoxicity Research, 19(2), 286–307. Martino, D., Chiarotti, F., Buttiglione, M., Cardona, F., Creti, R., Nardocci, N., et al. (2011). The relationship between group A streptococcal infections and Tourette syndrome: A study on a large service-based cohort. Developmental Medicine and Child Neurology, 53(10), 951–957. Martino, D., & Giovannoni, G. (2004). Antibasal ganglia antibodies and their relevance to movement disorders. Current Opinion in Neurology, 17(4), 425–432. Mas´li nska, D. (2001). The cytokine network and interleukin-15 (IL-15) in brain development. Folia Neuropathologica/Association of Polish Neuropathologists and Medical Research Centre, Polish Academy of Sciences, 39(2), 43–47. McKenzie, G. M., Gordon, R. J., & Viik, K. (1972). Some biochemical and behavioural correlates of a possible animal model of human hyperkinetic syndromes. Brain Research, 47(2), 439–456. Meaney, M. J., Diorio, J., Francis, D., Weaver, S., Yau, J., Chapman, K., et al. (2000). Postnatal handling increases the expression of cAMP-inducible transcription factors in the rat hippocampus: The effects of thyroid hormones and serotonin. The Journal of neuroscience, 20(10), 3926–3935.

234

Simone Macrì et al.

Mell, L. K., Davis, R. L., & Owens, D. (2005). Association between streptococcal infection and obsessive-compulsive disorder, Tourette’s syndrome, and tic disorder. Pediatrics, 116(1), 56–60. Morley-Fletcher, S., Rea, M., Maccari, S., & Laviola, G. (2003). Environmental enrichment during adolescence reverses the effects of prenatal stress on play behaviour and HPA axis reactivity in rats. The European Journal of Neuroscience, 18(12), 3367–3374. Morris, C. M., Pardo-Villamizar, C., Gause, C. D., & Singer, H. S. (2009). Serum autoantibodies measured by immunofluorescence confirm a failure to differentiate PANDAS and Tourette syndrome from controls. Journal of the Neurological Sciences, 276(1–2), 45–48. Nilsson, C., Jennische, E., Ho, H.-P., Eriksson, E., Bjorntorp, P., & Holmang, A. (2002). Postnatal endotoxin exposure results in increased insulin sensitivity and altered activity of neuroendocrine axes in adult female rats. European Journal of Endocrinology/European Federation of Endocrine Societies, 146(2), 251–260. Nordstrom, E. J., & Burton, F. H. (2002). A transgenic model of comorbid Tourette’s syndrome and obsessive-compulsive disorder circuitry. Molecular Psychiatry, 7(6), 617–625, 524. Nyuyki, K. D., Beiderbeck, D. I., Lukas, M., Neumann, I. D., & Reber, S. O. (2012). Chronic subordinate colony housing (CSC) as a model of chronic psychosocial stress in male rats. PLoS one, 7(12), e52371. O’Rourke, J. A., Scharf, J. M., Yu, D., & Pauls, D. L. (2009). The genetics of Tourette syndrome: A review. Journal of Psychosomatic Research, 67(6), 533–545. Paschou, P. (2013). The genetic basis of Gilles de la Tourette Syndrome. Neuroscience and Biobehavioral Reviews, 37(6), 1026–1039. Patel, C., Cooper-Charles, L., McMullan, D. J., Walker, J. M., Davison, V., & Morton, J. (2011). Translocation breakpoint at 7q31 associated with tics: Further evidence for IMMP2L as a candidate gene for Tourette syndrome. European Journal of Human Genetics, 19(6), 634–639. Patel, S., & Slater, P. (1987). Analysis of the brain regions involved in myoclonus produced by intracerebral picrotoxin. Neuroscience, 20(2), 687–693. Pauls, D. L. (1992). Issues in genetic linkage studies of Tourette syndrome. Phenotypic spectrum and genetic model parameters. Advances in Neurology, 58, 151–157. Pellow, S., Chopin, P., File, S. E., & Briley, M. (1985). Validation of open: Closed arm entries in an elevated plus-maze as a measure of anxiety in the rat. Journal of Neuroscience Methods, 14(3), 149–167. Perry, V. H., Newman, T. A., & Cunningham, C. (2003). The impact of systemic infection on the progression of neurodegenerative disease. Nature Reviews Neuroscience, 4(2), 103–112. Petek, E., Windpassinger, C., Vincent, J. B., Cheung, J., Boright, A. P., Scherer, S. W., et al. (2001). Disruption of a novel gene (IMMP2L) by a breakpoint in 7q31 associated with Tourette syndrome. American Journal of Human Genetics, 68(4), 848–858. Poliak, S., Gollan, L., Martinez, R., Custer, A., Einheber, S., Salzer, J. L., et al. (1999). Caspr2, a new member of the neurexin superfamily, is localized at the juxtaparanodes of myelinated axons and associates with K þ channels. Neuron, 24(4), 1037–1047. Ponzio, N. M., Servatius, R., Beck, K., Marzouk, A., & Kreider, T. (2007). Cytokine levels during pregnancy influence immunological profiles and neurobehavioral patterns of the offspring. Annals of the New York Academy of Sciences, 1107, 118–128. Price, R. A., Kidd, K. K., Cohen, D. J., Pauls, D. L., & Leckman, J. F. (1985). A twin study of Tourette syndrome. Archives of General Psychiatry, 42(8), 815–820. Proenca, C. C., Gao, K. P., Shmelkov, S. V., Rafii, S., & Lee, F. S. (2011). Slitrks as emerging candidate genes involved in neuropsychiatric disorders. Trends in Neurosciences, 34(3), 143–153.

Animal Models of Tourette Syndrome

235

Randolph, C., Hyde, T. M., Gold, J. M., Goldberg, T. E., & Weinberger, D. R. (1993). Tourette’s syndrome in monozygotic twins. Relationship of tic severity to neuropsychological function. Archives of Neurology, 50(7), 725–728. Reber, S. O., Birkeneder, L., Veenema, A. H., Obermeier, F., Falk, W., Straub, R. H., et al. (2007). Adrenal insufficiency and colonic inflammation after a novel chronic psychosocial stress paradigm in mice: Implications and mechanisms. Endocrinology, 148(2), 670–682. Robertson, M. M. (2008). The prevalence and epidemiology of Gilles de la Tourette syndrome. Part 1: The epidemiological and prevalence studies. Journal of Psychosomatic Research, 65(5), 461–472. Robertson, H. R., & Feng, G. (2011). Annual research review: Transgenic mouse models of childhood-onset psychiatric disorders. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 52(4), 442–475. Scahill, L., Riddle, M. A., King, R. A., Hardin, M. T., Rasmusson, A., Makuch, R. W., et al. (1997). Fluoxetine has no marked effect on tic symptoms in patients with Tourette’s syndrome: A double-blind placebo-controlled study. Journal of Child and Adolescent Psychopharmacology, 7(2), 75–85. Schmidt, D., Reber, S. O., Botteron, C., Barth, T., Peterlik, D., Uschold, N., et al. (2010). Chronic psychosocial stress promotes systemic immune activation and the development of inflammatory Th cell responses. Brain, Behavior, and Immunity, 24(7), 1097–1104. Schmidt, M. V., Sterlemann, V., Ganea, K., Liebl, C., Alam, S., Harbich, D., et al. (2007). Persistent neuroendocrine and behavioral effects of a novel, etiologically relevant mouse paradigm for chronic social stress during adolescence. Psychoneuroendocrinology, 32(5), 417–429. Schrag, A., Gilbert, R., Giovannoni, G., Robertson, M. M., Metcalfe, C., & Ben-Shlomo, Y. (2009). Streptococcal infection, Tourette syndrome, and OCD: Is there a connection? Neurology, 73(16), 1256–1263. Seuchter, S. A., Hebebrand, J., Klug, B., Knapp, M., Lehmkuhl, G., Poustka, F., et al. (2000). Complex segregation analysis of families ascertained through Gilles de la Tourette syndrome. Genetic Epidemiology, 18(1), 33–47. Singer, H. S., Butler, I. J., Tune, L. E., Seifert, W. E., & Coyle, J. T. (1982). Dopaminergic dsyfunction in Tourette syndrome. Annals of Neurology, 12(4), 361–366. Singer, Harvey S., Gause, C., Morris, C., & Lopez, P. (2008). Serial immune markers do not correlate with clinical exacerbations in pediatric autoimmune neuropsychiatric disorders associated with streptococcal infections. Pediatrics, 121(6), 1198–1205. Singer, H. S., Giuliano, J. D., Hansen, B. H., Hallett, J. J., Laurino, J. P., Benson, M., et al. (1998). Antibodies against human putamen in children with Tourette syndrome. Neurology, 50(6), 1618–1624. Singer, Harvey S., Hong, J. J., Yoon, D. Y., & Williams, P. N. (2005). Serum autoantibodies do not differentiate PANDAS and Tourette syndrome from controls. Neurology, 65(11), 1701–1707. Singer, Harvey S., & Minzer, K. (2003). Neurobiology of Tourette’s syndrome: Concepts of neuroanatomic localization and neurochemical abnormalities. Brain and Development, 25(Suppl. 1), S70–S84. Smith, S. E. P., Li, J., Garbett, K., Mirnics, K., & Patterson, P. H. (2007). Maternal immune activation alters fetal brain development through interleukin-6. The Journal of Neuroscience, 27(40), 10695–10702. State, M. W. (2011). The genetics of Tourette disorder. Current Opinion in Genetics and Development, 21(3), 302–309. Sterlemann, V., Ganea, K., Liebl, C., Harbich, D., Alam, S., Holsboer, F., et al. (2008). Long-term behavioral and neuroendocrine alterations following chronic social stress in mice: Implications for stress-related disorders. Hormones and Behavior, 53(2), 386–394.

236

Simone Macrì et al.

Stillman, A. A., Krsnik, Z., Sun, J., Rasin, M.-R., State, M. W., Sestan, N., et al. (2009). Developmentally regulated and evolutionarily conserved expression of SLITRK1 in brain circuits implicated in Tourette syndrome. The Journal of Comparative Neurology, 513(1), 21–37. Swain, J. E., Scahill, L., Lombroso, P. J., King, R. A., & Leckman, J. F. (2007). Tourette syndrome and tic disorders: A decade of progress. Journal of the American Academy of Child and Adolescent Psychiatry, 46(8), 947–968. Swedo, S. E., Leonard, H. L., Garvey, M., Mittleman, B., Allen, A. J., Perlmutter, S., et al. (1998). Pediatric autoimmune neuropsychiatric disorders associated with streptococcal infections: Clinical description of the first 50 cases. The American Journal of Psychiatry, 155(2), 264–271. Taylor, J. R., Morshed, S. A., Parveen, S., Mercadante, M. T., Scahill, L., Peterson, B. S., et al. (2002). An animal model of Tourette’s syndrome. The American Journal of Psychiatry, 159(4), 657–660. Tse, S. Y., & Wei, E. T. (1986). Inhibition of the shake response in rats by adenosine and 2-chloroadenosine. Psychopharmacology, 90(3), 322–326. Van Der Staay, F. J., Arndt, S. S., & Nordquist, R. E. (2009). Evaluation of animal models of neurobehavioral disorders. Behavioral and Brain Functions, 5(1), 11. Verkerk, A., Mathews, C., Joosse, M., Eussen, B., Heutink, P., & Oostra, B. (2003). CNTNAP2 is disrupted in a family with Gilles de la Tourette syndrome and obsessive compulsive disorder. Genomics, 82(1), 1–9. Weinstock, M. (2001). Alterations induced by gestational stress in brain morphology and behaviour of the offspring. Progress in Neurobiology, 65(5), 427–451. Weisman, H., Qureshi, I. A., Leckman, J. F., Scahill, L., & Bloch, M. H. (2013). Systematic review: Pharmacological treatment of tic disorders—Efficacy of antipsychotic and alpha-2 adrenergic agonist agents. Neuroscience and Biobehavioral Reviews, 37(6), 1162–1171. Wendlandt, J. T., Grus, F. H., Hansen, B. H., & Singer, H. S. (2001). Striatal antibodies in children with Tourette’s syndrome: Multivariate discriminant analysis of IgG repertoires. Journal of Neuroimmunology, 119(1), 106–113. Willis, C. L. (2012). Imaging in vivo astrocyte/endothelial cell interactions at the blood-brain barrier. Methods in Molecular Biology, 814, 515–529. Willner, P. (2005). Chronic mild stress (CMS) revisited: Consistency and behaviouralneurobiological concordance in the effects of CMS. Neuropsychobiology, 52(2), 90–110. Wolf, S. S., Jones, D. W., Knable, M. B., Gorey, J. G., Lee, K. S., Hyde, T. M., et al. (1996). Tourette syndrome: Prediction of phenotypic variation in monozygotic twins by caudate nucleus D2 receptor binding. Science, 273(5279), 1225–1227. Wood, B. L., Klebba, K., Gbadebo, O., Lichter, D., Kurlan, R., & Miller, B. (2003). Pilot study of effect of emotional stimuli on tic severity in children with Tourette’s syndrome. Movement Disorders, 18(11), 1392–1395. Yaddanapudi, K., Hornig, M., Serge, R., De Miranda, J., Baghban, A., Villar, G., et al. (2010). Passive transfer of streptococcus-induced antibodies reproduces behavioral disturbances in a mouse model of pediatric autoimmune neuropsychiatric disorders associated with streptococcal infection. Molecular Psychiatry, 15(7), 712–726. Zalcman, S., Green-Johnson, J. M., Murray, L., Nance, D. M., Dyck, D., Anisman, H., et al. (1994). Cytokine-specific central monoamine alterations induced by interleukin-1, -2 and -6. Brain Research, 643(1–2), 40–49. Zalcman, S., Murray, L., Dyck, D. G., Greenberg, A. H., & Nance, D. M. (1998). Interleukin-2 and -6 induce behavioral-activating effects in mice. Brain Research, 811(1–2), 111–121.

Animal Models of Tourette Syndrome

237

Zalcman, S. S., Patel, A., Mohla, R., Zhu, Y., & Siegel, A. (2012). Soluble cytokine receptors (sIL-2Ra, sIL-2Rb) induce subunit-specific behavioral responses and accumulate in the cerebral cortex and basal forebrain. PLoS one, 7(4), e36316. Zhang-James, Y., Middleton, F. A., & Faraone, S. V. (2013). Genetic architecture of Wistar Kyoto rats (WKY) and Spontaneous hypertensive Rats (SHR) substrains from difference sources. Physiological Genomics, 45, 528–538. Zhuang, X., Oosting, R. S., Jones, S. R., Gainetdinov, R. R., Miller, G. W., Caron, M. G., et al. (2001). Hyperactivity and impaired response habituation in hyperdopaminergic mice. Proceedings of the National Academy of Sciences of the United States of America, 98(4), 1982–1987.

Animal models recapitulating the multifactorial origin of Tourette syndrome.

Tourette Syndrome (TS) is a neurological disorder characterized by motor and phonic tics affecting approximately 1% of the pediatric population. Behav...
462KB Sizes 0 Downloads 0 Views