Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2017 May 3.
Published in final edited form as: Neuroscience. 2015 Aug 24;321:163–188. doi: 10.1016/j.neuroscience.2015.08.041

Animal Models of Bipolar Mania: The Past, Present and Future

Ryan W Logan 1, Colleen A McClung 1,*
PMCID: PMC4766066  NIHMSID: NIHMS721486  PMID: 26314632

Abstract

Bipolar disorder (BD) is the sixth leading cause of disability in the world according to the World Health Organization and affects nearly 6 million (~2.5% of the population) adults in the United State alone each year. BD is primarily characterized by mood cycling of depressive (e.g., helplessness, reduced energy and activity, and anhedonia) and manic (e.g., increased energy and hyperactivity, reduced need for sleep, impulsivity, reduced anxiety and depression), episodes. The following review describes several animal models of bipolar mania with a focus on more recent findings using genetically modified mice, including several with the potential of investigating the mechanisms underlying ‘mood’ cycling (or behavioral switching in rodents). We discuss whether each of these models satisfy criteria of validity (i.e., face, predictive, and construct), while highlighting their strengths and limitations. Animal models are helping to address critical questions related to pathophysiology of bipolar mania, in an effort to more clearly define necessary targets of first-line medications, lithium and valproic acid, and to discover novel mechanisms with the hope of developing more effective therapeutics. Future studies will leverage new technologies and strategies for integrating animal and human data to reveal important insights into the etiology, pathophysiology, and treatment of BD.

Introduction

Bipolar disorder (BD) is a complex disease defined by periods of both mania and depression with euthymic or normal mood states between episodes. Manic episodes can consist of hyperactivity, elevated mood or agitation, racing thoughts, reckless behavior, little need for sleep, and sometimes psychosis. Depressive episodes as defined by the DSM V can include persistent sadness, fatigue, eating disturbances, sleep disturbances, suicidal thoughts, guilt and social withdrawal. The cause of BD is unknown and may involve both genetic and environmental factors (Shinozaki and Potash, 2014). The mood stabilizing therapeutic effects of lithium (a salt) and valproate (an anticonvulsant) were discovered by accident and in the absence of any significant mechanistic understanding of BD (Can et al., 2014). While current treatments are generally effective for the reversal of manic episodes and preventing future episodes, these medications have limited, if any, efficacy on their own in the acute treatment of depressive episodes (McInerney and Kennedy, 2014). Moreover, standard antidepressant medications used either as monotherapies, or in conjunction with mood stabilizers or antipsychotics, are generally ineffective for treating depressive episodes, and may induce mood switching in a subset of patients with rapid cycling BD (De Wilde and Doogan, 1982, Himmelhoch et al., 1982, Gijsman et al., 2004, Amsterdam and Shults, 2005, Sachs et al., 2007, McElroy et al., 2010, Sidor and Macqueen, 2011, McInerney and Kennedy, 2014). Although there are a few studies suggesting therapeutic efficacy of antidepressant monotherapy for bipolar depression, current recommendations indicate antidepressants be used only in combination with mood stabilizers if those first line medications fail (McInerney and Kennedy, 2014). Despite their effectiveness in the treatment of mania, chronic treatment with current mood stabilizing drugs often result in serious side effects that make patient compliance difficult and burdensome. Our understanding of the etiological mechanisms of BD is poor. Therefore, the use of appropriate animal models should ultimately aid in the development of novel, potentially more efficacious treatments for this complex disorder.

The screening of compound libraries in animal models could also prove fruitful in the search for new medications. Most of the early mechanistic studies of BD in animals have focused on changes that occur following administration of lithium or other agents often on animals that are comparatively “normal”, which may limit the interpretability and applicability of these studies since lithium has very little effect on healthy individuals while having therapeutic effects on those suffering from mania (Calil et al., 1990). Therefore the changes in the brain that occur in wild-type rodents may not represent the same changes that occur in a rodent displaying mania-like behavior.

Evaluating animal models of BD: Issues of face, predictive, and construct validity

The use of rodent models that cut across multiple types of validity are vitally important to our understanding of psychiatric diseases and the search for better treatments. Face validity refers to the extent to which an animal model recapitulates important features of the human disease, such as neuroanatomical, biochemical, and/or behavioral phenotypes. There are few, if any, neurobiological pathologies that are known, with any certainty, to be biomarkers of the psychiatric disease. Even single behaviors may not completely map onto the clinical symptomology they are intended to model, especially when considering diagnoses are ‘inaccurate’ and ‘imprecise’ due to poor diagnostic criteria. Face validity is typically considered to be the requirement for reasonable ‘symptomatic’ homology or overlap with diagnostic criteria. This could pose a problem when diagnostic criteria are poorly defined or based solely on phenomenology (Chesler and Logan, 2012). Another issue arises when adhering to strict definitions of face validity—excessive anthropomorphization of rodent behavior. Most, if not all, of the animal models discussed here recapitulate at least one, but usually several, of the core behavioral phenotypes of bipolar mania, such as reduced depression, and increased impulsivity, or risk-taking behaviors. Face validity means also recapitulating anatomical, biochemical, or neuropathological features of the disease. Another form of this is sometimes referred to as pathological validity, whereby animal models recapitulate observed postmortem pathology from brains or other tissues found in the human disease. Human postmortem studies, along with the advent of newer technologies to derive neurons from pluripotent stem cells collected from patients with disease, should provide a clearer understanding of the mechanisms involved in the pathophysiology of BD.

Construct validity is the degree to which a test measures what it claims to measure. It often refers to the relevance or translational nature of the methods by which a model is constructed. This might be achieved by exposing an animal to an environmental risk factor, knocking in a genetic mutation, or by altering the function of proteins, signaling pathways, or neurocircuitry that is postulated to be involved in the pathogenesis. However, we currently do not know with much certainty the disease-causing genes, proteins, or neurocircuitry for many mental illnesses, as most present with highly complex etiologies. It is true that many of the variants identified in large-scale GWAS studies in those with a psychiatric disease are of small effect. Transgenic animals expressing any of these common variants of small effect do provide some translational utility, although we would argue that these results are approached with some awareness of their limitations. Construct validity of any animal model of psychiatric disease should be interpreted with caution because our understanding of how genes correlate with disease phenotypes is continuously evolving. Throughout the review, we allude to type of validity and primarily highlight facets of each model with importance to bipolar mania (Table).

Predictive validity is usually defined as the extent to which a model responds to a treatment that is efficacious in humans. Animal models of mania are usually administered medications used for the treatment of BD, most commonly lithium or valproic acid, to investigate whether these compounds normalize or reverse these mania-like phenotypes. This too is complicated because of there are multiple factors involved in the interpretation of behavioral data. For example, an animal may show less hyperactivity in a particular assay because the medication was successful in reversing the “manic-like” neuronal activity, or because the animal is instead spending time compulsively grooming, or is sick or heavily sedated. Newer, highly validated cross-species translational approaches should improve the success for identifying novel, more effective therapeutics, where classical behavioral assays fail too often.

In this review we will discuss some of the pharmacological, environmental and genetic models that have been used over the years to study BD. We will discuss the rationale for these particular models as well as provide critiques related to their validity. The vast majority of these models attempt to recapitulate specific aspects related to mania, like hyperactivity or several different features of mania (Table 1). Although more recently, newer models are attempting to reproduce ‘mood’ switching, the hallmark feature of BD, such that we can better understand the precipitation of these episodes (Young and Dulcis, 2015). While no animal model will ever be able to fully, and precisely reproduce all the features of this complex disease, the use of multiple models could still be highly informative (Tables 2,3).

Table 1.

Modeling human bipolar mania in rodents

Behavioral endophenotype Rodent paradigm
Reduced anxiety, increased novel exploration Elevated plus maze; dark-light box; open field arena; social interaction test, novelty suppressed feeding
Reduced depression Forced swim test, tail suspension test, learned helplessness test
Impaired sensory processing Paired-pulse inhibition (PPI)—sensorimotor gating*
Risk-taking behavior Iowa Gambling Task (IGT)*
Impulsivity IGT*
Impaired decision-making IGT*
Psychostimulant-induced hyperactivity Cocaine or amphetamine injection, repeated injections for locomotor sensitization
Hedonia Sucrose preference; cocaine or amphetamine self-administration*; intracranial self-stimulation (ICSS) of medial forebrain bundle; two bottle free-choice alcohol drinking; cocaine conditioned place preference (CPP)
Hyperactivity Open field arena; homecage monitors; wheel-running; behavioral pattern monitor (BPM)*
Increased sexual activity Mounts, intermission, and ejaculation events
Increased goal-directed BPM*
Repetitive movements BPM*
Aggressive behavior Resident-intruder test
Reduced or disrupted sleep EEG; circadian activity; sleep-wake monitoring*
Disrupted circadian rhythms Circadian wheel-running activity; temporal patterns of behavior
*

Demonstrated direct translational relevance to human bipolar disorder

Table 2.

Pharmacological and environmental mouse models of bipolar mania

Validity
Manipulation Face Predictive Construct
Amphetamine-induced hyperactivity Hyperactivity, cognitive deficits with chronic treatment Activity reversed by acute lithium and haloperidol Hyperdopaminergia
Amphetamine + CDP Hyperactivity Activity reversed with acute lithium Hyperdopaminergia, anxiolytic effects of CDP
Ouabain Hyperactivity, spatial learning deficits Activity reversed by haloperidol, lithium and valproate. Some neurodegenerative markers were reversed by lithium and valproate Increased levels of oxidative stress, changes in BDNF levels
D2 receptor stimulation Hyperactivity Activity reversed by valproate and carbamazepine Increased dopaminergic signaling
Sleep deprivation Hyperactivity, insomnia, aggressive behavior, hypersexuality, increased stereotypy, cognitive deficits, circadian rhythm disruption Sleep latency and hyperactivity reversed by lithium and haloperidol. PKC changes reversed by lithium Acute circadian rhythm and sleep changes can precipitate mania in humans. PKC signaling is altered.
Resident-intruder paradigm Increased aggression Aggression reduced by lithium and valproate Stress is associated with changes in mood state

Table 3.

Genetic mouse models of bipolar mania

Validity
Manipulation Face Predictive Construct
ClockΔ19 mutant Reduced anxiety, depression; increased impulsivity, reward-seeking; hyperactivity; impaired decision-making, sensorimotor gating; disrupted circadian rhythms (phase and amplitude) and sleep; episodic mania (mania and euthymia) Lithium and valproic acid normalized anxiety and depression behavior. Lithium normalized dopaminergic activity and defects in cross frequency phase coupling in the NAc. CLOCK polymorphisms; disrupted phase coherence, synchronization, and communication of cortico-striatal circuitry; hyperdopaminergia and altered glutamatergic neurotransmission
GSK-3β OX Hyperactivity; reduced depression Lithium reduced activity GSK-3β polymorphisms; reduced expression human bipolar DLPFC and temporal cortex; disrupted downstream targets of GSK-3β (e.g., β-catenin); hyperdopaminergia
DAT-KD Hyperactivity (reduced spatial d); increased goal-directed behavior; repetitive locomotor patterns; impaired decision-making Valproic acid reduced hyperactivity DAT polymorphisms; impaired DAT function; hyperdopaminergia
SHANK3-OX Hyperactivity; hypersensitivity to reward stimuli (e.g., amphetamine); elevated acoustic startle response; impaired sensorimotor gating; reduced depression; altered circadian rhythm behavior No effect of lithium; valproic acid reduced hyperactivity, amphetamine-induced locomotion, normalized sensorimotor deficits; reversed abnormal EEG patterns in frontal cortex and hippocampus SHANK3 polymorphisms; variants predict treatment response to ketamine in patients with bipolar depression; imbalance between excitatory and inhibitory neurotransmission; altered dopaminergic and glutamatergic signaling
ANK3 disruptions Hyperactivity; reduced anxiety; altered circadian activity rhythms; increased reward-seeking; stress-induced anxiety, anhedonia Lithium reduced hyperactivity, normalized anxiety Potential regulation of β-catenin and Wnt signaling pathways
GCLM-KO Hypersensitivity to amphetamine locomotor effects; altered social behavior; impaired sensorimotor gating Untested Reduced GSH levels in human postmortem PFC and other brain regions from bipolar patients; GSH alterations associated with GABA dysfunction, oxidative stress markers, and altered dopaminergic neurotransmission
Myshkin mutant Hyperactivity; hypersensitivity to amphetamine locomotor effects; impaired sensorimotor gating; disrupted sleep homeostasis; altered circadian rhythm behavior Lithium or valproic acid normalized anxiety and hyperactivity; exercise and melatonin improved sleep quality, reduced hyperactivity and normalized anxiety ATP1A3 polymorphisms, NA+K+ATPase a3 sodium pump dysfunction, multiple isoforms associated with bipolar disorder; altered ERK signaling and downstream effectors, such as BDNF, DISC1, GluR6, RASGRP1, and EGFR
Black Swiss Hyperactivity; hypersensitivity to amphetamine locomotor effects; reduced anxiety and depression; elevated sucrose preference Lithium and valproic acid reduced sucrose preference, amphetamine-induced locomotion, normalized depression behavior; asenapine (antipsychotic) normalized depression, reduced hyperactivity and amphetamine induced activity Genetic heterogeneity; reduced β-catenin expression in various brain regions
Madison Hyperactivity; reduced anxiety and depression; increased sexual behavior; disrupted circadian rhythm behaviors Lithium and olanzapine selectively reduced hyperactivity Genetic heterogeneity; abnormal expression patterns of numerous genes associated with bipolar disorder (e.g., Smarca4)
DBP-KO Abnormal circadian rhythm behaviors; disrupted sleep homeostasis; reduced activity; mania-like behaviors induced by chronic stress or acute sleep deprivation include hyperactivity, increased drug-taking Valproic acid blocked stress-induced hyperactivity Gene locus mapped near coding region for Dbp; ranked highly among candidate genes for bipolar disorder

Pharmacological models of mania

Amphetamine-induced hyperactivity

The initial animal models of human mania relied heavily on the induction of hyperactivity in response to drugs that modulate dopaminergic activity (Table 2). People with BD have higher urinary dopamine levels with the emergence of manic symptoms (Joyce et al., 1995). Furthermore, many studies have found that psychostimulants (such as amphetamine) can produce symptoms that resemble human mania in normal healthy subjects (Meyendorff et al., 1985, Peet and Peters, 1995, Cousins et al., 2009) as well as exacerbate symptoms or induce a manic episode in bipolar patients. Therefore, the effects of psychostimulants on behavior became widely used as an animal model of mania (Young et al., 2011a). Amphetamine-induced hyperactivity in rodents was thought to be the ‘gold-standard’ rodent model of mania for many years (Berggren et al., 1978, Gould et al., 2001). This is despite evidence showing that amphetamine-induced alterations of behavior in humans without BD are often quite different to those seen in rodents (Silverstone et al., 1998). Amphetamine-induced hyperactivity can be reversed with certain mood stabilizing medications. Recent studies found that microinjection of valproic acid (VPA) in the ventricle, amygdala, striatum, and prefrontal cortex blocked the hyperactivity induced by methamphetamine (Arent et al., 2011). Amphetamine-induced hyperactivity can also be reversed by acute lithium in certain mouse strains but not others. For example, Gould et al., found this effect in C57BL/6 mice but not in C3H mice (Gould et al., 2001). Moreover, this effect of lithium can be confounded by lithium-induced reductions in activity in untreated mice, particularly in habituated animals during the inactive phase of their circadian cycle (Berggren et al., 1978), Chronic treatment with lithium does not appear to reverse amphetamine-induced hyperactivity even though chronic treatments are often necessary in humans to obtain a therapeutic response (Fessler et al., 1982, Cappeliez and Moore, 1990). In randomized controlled trials, patients can still exhibit mania rating scores high enough to meet criteria for enrolling in the study even after 3 weeks of mood stabilizer treatment (Sachs and Gardner-Schuster, 2007). Another major limitation of this model is that human mania is characterized by a broad set of symptoms beyond simple hyperactivity. Thus, this model only captures one particular facet of this complex phenotype. Mania is also a chronic disease which may lead to significant neuroplasticity while this amphetamine treatment is normally acute (Machado-Vieira et al., 2014). However, locomotor sensitization to repeated doses of amphetamine has also been suggested as a model of human mania (Cappeliez and Moore, 1990). This model could be viewed as a more relevant model to mania than acute treatment with these drugs since the sensitization paradigm induces a greater locomotor response over time via neuroplasticity of the dopaminergic circuitry. Acute administration of lamotrigine, lithium, valproate, retigabine and carbamazepine all reverse amphetamine-induced sensitization (Dencker and Husum, 2010), although valproate alone also reduces activity in vehicle-treated mice. A recent study looked at cognitive behavior and levels of BDNF in several brain regions in response to either seven or thirty five days of amphetamine with the goal of modeling cognitive impairments often seen in bipolar subjects with progression of this disorder. They found that both treatments led to impaired habituation memory and inhibitory avoidance with greater levels of habituation memory impairment in the chronic group (Fries et al., 2015). Moreover, they found increased brain derived neurotrophic factor (BDNF) expression in the prefrontal cortex and amygdala and decreased BDNF protein in the hippocampus with these treatments (Fries et al., 2015). Thus, it is possible that very chronic treatment with amphetamine could be used to model the neuroprogression and cognitive impairment often seen in BD.

Amphetamine + CDP

Other animal models have used amphetamine in combination with other agents with the hopes of reproducing other symptoms of human mania beyond hyperactivity (Table 2). The benzodiazepine derivative chlordiazepoxide (CDP), which has anxiolytic and sedative effects alone, has been used in combination with amphetamine as a rodent model of mania (Poitou et al., 1975, Arban et al., 2005). However, the most common behavioral output measure used is still locomotor activity, since amphetamine and chlordiazepoxide (AMP + CDP) treatment increases activity in both rats and mice. Similar to amphetamine alone, AMP + CDP–induced hyperactivity is reversed by acute lithium, retigabine, lamotrigine and levetiracetam (Lamberty et al., 2001, Dencker et al., 2008, Redrobe and Nielsen, 2009). However, similar to amphetamine alone, the predictive validity of the AMP + CDP model is questionable, since both lamotrigine and levetiracetam have limited efficacy for treating mania (Goldberg and Burdick, 2002, Grunze et al., 2003, Kruger et al., 2008). One of the major concerns when using the AMP + CDP model of mania is the number of control groups required for each study (vehicle alone, amphetamine alone, CDP alone, AMP + CDP in combination, AMP + CDP plus each treatment, amphetamine plus each treatment and CDP plus each treatment) (Young et al., 2011a). Arban et al. (2005) conducted a complete control comparison study employing the AMP + CPD model in conjunction with several potential anti-manic agents (Arban et al., 2005). CPD alone often potentiated the locomotor activity-suppressant effects of known therapeutic treatments. Thus, it is unclear if the combined model provides any meaningful data on the action of anti-manic agents, or whether changes in activity simply represent a combination of additive drug effects. Furthermore, the complexity of controls required in the AMP + CPD model of mania makes it difficult to interpret the responses to novel putative anti-manic agents. For example, Kozikowski et al. (2007) examined the effects of a novel GSK-3β inhibitor in the AMP + CPD model of mania, and found that consistent with a high dose of valproate, the novel inhibitor attenuated but did not reverse AMP + CDP–induced hyperactivity (Kozikowski et al., 2007). However, the authors did not assess the effects of valproate treatment alone or in combination with either amphetamine or CPD, and the inhibitor effect was not compared to the vehicle treatment which appeared to reduce activity on its own. Therefore the complexity of factors make implementation and interpretation difficult in this model. Moreover, Kelly et al. (2009) found that the combination of amphetamine +CDP produced an inverted-U dose response in measures of hyperactivity in outbred CD1 and inbred C57BL/6N or 129S6 mice (Kelly et al., 2009). The important implications of this are in the interpretation of any effects seen with the addition of mood stabilizing drugs in that one cannot be certain as to whether the drug produces a “blockade” or a “potentiation” of the effects of amphetamine + CPD.

Oubain

The adenosine triphosphotase inhibitor, ouabain, administered ICV also induces hyperactivity in the open field and this treatment has been proposed as a model of mania (el-Mallakh et al., 1995) (Table 2). Moreover, this effect can last for seven days after a single ICV infusion (Riegel et al., 2009). OUA treatment is associated with severe neuronal damage and high levels of oxidative stress through the increased formation of lipid and protein oxidation products along with decreased BDNF levels in the brain (Riegel et al., 2009, Jornada et al., 2010). Some of these changes were modified by lithium and valproate (Jornada et al., 2011). For example, ouabain decreased levels of Bcl-2 and increased oxidative stress factors such as BAX and pp53 in rat brain (Valvassori et al., 2015). Treatment with both lithium and valproate improved most of these measures, however they were not protective against the increases in BAX in frontal cortex or hippocampus (Valvassori et al., 2015). In patients with BD, decreased BDNF levels have been described when compared with healthy controls (Machado-Vieira et al., 2007). The typical antipsychotic, haloperidol decreases ouabain-induced hyperactivity in rats, an effect which is confounded by haloperidol-induced reduction in activity alone (El-Mallakh et al., 2006). In contrast, the atypical antipsychotic olanzapine does not significantly reduce activity alone, nor does it attenuate ouabain-induced hyperactivity (El-Mallakh et al., 2006). Moreover, when one compares behaviour in a well lit open field and in a dark activity chamber (Decker et al., 2000); the results indicate that ouabain increases activity only in the open-field test. Thus, a novel and somewhat anxieogenic environment is needed to see the measurable effects of this compound. Furthermore, repeated exposure to an open-field chamber may induce a ‘floor effect’, where activity levels are so low that a drug effect alone could not lower them further. For example, Li et al. (1997) habituated rats for several days in the open field prior to quantifying ouabain-induced hyperactivity and acute lithium ‘reversal’ of effects. Lithium failed to decrease activity levels in this study since they were already minimal due to environmental familiarity (Li et al., 1997).

D2 receptor stimulation

Acute administration of the dopamine D2 receptor agonist, quinpirole can induce hyperactivity at certain doses and this treatment has also utilized as an animal model of mania. Chronic administration of the anti-manic drugs valproate and carbamazepine can reverse quinpirole-mediated locomotor activation (Shaldubina et al., 2002). The attenuation of quinpirole-induced hyperactivity was also found however with topiramate (Shaldubina et al., 2002), which is not a treatment for mania (Kushner et al., 2006, Vasudev et al., 2006) and may in fact induce a manic episode in certain patients (Jochum et al., 2002). It is relevant to note that topiramate has not been assessed in many other models of mania.

Environmental models of mania

Sleep deprivation and circadian rhythm disruption

Some labs have attempted to use environmental manipulations to induce manic-like symptoms in animals. Since it is well known that circadian rhythm disruption and sleep disturbances can often trigger manic episodes (Malkoff-Schwartz et al., 1998, McClung, 2007), researchers have used various sleep deprivation paradigms to model human mania (Gessa et al., 1995) (Table 2). These paradigms often involve placing a rat or mouse on a small platform (3–7 cm) surrounded by water for an extended period of time, typically 72 hours (Gessa et al., 1995). If the animal falls asleep it will fall into the water, therefore the animal stays awake to prevent this from happening. When rodents are placed back into their home cage after the sleep deprivation period, they exhibit a cluster of manic-like behaviours that include insomnia, hyperactivity, aggressive behavior, hypersexuality (an increase in mounting behaviour) and increased stereotypy (sniffing and rearing) (Morden et al., 1967, Hicks and Moore, 1979, Hicks et al., 1979, Moore et al., 1979). This abnormal behavior is at its peak 30–40 min period immediately after the sleep deprivation treatment but there can be lasting effects (Einat, 2006). A recent study looked at individual differences in the response to a 3 day circadian rhythm disruption protocol in mice and found that those mice that had a delay in the recovery of normal sleep/wake behavior displayed an increase in quinpirole-induced hyperactivity 10 days after the paradigm ended, suggesting that these particular mice showed a lasting “manic-like” response following a short circadian rhythm disruption (Jung et al., 2014). Gessa et al. (1995) found that 10 days of lithium treatment significantly reduced hyperactivity and decreased the latency to fall asleep compared with vehicle in 72 hour sleep-deprived rats (Gessa et al., 1995). In addition, rats given four injections of haloperidol in the last hours of the sleep deprivation period also demonstrated a decreased latency to fall asleep relative to vehicle-treated animals, which was interpreted as a reduction in manic-like insomnia. Thus in contrast with pharmacological models of mania, acute sleep deprivation induces a range of abnormal behaviors beyond simple hyperactivity and is responsive to chronic treatment with mood stabilizing drugs, making it perhaps more relevant for the study of human BD. However, while studies of sleep deprivation in healthy human controls can produce a temporary manic-like state, they rarely develop full-blown mania as seen with subjects with BD, and this state in healthy controls usually only lasts until the subject is able to sleep again.

While the sleep-deprivation paradigm described above may be useful in understanding the role of sleep in the development of manic-like behavior, it is important to note that this paradigm includes several stressors aside from simple sleep deprivation, including isolation, immobilization and the experience of falling into water. So it is not clear which of these stressors leads to the abnormal behaviors. A study by Benedetti et al. administered intervals of 24 hour sleep deprivation to CD1 mice while also including a stress control group with a larger stable platform enabling sleep, but enforcing isolation and immobilization (Benedetti et al., 2008); only the sleep-deprived mice exhibited increased locomotion and aggressive behaviour relative to controls, indicating that lack of sleep is key to the development of a manic-like phenotype.

How does sleep deprivation lead to a manic-like phenotype? One study examined the effect of lithium and sleep deprivation on brain protein kinase C (PKC) activity, an enzyme signaling pathway implicated in locomotor hyperactivity, risk-taking and excessive hedonic drive associated with mania (Szabo et al., 2009). Wistar-Kyoto rats subjected to a relatively short period of sleep deprivation (4 h during the light cycle) exhibited increased PKC phosphorylation of AMPA receptors, neurogranin and myristoylated alanine-rich C kinase substrate (MARCKS) in frontal cortex relative to non-sleep deprived animals. In contrast, C57BL/6 mice treated with lithium in food for 4 weeks showed significantly reduced phosphorylation of these PKC molecular targets compared with animals fed a normal diet. While behavioural symptoms such as hyperactivity or aggression were not assessed in this study, the data indicate that a sleep deprivation paradigm may have construct validity in representing the neurobiological alterations that may precipitate manic episodes.

Resident-intruder Paradigm

Human mania is often characterized by increased aggression, agitation and intrusive actions. These behaviors can be induced in rodents using variations of the resident–intruder test (Einat, 2006). This task typically starts with the introduction of a group-housed intruder rodent into the home cage of an isolated mouse or rat. The aggressive acts committed by the resident and the defensive acts and postures exhibited by the intruder are recorded (Miczek et al., 2001). Aggressive behaviours by the resident, including biting, threatening postures and tail rattling, can be increased by prolonged isolation or exposure to acute stressors such as foot shock (Legrand and Fielder, 1973, Miczek and O’Donnell, 1978). Typically the resident–intruder test has been employed as a model of depression by studying the social defeat characteristics displayed by the intruder. However, treatment with lithium can reduce resident aggression in both rats and mice (O’Donnell and Gould, 2007) suggesting that the aggressive nature of the resident may in some way model the aggressive behavior seen often in human mania (Table 1,2). Lithium administration has also consistently reduced aggression induced by mild electric foot shock in rats and mice, as well as attenuate combative behaviour augmented by stimulants such as D-amphetamine (Eichelman et al., 1973, Mukherjee and Pradhan, 1976, Prasad and Sheard, 1982, Kovacsics and Gould, 2010). However, lithium also decreases the response to shock in the jump-flinch test which could confound the interpretation of lithium’s effects on shock-mediated aggression paradigms (Kovacsics and Gould, 2010). Lithium also reduces isolation-induced fighting outside the home cage, as well as resident–intruder fighting within the home cage (Brain and Al-Maliki, 1979, Oehler et al., 1985). One study examined the effects of both lithium and valproate on aggression using a resident–intruder paradigm in C57BL/6 mice (Einat, 2007). Resident mice treated with lithium for 4 weeks before the test exhibited a significant reduction in the number of attacks against intruder mice compared with controls but did not show any difference in non-aggressive interactions such as rearing or sniffing. Finally, it is also worth mentioning that chronic treatment with a variety of antidepressant drugs increases aggression in the resident–intruder paradigm (Mitchell, 2005), suggesting that this may constitute a model of antidepressant-induced mania in individuals with BD (Koszewska and Rybakowski, 2009).

While the resident–intruder test has been in use for several decades (Miczek et al., 2001), a different model of dominant–submissive behaviour has also been proposed to induce symptoms of both depression and mania (Malatynska and Knapp, 2005). This paradigm is based on the notion that subordinate animals, similar to depressed humans, show increased defensive behaviour and reduced locomotor activity; in contrast, dominant animals, similar to humans in a manic state, are characterized by self-confident, assertive and aggressive behavior (Gardner, 1982). In this paradigm, two rats are placed in opposite chambers connected by a narrow tunnel allowing only one animal at a time access to a food source. When tested over 2 weeks in 5 min daily sessions, approximately half the animal pairs develop a dominant–submissive relationship, where one dominant animal monopolizes access to the food, defined as spending 40% more time at the feeder compared with the submissive rodent. When treated with antidepressant medications such as imipramine, desipramine and fluoxetine, submissive rats become more assertive, resulting in a significantly greater access to the food after 2 to 4 weeks of treatment. Conversely, dominant rats injected with lithium, carbamazepine or valproate become less aggressive and lose their dominant status (Malatynska and Knapp, 2005, Malatynska et al., 2007). In addition, the mood stabilizers exert an effect over a time course that is similar to that observed in bipolar patients in that lithium and carbamazepine treatment reduced rodent dominance only after 2–3 weeks of treatment. It is important to mention, however, that resident–intruder paradigms are also sensitive to the effects of anxiolytic and anxiogenic drugs, suggesting that behavioural changes in this model could be due to changes in anxiety, limiting the predictive validity of this model for BD (Vassout et al., 2000).

Genetic models of bipolar mania

Human genetic studies have identified many gene polymorphisms associated with psychiatric diseases, such as depression, anxiety, and BD, which implicate altered gene function in these disorders. Rodent models that leverage single gene mutation, knockout, or transgenic technologies are highly valuable for understanding the impact of a specific gene, or polymorphism, on behavior and the underlying mechanisms, although there are several limitations in their ability to fully resemble the pathogenesis in humans. For example, it is highly unlikely that BD is related to a single gene mutation. As stated earlier, psychiatric diseases emerge from complex etiologies of environmental and genetic origins (Bagot et al., 2014, Nestler, 2014, Pena et al., 2014). Nevertheless, there is a tremendous amount of scientific and translational utility in studying single gene mutation or knockout animal models for understanding basic mechanisms of brain function, behavior, and their relation to psychiatric diseases (Table 3). Several genetic animal models continue to be used for modeling specific endophenotypes that are common to many psychiatric diseases, or related to a particular illness. These models are also being used to identify the potential mechanisms underlying the therapeutic effects of existing treatments, or to screen the therapeutic potential of novel compounds.

BD is characterized by mood episodes accompanied by shifts in energy and activity. Over the past several years, there are several newly developed transgenic mice to model human mania. However, it has been incredibly difficult to develop models with spontaneous cycling between manic and depressive behavioral states (Young and Dulcis, 2015). We will discuss these new genetic models and recent attempts to investigate ‘mood’ switching. Some of these models are able to induce switching between manic and depressive related behaviors to different environmental stimuli or by probing the potential circadian variation of these behaviors, thus improving on their validity as a model of the human condition (Le-Niculescu et al., 2008, Sidor et al., 2015).

ClockΔ19 mutant mice

While the exact cause of BD is unknown, abnormalities in circadian rhythms may contribute to the pathogenesis of the disorder (Frank et al., 2000, Wirz-Justice, 2006, McClung, 2007), and markers of circadian disruption may predict susceptibility to mood changes, as well as treatment response (Kripke et al., 1978, Milhiet et al., 2011, McCarthy et al., 2013). A polymorphism in the 3′ flanking region of the Clock gene was associated with a greater frequency of manic episodes, insomnia, early morning waking, and reduced need for sleep in bipolar patients (Benedetti et al., 2003, Serretti et al., 2003). Efforts are underway in our laboratory to determine the consequences of these polymorphisms on CLOCK expression and function. The most well-characterized genetic mouse model of mania is the Clock mutant mouse (ClockΔ19) (King et al., 1997b, Vitaterna et al., 2006) (Table 3). CLOCK, along with its binding partner BMAL1, form heterodimers that drive the positive loop of the molecular clock. The molecular clock is comprised of interconnecting transcriptional-translational positive and negative feedback loops oscillating over a 24 hour period (Gustafson and Partch, 2015). ClockΔ19 mutant mice have a single base mutation that leads to a loss of exon 19 during an aberrant gene-splicing event, which effectively deletes the coding region for the transactivational domain that is required for CLOCK-mediated transcription (King et al., 1997a). Hence, circadian rhythms of gene transcription and behavior are significantly altered in Clock mutant mice (Antoch et al., 1997). It is worth noting that behavioral rhythms of Clock mutant mice almost fully breakdown under constant conditions, which is more severe than those measured in individuals with BD. However, Clock mutant mice share other circadian phenotypes with BD, including reduced circadian amplitude and delayed phase (Vitaterna et al., 1994, Challet et al., 2000, Vitaterna et al., 2006, Etain et al., 2014, Rock et al., 2014).

Recent efforts have begun to develop and employ translational approaches providing cross-species validation of behavioral phenotypes relevant to BD (Henry et al., 2010, van Enkhuizen et al., 2013b). Impaired sensorimotor gating and hyperactivity are core features of acutely manic patients, which can also be readily assessed in rodent models. Clock mutant mice also have sensorimotor deficits along with attenuated startle habituation during paired pulse inhibition (PPI) paradigms (van Enkhuizen et al., 2013b). More systematic studies are necessary to determine whether sensorimotor gating and startle responses vary based on sex and/or ‘mood’ state, as has been shown in BD patients (Barrett et al., 2005, Gogos et al., 2009). In addition to sensorimotor impairments, future studies should continue to expand on the performance of Clock mutant mice during ‘cognitive’ tasks, such as spatial working memory, attention, and decision-making, because cognitive impairments may be a hallmark of the disorder, at least in certain individuals. Indeed, neurocognitive deficits often accompany sensorimotor impairments in patients experiencing either manic or depressive episodes (Torres et al., 2011).

A change in activity is another core feature of BD, particularly during manic episodes, and can also be measured in humans within a laboratory setting. The behavioral pattern monitor (BPM) is essentially an open-field area designed for humans to explore, investigate and interact with different stimuli, such as familiar and novel objects, as a standardized paradigm to assess level, pattern, and organization of activity, in addition to goal-directed behavior (Young et al., 2007, Perry et al., 2009, Minassian et al., 2010, Young et al., 2010). The human BPM is an intriguing approach to investigating activity related phenotypes in BD and assessing whether rodent models display similar patterns (Geyer et al., 1986) van Enkhuizen et al., 2013c) (Table 1). Acutely manic patients are hyperactive, with more object exploration and goal-directed behavior (linear, direct movements) (Perry et al., 2010, Minassian et al., 2011). Clock mutant mice, on the other hand, display subtle dissimilarities with acutely manic patients. While mutant mice are hyperactive and have the ability to habituate to repeated testing environments similar to manic patients (Easton et al., 2003, McClung et al., 2005), their pattern and organization of locomotor activity are different (referred to as their ‘spatial d’, which ranges from 1 to 2, where 1 represents a straight path and 2, highly circumscribed small scale movements; (Perry et al., 2009). Acutely manic patients tended to have more linear movements, represented by a lower spatial d, whereas spatial d was higher, more circumscribed repetitive movements, in Clock mutant mice (Perry et al., 2009, van Enkhuizen et al., 2013c). These discrepancies require further study, such as elucidating the underlying neurobiological mechanisms regulating these activity phenotypes, which may be related to altered dopamine transmission in striatum and VTA. Nevertheless, Clock mutant mice display both hyperactivity and sensorimotor deficits similar to acutely manic patients.

Clinical symptoms used for diagnoses include reduced depression and increased risk-taking behavior during manic episodes. Much of our work has focused on discovering the cellular and molecular mechanisms that regulate anxiety, risk taking, and depression related behaviors in Clock mutant mice. Turek and colleagues were the first to report increased exploratory and escape seeking behaviors in both male and female Clock mutant mice (Easton et al., 2003). These findings have been replicated and expanded by our group and others and include a wide range of manic-like phenotypes (Table 3) (Roybal et al., 2007, van Enkhuizen et al., 2013b). We have found that these mania-like behaviors are driven by enhanced dopamine release from VTA neurons (e.g., enhanced dopamine cell bursting and firing and increased extracellular dopamine) (Coque et al., 2011, McClung et al., 2005, Sidor et al., 2015). Indeed, selectively driving dopamine neurons of the VTA with optogenetics recapitulates the hyperdopaminergic state and the anxiety and depression related behaviors of Clock mutants (Sidor et al., 2015).

These behaviors can be completely normalized to wild-type levels by chronic lithium treatment (Roybal et al., 2007). Interestingly, we found that cholecystokinin (Cck), a gene encoding for a peptide that is co-released with dopamine to inhibit further dopamine release by activation of presynaptic receptors, is a gene that is rhythmically transcribed by CLOCK:BMAL1 heterodimers, and significantly reduced in the VTA of Clock mutants (Arey et al., 2014). Lithium increased Cck expression in the VTA of Clock mutants through epigenetic modifications and recruitment of other transcription factors to the gene promoter (Arey et al., 2014). Interestingly, Cck expression was significantly increased in human postmortem VTA tissue collected from medicated patients with BD (Arey et al., 2014). Thus, Cck may be a primary target of lithium treatment that is relevant for its therapeutic effects on mania.

The mechanisms underlying CLOCK’s regulation of these behaviors is complicated, mostly due to our findings of a mixed behavioral state (greater novelty exploration, hyperactivity and increased depression-like behavior) following knockdown of Clock specifically within the VTA of wild-type mice (Mukherjee et al., 2010). Consistent with this, CLOCK overexpression (OX) in the VTA of Clock mutants only rescued hyperactivity and anxiety-related behaviors (Roybal et al., 2007). Other molecular mechanisms in alternative brain areas seem to be involved based on these findings and should be explore in the future. In fact, CLOCK in the hippocampus may be involved in depression related behaviors (Jiang et al., 2013).

Alterations in circadian rhythms may be a common factor in the vulnerability to highly comorbid psychiatric diseases, substance use and abuse disorders and BD (Sjoholm et al., 2010). It is well known that individuals with BD are more prone to developing substance use problems and addiction. In addition to the mania-like phenotypes, Clock mutant mice display ‘hyperhedonia’—an increase in reward sensitivity and motivation. At lower doses of cocaine, Clock mutant mice were more prone to developing conditioned place preference, a proxy of contextual reward conditioning (McClung et al., 2005). Additionally, these mice had increased locomotor sensitization to repeated cocaine injections and lowered reward thresholds during intracranial self-stimulation (ICSS) of the medial forebrain bundle, an intensely pleasurable and rewarding action (McClung et al., 2005, Carlezon and Chartoff, 2007). During operant self-administration paradigms, Clock mutant mice administered higher levels of cocaine across a range of doses and reached higher break-points than wild-type mice (i.e., enhanced seeking and motivation) (Ozburn et al., 2012). Diurnal patterns of cocaine self-administration found in wild-type mice were completely lost in Clock mutants (Ozburn et al., 2012), suggesting a ‘loss of control’ of normal circadian regulation of drug-taking behavior. Finally, Clock mutant mice consume more alcohol at higher concentrations during two-bottle free choice paradigms (Ozburn et al., 2013). Collectively, these data suggest a hyperdopaminergic state observed in the Clock mutants contributes to anxiety, depression, and addiction related phenotypes. Disrupted circadian rhythms contribute to the vulnerability and emergence of addiction and mood disorders (Logan et al., 2014). However, there are very few studies investigating the potential interplay between Clock gene variants, addiction, and mood disorders. One study reported specific variants of Clock may increase the vulnerability to depression in individuals with an alcohol use disorder (Sjoholm et al., 2010).

Communication across multiple brain areas is required for processing of information, decision-making, motor planning, and behavioral action. Disrupted connectivity or coherence between certain brain areas during discrete visual or auditory, or more demanding cognitive tasks has been repeatedly found in BD and may be related to mood state. Functional neuroimaging and electroencephalography (EEG) are techniques used to measure ‘connectivity’ and coherence. With EEG, coherence is defined as the coupling of or relationship between specific signals in a given frequency band, and thus represents the functional relationship or synchronization between cortical regions (Nunez et al., 1997, Sarnthein et al., 1998). EEG signals are filtered into delta (0.5–3.5 Hz), theta (4–7 Hz), alpha (9–13 Hz), beta (18–30 Hz), and gamma (28–48 Hz) oscillations. Disrupted coherence is thought to depict abnormal or poorly integrated brain function.

The generation of beta oscillations depend on the synchronization of gamma signals that originate from networks of inhibitory GABAergic interneurons (Gray and McCormick, 1996, Whittington et al., 2000). Long-distance gamma coherence was substantially reduced in both euthymic and manic patients (Ozerdem et al., 2010, Ozerdem et al., 2011), which may underlie elevated beta activity associated with mania (Ozerdem et al., 2008). Reduced gamma oscillations could reflect dysfunctional GABA signaling in cortical regions. Whether these markers are relevant for treatment response remains an open area of exploration (Ozerdem et al., 2010), these findings indicate altered beta and gamma activity are fairly robust representations of impaired functional connectivity of frontal and temporal cortical brain regions in BD (O’Donnell et al., 2004, Kam et al., 2013, Ozerdema et al., 2013). Neuroimaging studies have been able to more precisely identify the brains regions and define localities of the impaired connectivity. Dysfunction of subcortical, striatal-thalamic, and prefrontal networks, along with regions that send inputs to these areas, such as the amygdala, is associated with BD (Strakowski et al., 2005, Ongur et al., 2010). These networks may change over the course of the disease process. For example, abnormal activation patterns in the PFC, striatum, and amygdala preceded awry activation in the cerebellar vermis, lateral ventricles, and prefrontal regions that were more associated with repeated mood episodes (Strakowski et al., 2005).

Synchronization and coherence among areas of the cortico-striatal neurocircuitry is also disrupted in Clock mutant mice. Simultaneous extracellular recordings from nucleus accumbens (NAc), prelimbic cortex, and the VTA of freely behaving Clock mutant mice revealed severely impaired phase coupling and synchrony between limbic regions during the exploration of a novel environment (Dzirasa et al., 2010). Phasic entrainment, or coherence, of NAc low gamma (30–55 Hz) and delta (1–4 Hz) oscillations negatively correlated with exploratory behavior in wild-type mice, (i.e., more phase coherence, less exploratory drive). In Clock mutants, there were profound deficits in low gamma oscillations and single neuron phase coupling in the NAc, along with enhanced exploratory drive, which were reversed by chronic lithium treatment (Dzirasa et al., 2010). Reductions in gamma oscillations were also associated with reduced anxiety related behavior (Dzirasa et al., 2011). Overall, general circuit synchronization and coherence among the NAc and amygdala, or the VTA, predicted fear or anxiety related behavior, respectively, in wild-type mice, but failed to do so in Clock mutants (Dzirasa et al., 2011). Thus, low gamma coherence in the NAc and impaired phase synchrony within the limbic system may be contribute to the mania-like behaviors of the mutants. Although the precise mechanisms that drive these oscillations, or those that lead to impaired synchrony are unknown, we hypothesize an imbalance, or shift towards, dopaminergic and glutamatergic signaling within and from the limbic circuitry is primarily responsible. If so, this model may be useful for investigating the cellular and molecular mechanisms contributing to circuit level deficits associated with BD that likely involve the dysfunctional dopamine and glutamate crosstalk at the postsynaptic density believed to be a major contributor to the pathophysiology of the disease (de Bartolomeis et al., 2014).

GSK-3β overexpressing (OX) mice

At therapeutic concentrations, lithium inhibits several enzymes via competition with magnesium, including inositol monophosphatase, phosphoglucomutase, and glycogen synthase kinase-3 (GSK-3) (Can et al., 2014, Gould et al., 2004, Gould and Manji, 2005, Klein and Melton, 1996, Li et al., 2007, Phiel and Klien, 2001, Stambolic et al., 1996). Lithium inhibits GSK-3 in the mouse brain (De Sarno et al., 2002, Munoz-Montano et al., 1997, Noble et al., 2005, (O’Brien et al., 2004, O’Brien et al., 2011) and in peripheral blood cells of those with BD (Li et al., 2007, Polter et al., 2010). In several rodent models of mania, synthetic inhibitors of GSK-3β mimic the ‘therapeutic’ effects of lithium (Kozikowski et al., 2007, 2011). Furthermore, polymorphisms in GSK-3β are associated with certain symptoms of BD, as well as treatment response to lithium (Jope and Roh, 2006, Lin et al., 2013, Tang et al., 2013). The antipsychotic valproic acid also inhibits GSK-3β activity via mechanisms less characterized than lithium action (Chen et al., 1999, De Sarno et al., 2002, Jin et al., 2005, Sintoni et al., 2013).

GSK-3β has numerous cellular and molecular functions, including cell development, gene transcription, metabolic homeostasis, neurogenesis, and apoptosis (Doble and Woodgett, 2003). GSK-3β signaling pathways also form feedback loops with the molecular circadian clock to modulate circadian amplitude and period (Besing et al., 2014, Lavoie et al., 2013). The molecular clock regulates the activity of GSK-3β, and almost every core circadian protein, including BMAL1 and CLOCK, is directly phosphorylated by GSK-3β (Besing et al., 2014, Spengler et al., 2009). Interestingly, GSK-3β variants predicted the effects of lithium on molecular rhythms of skin fibroblasts cultured from bipolar patients (McCarthy et al., 2013). In general, lithium tended to resynchronize dampened rhythms of only those cells from bipolar patients (McCarthy et al., 2013). Thus, GSK-3β regulation of circadian rhythms may contribute to the mood-stabilizing properties of lithium (Kaladchibachi et al., 2007, Lamont et al., 2010, Li et al., 2012).

Perhaps the most investigated pathway by which GSK-3β might exert therapeutic action is via the Wnt signaling pathway that controls the expression and activity β-catenin, a nuclear transcription factor. GSK-3β inhibits the activation of this pathway by phosphorylating β-catenin for ubiquitin dependent degradation (Aberle et al., 1997, Orford et al., 1997). Protein expression of GSK-3β, pGSK-3β-ser-9, and β-catenin is significantly reduced in the dorsolateral prefrontal cortex and the temporal cortex of human post-mortem brains from bipolar patients (Pandey et al., 2015). A substantial effort needs to be put forth to investigate the transcriptional networks regulated by GSK-3β control of β-catenin-mediated gene transcription. Early efforts suggest there is tremendous overlap between genes previously identified as candidates involved in the pathophysiology of BD and those genes directly regulated by β-catenin-mediated transcription (Pedrosa et al., 2010), although this likely depends on brain region, treatment duration, and the course of the disorder. During chronic lithium treatment, GSK-3β is inactivated, preventing β-catenin degradation, and thus increasing protein expression (Yost et al., 1996, Behrens et al., 1998). A loss of GSK3β activity has consequences on synaptic plasticity and behavior via β-catenin signaling. Recent evidence suggests increased β-catenin activity destabilizes the formation of new dendritic spines and attenuates excitatory transmission (Ochs et al., 2015). β-catenin overexpression in the mouse brain leads to reduced depression related behaviors, and attenuated amphetamine induced hyperlocomotion and sensitization, quite similar to the effects of lithium treatment (Gould et al., 2007). GSK-3β haplosufficient mutant mice, lacking one copy of the GSK-3β gene, display an antidepressant-like behavioral phenotype, which also mimics the effects of lithium treatment mice (O’Brien et al., 2004). Together these results suggest the therapeutic action of lithium depends on inhibiting GSK-3β activity (O’Brien et al., 2011), possibly through the regulation of dopaminergic activity and signaling pathways downstream of dopamine receptor activation (Beaulieu et al., 2011, Beaulieu, 2012, de Bartolomeis et al., 2014, Gomez-Sintes et al., 2014).

GSK-3β-OX mice may have some relevance for modeling certain behaviors related to BD (Table 3). These mice are hyperactive and have reduced food intake across the circadian cycle (Prickaerts et al., 2006). Although, it is worth noting that both hyper- and hypo-phagia are associated with BD (Kishi and Elmquist, 2005, Krishnan, 2005). GSK-3β-OX mice are also hyperactive with impaired habituation to a novel environment and reduced depression related behavior (Prickaerts et al., 2006). These mice also have an increased startle response, which is in contrast to reports of low startle response, in acutely manic (Perry et al., 2001), mixed episode (Carroll et al., 2007), and remitted bipolar patients (Giakoumaki et al., 2010).

Other discrepancies between GSK-3β-OX mice and BD patients are worth noting—the corticosterone response to acute stress is normal in these mice (Prickaerts et al., 2006), while a blunted hormonal response is observed in BD patients (Houtepen et al., 2015), representative of altered HPA axis and autonomic nervous system regulation in response stress. In line with this, knockin mice carrying a transgene that impairs serine phosphorylation of GSK-3β (P-GSK-3β-KI), which effectively prevents GSK-3β inactivation, are more susceptible to stress-induced depression related behaviors (Polter et al., 2010). The relationship between stress and BD is complex and likely depends on a number of factors, such as chronicity of the disorder, treatment regimen, mood state, and uncontrollable environmental factors. The role of GSK-3β in stress regulation of neurocircuitry related to behavior relevant to psychiatric disorders requires further investigation (e.g., brain region or cell type specificity, compensatory signaling or plasticity). Indeed, mouse models of increased GSK-3β activity appear to have different behavioral phenotypes. Unlike GSK-3β-OX mice, P-GSK-3β-KI mice were hyperactive and more sensitive to amphetamine-induced hyper-locomotion (i.e., mania-like), yet exhibited increased anxiety and depression related behaviors (i.e., depressive-like) (Polter et al., 2010). As previously mentioned, this discrepancy may be due to enhanced sensitivity of the P-GSK-3β-KI mice to acute stress experienced during FST, TST, and learned helplessness paradigms (Polter et al., 2010). Acute stress decreased P-GSK-3β in the brains of wild-type mice, which indicates stress-induced GSK-3β activity (Polter et al., 2010). These effects suggest other molecular pathways are involved, but do lead to the hypothesis that the level of GSK-3β activity may be correlated with mood state. In peripheral blood mononuclear cells (PBMCs) from BD patients, P-GSK-3β levels negatively correlated with mania symptom severity, but not depression (Polter et al., 2010). The inhibitory effects of lithium on GSK-3β activity depend on serine phosphorylation, and thus offer a putative mechanism for mood stabilization, especially when considering lithium’s effectiveness for treating bipolar mania.

Dopamine transporter knockdown (DAT-KD) mice

Dysfunction in DAT has been proposed as a putative mechanism for altered dopaminergic neurotransmission in BD. Based on genetic, molecular, and imaging evidence indicating reduced DAT expression in BD (Horschitz et al., 2005, Greenwood et al., 2006, Anand et al., 2011, Pinsonneault et al., 2011), Young, Geyer, and colleagues have demonstrated mice with reduced DAT expression (~10% of wild-type DAT levels; Zhuang et al., 2001), exhibit a wide range of mania-like behaviors (Young et al., 2010, Young et al., 2011b, van Enkhuizen et al., 2013a, van Enkhuizen et al., 2014a, van Enkhuizen et al., 2014b) (Table 3). As mentioned previously, acutely manic bipolar patients display reduced spatial d (more linear movements) and increased goal directed behavior (object exploration) than healthy volunteers and schizophrenic patients in the hBPM (Perry et al., 2009). DAT-KD mice, however, display increased spatial d relative to other animal models of reduced dopamine uptake, including pharmacological inhibition of DAT by GBR12909 and acute amphetamine administration (Perry et al., 2009, Young et al., 2010). Mice treated with GBR-12909 were strikingly similar to bipolar patients in the BPM—hyperactive, increased exploration, with more predictable and repetitive locomotor patterns (Perry et al., 2009). When compared to their wild-type counterparts, DAT-KD mice displayed modest and sometimes inconsistently reduced spatial d, which is sometimes reflected by increased perseverative behavior (Ralph-Williams et al., 2003), Young et al., 2007, 2010, van Enkhuizen et al., 2013). While DAT-KD mice display these activity phenotypes, these mice fail to show sensorimotor deficits during PPI tests (Ralph-Williams et al., 2003), in contrast to other rodent models of mania, including the Clock mutant mice. Intriguingly, DAT knock-out mice (KO) which lack the gene encoding DAT, display significant sensorimotor deficits that could be reversed by clozapine and quetiapine treatment, two antipsychotics used for schizophrenia or BD (Powell et al., 2008). Chronic administration of valproic acid attenuated hyperactivity, while goal directed behavior remained unaffected and spatial d was further reduced (van Enkhuizen et al., 2013a). Acute treatment with the dopamine depleting agent, alpha-methyl-p-tyrosine (AMPT), attenuated hyperactivity, increased goal directed behavior, and failed to have any effect on spatial d (van Enkhuizen et al., 2014a). In patients with BD, the effects of AMPT on mood symptoms are mixed (Bunney et al., 1977, Anand et al., 1999). Thus, valproic acid selectively reverses some phenotypes related to locomotor activity potentially through dopamine independent mechanisms. The BPM has tremendous utility for investigating certain endophenotypes of BD and effects of common or novel treatments, and has the potential for integrating analyses across human patients and animal models of BD. The BPM represents a tractable translational approach for studying new and exciting endophenotypes important for BD.

Other features of bipolar mania include increased impulsivity and risk-taking behaviors, which may impair decision-making and resemble underlying cognitive deficits. One strategy for assessing impulsivity and its relation to decision-making in patients with BD has been the structured Iowa Gambling Task (IGT) (Bechara et al., 1994, Clark et al., 2001, Kim et al., 2006) (Table 1). In the IGT, participants have to select the best of four options varying both the size and probability of the reward and loss (Bechara et al., 1997, 2005). The IGT has proved reliable for investigating the relationships between impulsivity and risk-taking, decision-making, response inhibition, and drug use and dependence in BD (Christodoulou et al., 2006, Adida et al., 2008, Yechiam et al., 2008, Nejtek et al., 2013).

Researchers have developed a version of the IGT to study decision-making in rodents (Homberg et al., 2008, Rivalan et al., 2009, de Visser et al., 2011, Young et al., 2011b). An animal is presented with four distinct options with different probabilities and magnitudes of expected gains and losses—two are advantageous and the other two are distinctly disadvantageous. Impulsivity is defined as the frequency of premature responses (before options are presented) and risk-taking behavior is measured as the percentage of advantageous choices (advantageous-safe versus disadvantageous-risky choices). The impulsive choice or risk-taking behavior has the most translational validity to the human task (Evenden, 1999, van Enkhuizen et al., 2013a).

The rodent IGT has been used to investigate the role of serotonergic and dopaminergic mechanisms in impulsivity, risk-taking, and decision-making. Various DAT inhibitors have differential effects on IGT task performance in mice. Amphetamine treatment potentiated the choice of the advantageous-safe options, whereas more selective DAT inhibitors, modafinil or GBR12909, increased impulsivity in ‘normal’ C57BL/6 mice (van Enkhuizen et al., 2012, 2014). Similar to the effects of selective dopamine inhibitors, DAT KD mice are more impulsive and more motivated during the IGT, which impairs decision-making during risk-reward tasks (van Enkhuizen et al., 2014). Similar to DAT-KD mice, bipolar patients have impaired decision-making during the IGT, often exhibiting impulsive choices (van Enkhuizen et al., 2014). The IGT is another example besides the BPM of a cross-species, translational, and entirely tractable approach to investigating relevant disease constructs of BD. These approaches are necessary for understanding the neurobiological mechanisms driving disease-related functional impairments. Initial studies strongly suggest that hyperdopaminergic states in the brain, potentially due to reduced DAT efficiency, impairs decision-making processes by promoting the ‘system’s that drive impulsivity and motivation, or reward-seeking behavior. A neuroimaging study has shown reduced activation of ventral and dorsal PFC and heightened activation of the lateral temporal regions during the gambling task (Frangou et al., 2008). The ventral and dorsal PFC are involved in working memory and incentive value of stimuli, and communication between these two areas modulates reward—ventral PFC processes salience and incentive value, followed by input to the dorsal PFC, which influences the behavioral action. In light of these findings, ventral-dorsal PFC communication during decision-making tasks may be disrupted, requiring compensatory activation from other brain regions. Dysfunctional dopamine transmission in these patients could be responsible for these deficits (St Onge et al., 2012).

SHANK3-OX mice

Several human studies have suggested the SHANK3 protein is involved in the etiology of BD. Altered expression or function of this scaffolding protein, and others of the same family, are linked to several human synaptopathies, including schizophrenia and BD (Guilmatre et al., 2014). Case studies have provided further support for a link between SHANK3 dysfunction and BD (Vucurovic et al., 2012). The SHANK family of proteins is crucial for synapse formation, maintenance, and modulation of excitatory and inhibitory balance (Sudhof and Malenka, 2008). These proteins are primarily localized to excitatory synapses, which form large protein complexes at the postsynaptic density to connect with the actin cytoskeleton of dendritic spines (Han et al., 2013, Guilmatre et al., 2014). Mice that overexpress SHANK3 were hyperactive in both homecage and novel environments, and hypersensitive to the locomotor stimulating effects of amphetamine (Han et al., 2013). SHANK3-OX mice also had elevated acoustic startle response, sensorimotor deficits, reduced depression related behaviors in the TST, and altered circadian rhythms of locomotor activity (Han et al., 2013), much like some of other animal models of mania (Table 3).

A few of these behaviors of the SHANK3-OX mice were reversed by valproic acid treatment, including hyperactivity, amphetamine-induced locomotor activity, sensorimotor deficits, and abnormal EEG in the frontal cortex and hippocampus (Han et al., 2013). Surprisingly, lithium treatment had no effect on any of these phenotypes (Han et al., 2013), which could be consistent with a subset of those with BD who fail to respond to lithium treatment. Further investigation into the neurophysiological correlates of the behavior indicated these mice have an imbalance towards enhanced excitatory signaling within the hippocampus. SHANK3-OX mice had increased VGLUT1 and decreased VGAT postsynaptic markers, along with reduced GABA-A initiated miniature inhibitory postsynaptic frequencies, and enhanced amplitudes of spontaneous excitatory postsynaptic currents (EPSCs), without changes in AMPA/NMDA ratios (Han et al., 2013). This shift from synaptic excitatory-inhibitory balance towards excitation was directly due to enhanced interactions between SHANK3 and ARP2/3 complexes, which promoted the formation of excitatory dendritic spines along with a reduction in the number of inhibitory synapses (Han et al., 2013). Impaired GABAergic signaling may be a robust neurobiological endophenotype of schizophrenia and BD (Benes, 2007). A shift toward excitatory signaling in the hippocampus or striatum may further alter dopaminergic and glutamatergic signaling at the postsynaptic density, or may even be a consequence of a dysfunctional dopamine system (de Bartolomeis et al., 2014). Other mouse models with altered excitatory signaling in the brain display similar behavioral phenotypes as the SHANK3-OX mice, such as the glutamate receptor subtype 6 knockout mouse (Shaltiel et al., 2008). Importantly, a recent study reported certain SHANK3 variants as potential biomarkers for predicting treatment response with ketamine in patients with bipolar depression (Ortiz et al., 2014).

ANK3 disruptions

The ANK3 gene encodes isoforms of the scaffold protein, Ankyrin G, which is involved in the formation and maintenance of the axon initial segment of neurons. Ankyrin G links subsets of sodium and potassium channels to the cytoskeleton and is necessary for action potential propagation (Zhou et al., 1998, Bennett and Lambert, 1999). A meta-analysis of several large genome wide association studies of BD identified ANK3 as among the most significantly associated genes, though the results were not totally consistent across samples, and it may be a shared risk factor between schizophrenia and BD (Ripke et al., 2011, Lee et al., 2012, Lee et al., 2013). ANK3 levels are also downregulated in postmortem brain of schizophrenics (Roussos et al., 2012). Leussis et al.(2013) used RNA interference to knock-down the expression of Ank3 specifically in the dentate gyrus of the hippocampus. They found that this manipulation produced a reduction in anxiety-related behavior in the elevated plus maze and light/dark test (Leussis et al., 2013). It also produced increased locomotor activity specifically during the light phase, suggesting a disruption in the circadian rhythm amplitude of these mice (Leussis et al., 2013). These behavioral changes were reversed with lithium treatment. The authors also tested a heterozygote mouse in which Ank3 exon1b is disrupted, resulting in a loss of ANk3 transcript variants that are exclusively expressed in the brain (Leussis et al., 2013). These mice had a similar behavioral profile to those with the DG specific knock-down. In addition, these mice also had an increase in sucrose preference compared to wild type littermates. Importantly, in contrast to some of the other models discussed, both the knock-down animals and the heterozygotes show no behavioral abnormalities at baseline when measured in open field, forced swim test, acoustic startle, PPI, contextual and cued fear conditioning or overall motor activity in the EPM or dark-light box tasks (Leussis et al., 2013) (Table 3). The authors then measured behavior in the heterozygote animals following 6 weeks of social isolation stress and found increased latencies to enter the open arm of the EPM and light side of the dark-light box compared to wild type animals, suggestive of increased anxiety. However, given this interpretation, wild type animals tended to have less anxiety following social isolation in this study, which is not typical. Nevertheless, social isolation stress does appear to produce a reversal of phenotypes in the heterozygote mice, suggesting a ‘mood’ switch. Social isolation also decreased their sucrose preference to that of wild type animals and increased immobility of the Ank3 heterozygotes in the FST (Leussis et al., 2013). Thus, these animals seem to be very responsive to stress and indeed the mutants had greater levels of corticosterone following acute stress (Leussis et al., 2013). Taken together, ANK3 appears to be an important mediator of mood-related behavior and this model has great construct validity since the choice of gene for manipulation came directly from a human GWAS meta analysis. The fact that knock-down of Ank3 specifically in the DG leads to several manic-like behaviors is intriguing given the fact that most other models of mania have focused primarily on the manipulation of dopaminergic activity. It will be interesting in future studies to determine how this DG specific manipulation changes overall hippocampal function and if it leads to any changes in dopaminergic or glutamatergic signaling in the ventral striatum. A recent study found that ankyrin-G appears to regulate β-catenin anchoring in the progenitor cell membrane through an interaction with E-cadherin (Durak et al., 2015). This controls the level of β-catenin available for Wnt signaling and progenitor cell proliferation during development (Durak et al., 2015). Given the effects of lithium on Wnt signaling described earlier in this review, it is interesting to speculate that the behavioral effects caused by Ank3 knock-down in the DG may be mediated through changes in Wnt signaling and may potentially involve changes in adult neurogenesis.

Redox signaling mutants

Dysfunctional redox signaling and elevated oxidative stress have been reported in brains of patients with schizophrenia or BD. One of the primary intracellular regulators of redox and antioxidant signaling is glutathione (GSH). GSH protects against oxidative damage caused by reactive oxygen species (ROS) and reactive nitrogen species (RNS). Maintenance of cellular redox state is important for modulating receptor activation, signal transduction, and transcription factor binding to DNA. In human postmortem PFC, GSH levels were reduced in subjects with BD compared to controls (Gawryluk et al., 2011a, Gawryluk et al., 2011b). GSH deficits were associated with GABA dysfunction in parvalbumin (PV) neurons within the ventral hippocampus and the anterior cingulate cortex of bipolar subjects (Chen et al., 2010). Evidence from both GSH-deficient mice and rats suggests these PV neurons are particularly sensitive to oxidative stress, while other GABA interneurons are unaffected (Cabungcal et al., 2006, Steullet et al., 2010). Clinical trials have reported that adjunct treatment with the GSH precursor antioxidant N-acetylcysteine improves the remission of depressive symptoms and general functioning in bipolar patients (Dean et al., 2011, Berk et al., 2012). Thus impaired redox signaling and increased oxidative stress could contribute to dysfunction of PV neurons that is associated with schizophrenia and BD and may be secondary to poor neuronal metabolism and mitochondrial dysfunction.

Glutamate-cysteine ligase modifier unit (GCLM) KO mice have ~70–80% reductions in GSH levels in the brain relative to wild-type mice (Yang et al., 2002, Cole et al., 2011, Kulak et al., 2012). GCLM-KO mice display a range of mania-like behaviors, including enhanced sensitivity to the locomotor stimulating effects of amphetamine, altered social behavior, and impaired sensorimotor gating (Kulak et al., 2012) (Table 3). In response to mild stress, GCLM-KO mice become more exploratory, risky, and less ‘anxious’ and ‘depressed’ (Kulak et al., 2012). Transient GSH depletion in the adult rodent brain impaired performance during cognitive tasks, including object memory, place discrimination, and spatial working memory (Jacobsen et al., 2005, Mamiya et al., 2008). Thus chronic or transient deficits of GSH lead to different behavioral phenotypes, where transient depletion of GSH may have more selective impairments on cognitive performance. Low GSH may increase oxidative burden in the brain and impair mitochondrial function, both of which are posited to be involved in the pathophysiology of bipolar disorder and schizophrenia (Dean et al., 2009).

Changes in excitatory signaling and dopamine transmission may underlie the mania-like behaviors of GCLM-KO mice. The exaggerated locomotor response to amphetamine suggests hyperdopaminergia in these mice (Kulak et al., 2013). Fluctuations in neuronal redox state have functional consequences on excitatory and inhibitory signaling via NMDA and GABAA receptors, and L-type voltage gated calcium channels. For example, NMDA receptor mediated excitatory postsynaptic potentials were weaker in hippocampal slices from GSH depleted rats, while basal AMPA receptor mediated transmission was unaltered (Steullet et al., 2006). Transient reductions in GSH elevated dopamine metabolites in the mouse PFC, NAc, and hippocampus and altered dopamine modulation of neuronal calcium signaling (Jacobsen et al., 2005). In neurons, redox state and GSH also epigenetically regulate a host of genes directly involved in dopamine and glutamate signaling pathways (Trivedi and Deth, 2014). Dopamine fluctuations alter intracellular levels of GSH-based cellular redox status, subsequently affecting DNA methylation (Trivedi and Deth, 2014). Thus, redox and dopamine signaling pathways are intimately connected to regulate gene transcription and neuronal stimulation. Together, these studies support the hypothesis that impaired redox signaling, increased oxidative stress, and metabolic dysfunction are involved in the pathophysiology of BD. Currently, it is unknown whether these deficits precede or are a consequence, or function, of the chronicity and progression of the disease and secondary to abnormalities within other, more primary systems. GSH deficits selectively within PV neurons of bipolar subjects warrants further study of cell-type specific mechanisms regulating redox status in rodent models of mania. Targeted treatments to restore redox balance may prove to have therapeutic utility.

Myshkin mutant mice

Originally identified through ENU mutagenesis and developed as a preclinical model of epilepsy, the Myshkin mutant mouse has also been proposed as a model of bipolar mania (Snead et al., 2006) (Table 3). The Myshkin allele contains a mutation in the Na+, K+-ATPase α3-isoform that inactivates its enzymatic properties and consequently promotes neuronal excitability and seizure-susceptibility (Grisar et al., 1992, Clapcote et al., 2009). The ATP1A3 gene encodes for Na+, K+-ATPase α3 sodium pump and has been linked to BD (Goldstein et al., 2009) and human post-mortem studies have reported significantly reduced expression of the Na+, K+-ATPase α2-isoform in the temporal cortex (Rose et al., 1998) and the α2-isoform in the prefrontal cortex (Tochigi et al., 2008). Each of these α-isoforms (α1-3) has been associated with BD (Mynett-Johnson et al., 1998, Goldstein et al., 2009).

More recently, the Myshkin allele has been backcrossed to the seizure-resistant C57BL/6NCr strain, effectively producing mutants without seizures. Myshkin mutant mice (Myk/+) are hyperactive and display exaggerated locomotor response to repeated amphetamine, increased exploration of novel objects, reduced anxiety related behavior, impaired sensorimotor gating, and disrupted sleep patterns with altered circadian behavioral rhythms (Kirshenbaum et al., 2012). Chronic treatment with lithium or valproic acid normalized only the anxiety and hyperactivity phenotypes (Kirshenbaum et al., 2012). Interestingly, adjunct therapies for BD and other mood disorders, including exercise and melatonin, are effective for reversing mania-like behaviors in Myk/+ mice (Ng et al., 2007, Kucyi et al., 2010, Kirshenbaum et al., 2012, Sylvia et al., 2012, Kirshenbaum et al., 2014). More than a month of running wheel access or chronic treatment with melatonin supplementation improved sleep quality and reduced locomotor activity and normalized anxiety related behaviors in these mice (Kirshenbaum et al., 2014), suggesting these mice as a relevant model for studying the potential mechanisms underlying the therapeutic effects of these treatments. These treatments have not been systematically investigated in other genetic models of mania.

The mechanisms underlying these behaviors are due to impaired α3 sodium pump function. Transgenic restoration of the functional Na+, K+-ATPase α3-isoform completely normalized the anxiety related behaviors in the EPM, open-field, and dark-light box tests (Kirshenbaum et al., 2012). Activation of Na+, K+-ATPase α3 sodium pumps initiates intracellular calcium signaling and the phospho-activation of extracellular signal regulated kinase (P-ERK) and Akt (Dietz et al., 2007, Kim et al., 2008). Not surprisingly, Myk/+ mice have increased P-ERK and acute treatment with an ERK inhibitor (SL327) normalized anxiety behavior (Kirshenbaum et al., 2012). The role of ERK signaling is of particular interest for designing novel treatments because many mood stabilizers activate both ERK1 and ERK2 pathways, and many of the upstream regulators of ERK, such as BDNF, DISC1, GluR6, RASGRP1, and EGFR, are believed to be involved in the pathophysiology of BD (Dick et al., 2003, Chubb et al., 2008, Engel et al., 2009, Shinozaki and Potash, 2014). Moreover, ERK1 KO mice display a similar behavioral repertoire as Myk/+ mice (Engel et al., 2009). Other mice with impaired Na+, K+-ATPase signaling are susceptible to depression behavior following chronic stress (Kirshenbaum et al., 2011). It may be worthwhile to investigate whether Myk/+ mice are also vulnerable to environmental perturbations, such as chronic stress or acute sleep deprivation, to induce depression phenotypes, which could lead to more mechanistic studies underlying the susceptibility to triggers that induce mood cycling.

Mouse models amenable to systems neuroscience and functional genomics approaches

The Black Swiss mice

The major disadvantages of using single gene knockout, knockdown, or overexpressing, transgenic mouse models to study bipolar mania or depression are sometimes poor construct validity and poor translational interpretations of the findings. Most, if not all, of the genetic association studies, which have identified polymorphisms associated with BD, have small effect sizes, explaining very little of the genetic variance. Highly penetrant, large effect alleles associated with psychiatric disorders are rare. Psychiatric disorders are genetically heterogeneous, complex, and emerge from multiple trajectories, influenced by many, often nuanced, environmental and biological factors.

Inbred mouse strains have been used over decades of behavior genetics research to discover genetic loci associated with complex behavioral and biological phenotypes, albeit with very little success in elucidating the functional consequences of those polymorphisms. Inbred strains have also been investigated for their differential response to lithium following amphetamine induced hyperlocomotion (Gould et al., 2007). Using a mouse model with more genetic heterogeneity, may have distinct advantages over single gene mutant mice for modeling BD, including improved construct validity (i.e., polygenic mechanisms of complex behaviors).

Several groups have identified outbred mouse populations that model multiple domains of bipolar mania (Table 3). The Black Swiss mouse population was developed from crossing Swiss outbred and C57BL6/JN mice. When compared to the founder strains, Black Swiss mice are hyperactive and display elevated sucrose preference, reduced anxiety and depression behavior, enhanced sensitivity to amphetamine, and increased aggression (Flaisher-Grinberg and Einat, 2010, Ene et al., 2015). Therefore these mice may provide certain advantages for investigating the complex etiology of the disorder by modeling multiple behavioral domains and presumably more heterogeneous genetics (Flaisher-Grinberg et al., 2010).

Treatment with either lithium or valproic acid normalized certain mania-like behaviors of Black Swiss mice. Both lithium and valproic acid reduced sucrose preference, increased immobility time of the FST, and reduced amphetamine-induced hyperlocomotion, with no measurable effect on hyperactivity or anxiety behavior (Flaisher-Grinberg and Einat, 2010, Kalinichev and Dawson, 2011). Notably, GSK-3β inhibitors selectively attenuated amphetamine-induced hyperactivity in Black Swiss mice (Kalinichev and Dawson, 2011), consistent with the proposed GSK-3β-mediated mechanism of action for mood stabilizers. The atypical antipsychotic asenapine, which is indicated for schizophrenia and bipolar mania or mixed mood states (Marston et al., 2009, Gonzalez et al., 2011, Azorin et al., 2013), was more effective in normalizing multiple mania-like phenotypes of Black Swiss mice (Ene et al., 2015). In common with other antipsychotics, asenapine has a high affinity for dopamine receptor subtype 2 (D2R), which most likely mediates the anti-manic mechanism (Shahid et al., 2009). Differential responses to lithium and between antipsychotics of Black Swiss mice can be explained by inherent genetic factors (Gould et al., 2007).

To date, there has been only a single study investigating the construct (or biological) validity of Black Swiss mice. As mentioned previously, β-catenin is proposed as a downstream target of GSK-3β inhibition by lithium and Black Swiss mice have reduced β-catenin protein expression in the hippocampus, but not in the PFC, relative to C57BL6/J (Hannah-Poquette et al., 2011). However, BDNF is expressed normally in the hippocampus and frontal cortex of Black Swiss mice (Hannah-Poquette et al., 2011). In human postmortem tissue, BDNF transcripts (I, IIc, and IV) were reduced in hippocampus, but not the PFC or striatum, in subjects with BD (Reinhart et al., 2015). Circulating blood levels of BDNF in bipolar subjects correlated with mood state (Hashimoto, 2014) and treatment response (Dwivedi and Zhang, 2014, Tunca et al., 2014, Polyakova et al., 2015). Collectively, support the use of Black Swiss mice for modeling bipolar mania with high face and predictive validity, and possible construct validity.

The Madison mice

Another mouse model of mania derived from outbred stocks is the Madision (MSN) strain (Table 3). MSN mice display reduced anxiety and depression, increased sexual behavior, hyperactivity, and circadian abnormalities (Scotti et al., 2011, Saul et al., 2012, Saul et al., 2013). Lithium and olanzapine, an atypical antipsychotic, only reduced hyperactivity, with no changes observed for anxiety behavior in these mice (Scotti et al., 2011). In addition to face and predictive validity, MSN mice may also have high construct validity. MSN mice share genetic correlates with BD (Saul et al., 2012). A number of transcripts show abnormal expression levels in hippocampus of MSN mice, many of which are previously associated with BD (Saul et al., 2012). For example, genes encoding proteins involved in chromatin packaging and remodeling are altered in MSN mice, such as Smarca4, which has been implicated in neurodevelopmental disorders, including autism schizophrenia, and BD (Li et al., 2013). MSN mice also share synteny with several loci identified in human BD (Saul et al., 2012). Very little is known as to whether these loci are causative or structural variants with any functional consequences—MSN mice provide a model for investigating these mechanisms. Further studies are necessary to expand on the validity of the model, as well as to understand the polygenic contributions to mania-like behaviors (Saul et al., 2012). MSN mice have potential for being a tractable and highly valid model of bipolar mania to leverage the power of genomics and systems neuroscience approaches to begin to understand some of the mechanisms underlying the biological complexity of phenotypes related to mania.

Mood cycling mouse models: A focus on the role of circadian mechanisms of behavioral state switching

Over the past several decades, there have been attempts to model mood cycling in rodents. Many have used pharmacological or environmental triggers to induce behavioral state switching (see (Young and Dulcis, 2015)). More recently, researchers have employed circadian and/or sleep perturbations to affect behavior in rodent models with the intention to achieve higher construct, or etiological validity. Sleep and circadian disruptions are hallmarks of those suffering with BD and used as diagnostic criteria (Wirz-Justice, 2006, Kripke et al., 2009, McCarthy and Welsh, 2012, McClung, 2013, Gonzalez et al., 2014). Social rhythm therapy (SRT), which aims to stabilize circadian and sleep domains via daily scheduling, is efficacious for treating BD (Frank et al., 2008, Goldstein et al., 2014). The social rhythm disruption (SRD) theory of affective disorders hypothesizes rhythm disruption, whether social or biological, induces mood changes in vulnerable individuals (Ehlers et al., 1988, Hlastala, 2003, Boland et al., 2012). Stress can be a precipitating factor for rhythm disruption, sleep disturbances, and mood changes (Ellicott et al., 1990, Post, 1992). Rodent models have really just begun to investigate the mechanisms underlying the effects of circadian disruption and sleep disturbances on behavioral switching related to BD.

It has been evident for some time that seasonality in related to changes in mood in those with BD and other mood disorders (Geoffroy et al., 2013, Wang and Chen, 2013, Moore et al., 2014). Seasonal changes in mood may be driven by changes in circadian photoperiod. The most consistent environmental trigger of depressive episodes in BD is the shortening of day length (Wang and Chen, 2013). In diurnal rodents, shortened photoperiods induced depression related behaviors (Einat et al., 2006), which were reversed with bright light therapy administered at specific circadian phases (Ashkenazy et al., 2009), similar to the antidepressant effects of bright light therapy in human BD (Postolache and Oren, 2005, Wu et al., 2009).

Several mechanisms likely regulate the effects of photoperiod on mood related behavior. In the adult rat brain, long and short day photoperiods induced neurotransmitter plasticity and bidirectionally modulated anxiety and depression behavior (Dulcis et al., 2013). Under long days (19 h of light and 5 h of dark) over 7 days, hypothalamic neurons switched from dopamine to somatostatin phenotype, while the opposite was found under short days (19 h of dark and 5 h of light). Long days induced D2R expression on postsynaptic corticotropin release factor (CRF) hypothalamic neurons and increased CRF levels in the CSF and periphery. Long days increased anxiety and depression behavior. Lesions to presynaptic dopaminergic hypothalamic neurons or pharmacological blockade of dopamine D1 or D2 receptors recapitulated the behavioral effects of long days and could be partially rescued via photoperiodic induction of newly switched dopaminergic neurons (Dulcis et al., 2013). Short day exposure reduced anxiety and depression behaviors. Thus, the photoperiodic dependent switch between dopamine and somatostatin neuronal phenotypes, along with CRF levels, can modulate mood related behavior, revealing a potential mechanism by which seasonal variation could rapidly induce neurocircuit level changes that lead to shifts in mood.

Genetic models of circadian disruption have also shown promise for modeling mood switching in rodents. As previously mentioned, Clock mutant mice have disrupted molecular and behavioral rhythms (Vitaterna et al., 2006) and display a mania-like behavioral repertoire (McClung et al., 2005, Roybal et al., 2007). We have recently reported that anxiety and depression behaviors of Clock mutant mice naturally varied across the circadian day under entrained 12:12 light-dark conditions (Sidor et al., 2015). During the day, anxiety and depression behavior was reduced, while during the night, these behaviors normalized to wild-type levels (Sidor et al., 2015). Mania-like behaviors were found only when dopamine cell firing was highest during the day and these behaviors were normalized by dopamine depletion with AMPT treatment. Optogenetic stimulation of VTA dopamine neurons of wild-type mice during the day almost fully mimicked the mania-like behaviors of Clock mutant mice (Sidor et al., 2015). Thus diurnal variation of neurocircuit function underlies ‘normal’ variation of mood and altered circadian regulation of dopaminergic transmission could promote mania. We consider this ‘rapid cycling’ to be mostly related to an individual cycling from a manic to euthymic states. Sensorimotor gating and wheel-running activity, which is inherently rewarding, remain elevated during the night phase in Clock mutant mice (van Enkhuizen et al., 2014), suggesting diurnal variation may be specific to mood related behaviors and disparate mechanisms regulate these behaviors. Future studies will determine whether chronic stress can induce anxiety and depression behavior in Clock mutant mice. Preliminary also data suggests that Clock mutants may be more vulnerable to stress-induced mood related behaviors (unpublished results) similar to those described in the Ank3 mutant mice (Leussis et al., 2013).

Another circadian genetic model of BD is the D-binding protein (Dbp) KO mouse and actually has the opposite phenotypes to Clock mutant mice (Table 3). Dbp is a basic leucine zipper transcription factor involved in the regulation of circadian locomotor behavior and a component of the molecular clock (Lopez-Molina et al., 1997, Ripperger et al., 2000). A genetic locus has been mapped near the chromosomal region of Dbp for BD (Morissette et al., 1999). Using convergent functional genomics approaches to integrate gene expression data with human genetic linkage and association studies, highlighted Dbp among the highest ranked candidate genes related to BD (Le-Niculescu et al., 2008).. Dbp-KO mice had abnormal circadian behavior, disrupted homeostatic sleep regulation, reduced locomotor activity, a blunted locomotor response, and stereotopy following acute amphetamine (Le-Niculescu et al., 2008). Intriguingly, following chronic stress (i.e,. social isolation with intermittent acute stress exposure), Dbp-KO mice became hyperactive, whereas wild-type mice were hypoactive. Acute sleep deprivation induced a similar phenotype in these mice and was completely blocked by valproic acid treatment prior to sleep deprivation (Le-Niculescu et al., 2008). Additionally, Dbp-KO mice switched to a hyperhedonic phenotype during long periods of continuous access to alcohol containing drinking fluids (Le-Niculescu et al., 2008). It is unclear whether stress or acute sleep deprivation also induces anxiety and depression behaviors in Dbp-KO mice, or whether these mice are more vulnerable to circadian disrupting effects of changes in photoperiod. The Dbp-KO mouse model of depression and stress-induced mania seems promising for studying the cellular and molecular mechanisms regulating behavioral switching in mice.

Future directions

The major challenge for the field of translational neuroscience will continue to be the development, investigation, and interpretability of animal models of psychiatric diseases. In the absence of known etiology of BD, models based on clinical observations should be positioned to provide greater impact on our understanding of the disease process. Collaborative efforts are necessary to integrate data across many pharmacological, environmental, and genetic models of BD in order to elucidate shared and distinct disease related mechanisms. Further integration of these changes in animal models with data from patient samples should identify common mechanisms of a heterogeneous disease and potentially narrow the targets for therapeutic development. For example, transcriptome profiling in animals and human post-mortem tissues of well-defined disease cohorts will identify shared molecular patterns (Krishnan and Nestler, 2011).

With new technology come new opportunities, and neuroscience is benefiting tremendously from new tools and approaches. Pharmacological models that recapitulated a few of the core behavioral phenotypes of BD were the research standard. Despite their limitations and confounds, genetically engineered mutant, knockin, or knockout mice pushed the research into investigating the causal relationship between risk loci associated with the disease and their interaction with environmental factors. Now with readily available and more affordable genome editing tools, such as TALEN or CRISPR/Cas9 technologies, investigators can recreate the precise genetic polymorphisms associated with BD. For example, ‘humanized’ mice carrying either Val or Met alleles in the catechol-o-methyltransferase (COMT) gene, an enzyme that degrades cathecholamines, dopamine, epinephrine, and norepinephrine, support the human genetic findings suggesting the human COMT Val158Met polymorphism modulates behavioral and cognitive functioning (Risbrough et al., 2014). ‘Humanized’ rodent models can systematically investigate the effects of risk loci on the developmental and behavioral factors of BD (Malkesman et al., 2009). These models still present with a major challenge facing single gene knockout mice, whereby the selection of genetic background can be critical for observing any relevant phenotype and should be considered when creating these models of human disease.

Newly available genetically and phenotypically heterogenous and diverse mouse populations, such as the Collaborative Cross (Threadgill and Churchill, 2012) and Diversity Outbred (Svenson et al., 2012), are valuable resources for discovering and interrogating new loci related to mania and depression (Logan et al., 2013). High-density mapping and high-throughput genotyping tools are readily accessible for investigating the genetics of common or disparate behavioral phenotypes. Integrative functional genomics can then be used to eludicate biological and genetics mechanisms that cut across animal models and human disease. An example of this type of strategy has identified new mechanisms common to bipolar and addiction disorders that could be recapitulated in a mouse model (Le-Niculescu et al., 2008). Collaborative Cross and Diversity Outbred mice may also help in our understanding of how genetics contributes to differential treatment responses, such as in the case of haloperidol response (Crowley et al., 2014). Thus, the major advantage of using these mice to investigate disease relevant phenotypes and treatment response is to study the continuous and more complete spectrum of the behavior by maximizing the genetic and phenotype variance, eventually moving the field from the classical approach of the simple presence or absence of an intended phenotype. The Black Swiss mice model of mania has more genetic diversity than standard inbred strains and other genetic mouse models, making them a positive step in this direction.

In addition to developing better animal models, there should be equal efforts into creating more tractable translational behavioral assays. Efforts in this arena will provide cross-species validation and interpretation of animal findings. The BPM, PPI, and drug self-administration models are examples where behavioral phenotypes observed in patients can be readily tested and interpreted in animals. The field also needs cross-species validated assays of cognition, particularly those that assess risk-taking and impulsivity—constructs that interfere with problem solving and decision-making abilities. As discussed previously, BD patients are more impulsive and have impaired decision-making during risky situations, and cognitive deficits correlate with their functional outcome (Green, 2006, Burdick et al., 2011).

New technologies, tools, and cross-species models may push the field beyond past shortcomings. As technology improves and researchers adopt approaches that cut across multiple domains, translational research of neuroscience and psychiatry will make huge strides forward. Integrating data from self-report, neuroimaging, optogenetics, epigenetics, genetics, and postmortem, along with other approaches, will be critical for uncovering the etiology of BD and the development of novel therapeutics.

Conclusions

Rodent models are limited in their ability to capture the entirety of human BD. The combination of pharmacological, environmental, and genetic approaches will continue to bring the field closer to understanding the environmental, biological, and genetic etiological factors of the human disease. Preclinical models are also useful for screening novel compounds or discovering the therapeutic mechanisms of current treatments. It is important to continue to elucidate the cellular and molecular pathways regulating motivation, reward, and mood, and how alterations to these lead to behavioral and biological phenotypes related to BD. With the advent of new, more heterogeneous mouse populations, we can use forward genetics to begin to unravel the complex genetic mechanisms and environmental factors of mania related behaviors. Models of behavioral switching should aid in our understanding of biological mechanisms that contribute to mood cycling, which is the hallmark of BD.

Highlights.

In this review we highlight the work that has been done on multiple animal models of bipolar mania.

  1. Pharmacological approaches to induce hyperactivity as a model of human mania have been used for many years.

  2. Environmental manipulations are used to induce manic-like states which mimic multiple facets of this complex disorder.

  3. Most recent studies have focused on genetic models.

  4. Certain outbred strains of mice may be useful in modeling the polygenic nature of this disease.

  5. Ultimately, models that cycle through various mood states are needed.

Acknowledgments

Studies from our group were funded by IMHRO, NARSAD, MH082876, DA023988, NS058339 and The McKnight Foundation.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  1. Aberle H, Bauer A, Stappert J, Kispert A, Kemler R. beta-catenin is a target for the ubiquitin-proteasome pathway. EMBO J. 1997;16:3797–3804. doi: 10.1093/emboj/16.13.3797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Adida M, Clark L, Pomietto P, Kaladjian A, Besnier N, Azorin JM, Jeanningros R, Goodwin GM. Lack of insight may predict impaired decision making in manic patients. Bipolar Disord. 2008;10:829–837. doi: 10.1111/j.1399-5618.2008.00618.x. [DOI] [PubMed] [Google Scholar]
  3. Amsterdam JD, Shults J. Comparison of fluoxetine, olanzapine, and combined fluoxetine plus olanzapine initial therapy of bipolar type I and type II major depression--lack of manic induction. J Affect Disord. 2005;87:121–130. doi: 10.1016/j.jad.2005.02.018. [DOI] [PubMed] [Google Scholar]
  4. Anand A, Barkay G, Dzemidzic M, Albrecht D, Karne H, Zheng QH, Hutchins GD, Normandin MD, Yoder KK. Striatal dopamine transporter availability in unmedicated bipolar disorder. Bipolar Disord. 2011;13:406–413. doi: 10.1111/j.1399-5618.2011.00936.x. [DOI] [PubMed] [Google Scholar]
  5. Anand A, Darnell A, Miller HL, Berman RM, Cappiello A, Oren DA, Woods SW, Charney DS. Effect of catecholamine depletion on lithium-induced long-term remission of bipolar disorder. Biol Psychiatry. 1999;45:972–978. doi: 10.1016/s0006-3223(98)00293-5. [DOI] [PubMed] [Google Scholar]
  6. Antoch MP, Song EJ, Chang AM, Vitaterna MH, Zhao Y, Wilsbacher LD, Sangoram AM, King DP, Pinto LH, Takahashi JS. Functional identification of the mouse circadian Clock gene by transgenic BAC rescue. Cell. 1997;89:655–667. doi: 10.1016/s0092-8674(00)80246-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Arban R, Maraia G, Brackenborough K, Winyard L, Wilson A, Gerrard P, Large C. Evaluation of the effects of lamotrigine, valproate and carbamazepine in a rodent model of mania. Behav Brain Res. 2005;158:123–132. doi: 10.1016/j.bbr.2004.08.015. [DOI] [PubMed] [Google Scholar]
  8. Arent CO, Valvassori SS, Fries GR, Stertz L, Ferreira CL, Lopes-Borges J, Mariot E, Varela RB, Ornell F, Kapczinski F, Andersen ML, Quevedo J. Neuroanatomical profile of antimaniac effects of histone deacetylases inhibitors. Mol Neurobiol. 2011;43:207–214. doi: 10.1007/s12035-011-8178-0. [DOI] [PubMed] [Google Scholar]
  9. Arey RN, Enwright JF, 3rd, Spencer SM, Falcon E, Ozburn AR, Ghose S, Tamminga C, McClung CA. An important role for cholecystokinin, a CLOCK target gene, in the development and treatment of manic-like behaviors. Mol Psychiatry. 2014;19:342–350. doi: 10.1038/mp.2013.12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Ashkenazy T, Einat H, Kronfeld-Schor N. Effects of bright light treatment on depression- and anxiety-like behaviors of diurnal rodents maintained on a short daylight schedule. Behav Brain Res. 2009;201:343–346. doi: 10.1016/j.bbr.2009.03.005. [DOI] [PubMed] [Google Scholar]
  11. Azorin JM, Sapin C, Weiller E. Effect of asenapine on manic and depressive symptoms in bipolar I patients with mixed episodes: results from post hoc analyses. J Affect Disord. 2013;145:62–69. doi: 10.1016/j.jad.2012.07.013. [DOI] [PubMed] [Google Scholar]
  12. Bagot RC, Labonte B, Pena CJ, Nestler EJ. Epigenetic signaling in psychiatric disorders: stress and depression. Dialogues Clin Neurosci. 2014;16:281–295. doi: 10.31887/DCNS.2014.16.3/rbagot. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Barrett SL, Kelly C, Watson DR, Bell R, King DJ. Normal levels of prepulse inhibition in the euthymic phase of bipolar disorder. Psychol Med. 2005;35:1737–1746. doi: 10.1017/S0033291705005702. [DOI] [PubMed] [Google Scholar]
  14. Beaulieu JM. A role for Akt and glycogen synthase kinase-3 as integrators of dopamine and serotonin neurotransmission in mental health. J Psychiatry Neurosci. 2012;37:7–16. doi: 10.1503/jpn.110011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Beaulieu JM, Del’guidice T, Sotnikova TD, Lemasson M, Gainetdinov RR. Beyond cAMP: The Regulation of Akt and GSK3 by Dopamine Receptors. Front Mol Neurosci. 2011;4:38. doi: 10.3389/fnmol.2011.00038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Bechara A, Damasio AR, Damasio H, Anderson SW. Insensitivity to future consequences following damage to human prefrontal cortex. Cognition. 1994;50:7–15. doi: 10.1016/0010-0277(94)90018-3. [DOI] [PubMed] [Google Scholar]
  17. Bechara A, Damasio H, Tranel D, Damasio AR. Deciding advantageously before knowing the advantageous strategy. Science. 1997;275:1293–1295. doi: 10.1126/science.275.5304.1293. [DOI] [PubMed] [Google Scholar]
  18. Bechara A, Damasio H, Tranel D, Damasio AR. The Iowa Gambling Task and the somatic marker hypothesis: some questions and answers. Trends Cogn Sci. 2005;9:159–162. doi: 10.1016/j.tics.2005.02.002. discussion 162–154. [DOI] [PubMed] [Google Scholar]
  19. Behrens J, Jerchow BA, Wurtele M, Grimm J, Asbrand C, Wirtz R, Kuhl M, Wedlich D, Birchmeier W. Functional interaction of an axin homolog, conductin, with beta-catenin, APC, and GSK3beta. Science. 1998;280:596–599. doi: 10.1126/science.280.5363.596. [DOI] [PubMed] [Google Scholar]
  20. Benedetti F, Fresi F, Maccioni P, Smeraldi E. Behavioural sensitization to repeated sleep deprivation in a mice model of mania. Behav Brain Res. 2008;187:221–227. doi: 10.1016/j.bbr.2007.09.012. [DOI] [PubMed] [Google Scholar]
  21. Benedetti F, Serretti A, Colombo C, Barbini B, Lorenzi C, Campori E, Smeraldi E. Influence of CLOCK gene polymorphism on circadian mood fluctuation and illness recurrence in bipolar depression. American journal of medical genetics Part B, Neuropsychiatric genetics: the official publication of the International Society of Psychiatric Genetics. 2003;123B:23–26. doi: 10.1002/ajmg.b.20038. [DOI] [PubMed] [Google Scholar]
  22. Benes FM. Searching for unique endophenotypes for schizophrenia and bipolar disorder within neural circuits and their molecular regulatory mechanisms. Schizophr Bull. 2007;33:932–936. doi: 10.1093/schbul/sbm064. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Bennett V, Lambert S. Physiological roles of axonal ankyrins in survival of premyelinated axons and localization of voltage-gated sodium channels. J Neurocytol. 1999;28:303–318. doi: 10.1023/a:1007005528505. [DOI] [PubMed] [Google Scholar]
  24. Berggren U, Tallstedt L, Ahlenius S, Engel J. The effect of lithium on amphetamine-induced locomotor stimulation. Psychopharmacology (Berl) 1978;59:41–45. doi: 10.1007/BF00428028. [DOI] [PubMed] [Google Scholar]
  25. Berk M, Dean OM, Cotton SM, Gama CS, Kapczinski F, Fernandes B, Kohlmann K, Jeavons S, Hewitt K, Moss K, Allwang C, Schapkaitz I, Cobb H, Bush AI, Dodd S, Malhi GS. Maintenance N-acetyl cysteine treatment for bipolar disorder: a double-blind randomized placebo controlled trial. BMC Med. 2012;10:91. doi: 10.1186/1741-7015-10-91. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Boland EM, Bender RE, Alloy LB, Conner BT, Labelle DR, Abramson LY. Life events and social rhythms in bipolar spectrum disorders: an examination of social rhythm sensitivity. J Affect Disord. 2012;139:264–272. doi: 10.1016/j.jad.2012.01.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Brain PF, Al-Maliki S. Effects of lithium chloride injections on rank-related fighting, maternal aggression and locust-killing responses in naive and experienced ‘TO’ strain mice. Pharmacol Biochem Behav. 1979;10:663–669. doi: 10.1016/0091-3057(79)90318-6. [DOI] [PubMed] [Google Scholar]
  28. Bunney WE, Jr, Kopanda RT, Murphy DL. Sleep and behavioral changes possibly reflecting central receptor hypersensitivity following catecholamine synthesis inhibition in man. Acta Psychiatr Scand. 1977;56:189–203. doi: 10.1111/j.1600-0447.1977.tb03562.x. [DOI] [PubMed] [Google Scholar]
  29. Burdick KE, Goldberg TE, Cornblatt BA, Keefe RS, Gopin CB, Derosse P, Braga RJ, Malhotra AK. The MATRICS consensus cognitive battery in patients with bipolar I disorder. Neuropsychopharmacology. 2011;36:1587–1592. doi: 10.1038/npp.2011.36. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Cabungcal JH, Nicolas D, Kraftsik R, Cuenod M, Do KQ, Hornung JP. Glutathione deficit during development induces anomalies in the rat anterior cingulate GABAergic neurons: Relevance to schizophrenia. Neurobiol Dis. 2006;22:624–637. doi: 10.1016/j.nbd.2006.01.003. [DOI] [PubMed] [Google Scholar]
  31. Calil HM, Zwicker AP, Klepacz S. The effects of lithium carbonate on healthy volunteers: mood stabilization? Biol Psychiatry. 1990;27:711–722. doi: 10.1016/0006-3223(90)90586-q. [DOI] [PubMed] [Google Scholar]
  32. Can A, Schulze TG, Gould TD. Molecular actions and clinical pharmacogenetics of lithium therapy. Pharmacol Biochem Behav. 2014;123:3–16. doi: 10.1016/j.pbb.2014.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Cappeliez P, Moore E. Effects of lithium on an amphetamine animal model of bipolar disorder. Prog Neuropsychopharmacol Biol Psychiatry. 1990;14:347–358. doi: 10.1016/0278-5846(90)90023-a. [DOI] [PubMed] [Google Scholar]
  34. Carlezon WA, Jr, Chartoff EH. Intracranial self-stimulation (ICSS) in rodents to study the neurobiology of motivation. Nat Protoc. 2007;2:2987–2995. doi: 10.1038/nprot.2007.441. [DOI] [PubMed] [Google Scholar]
  35. Carroll CA, Vohs JL, O’Donnell BF, Shekhar A, Hetrick WP. Sensorimotor gating in manic and mixed episode bipolar disorder. Bipolar Disord. 2007;9:221–229. doi: 10.1111/j.1399-5618.2007.00415.x. [DOI] [PubMed] [Google Scholar]
  36. Challet E, Takahashi JS, Turek FW. Nonphotic phase-shifting in clock mutant mice. Brain Res. 2000;859:398–403. doi: 10.1016/s0006-8993(00)02040-0. [DOI] [PubMed] [Google Scholar]
  37. Chen G, Huang LD, Jiang YM, Manji HK. The mood-stabilizing agent valproate inhibits the activity of glycogen synthase kinase-3. J Neurochem. 1999;72:1327–1330. doi: 10.1046/j.1471-4159.2000.0721327.x. [DOI] [PubMed] [Google Scholar]
  38. Chesler EJ, Logan RW. Opportunities for bioinformatics in the classification of behavior and psychiatric disorders. Int Rev Neurobiol. 2012;104:183–211. doi: 10.1016/B978-0-12-398323-7.00008-2. [DOI] [PubMed] [Google Scholar]
  39. Christodoulou T, Lewis M, Ploubidis GB, Frangou S. The relationship of impulsivity to response inhibition and decision-making in remitted patients with bipolar disorder. Eur Psychiatry. 2006;21:270–273. doi: 10.1016/j.eurpsy.2006.04.006. [DOI] [PubMed] [Google Scholar]
  40. Chubb JE, Bradshaw NJ, Soares DC, Porteous DJ, Millar JK. The DISC locus in psychiatric illness. Mol Psychiatry. 2008;13:36–64. doi: 10.1038/sj.mp.4002106. [DOI] [PubMed] [Google Scholar]
  41. Clapcote SJ, Duffy S, Xie G, Kirshenbaum G, Bechard AR, Rodacker Schack V, Petersen J, Sinai L, Saab BJ, Lerch JP, Minassian BA, Ackerley CA, Sled JG, Cortez MA, Henderson JT, Vilsen B, Roder JC. Mutation I810N in the alpha3 isoform of Na+, K+-ATPase causes impairments in the sodium pump and hyperexcitability in the CNS. Proc Natl Acad Sci U S A. 2009;106:14085–14090. doi: 10.1073/pnas.0904817106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Clark L, Iversen SD, Goodwin GM. A neuropsychological investigation of prefrontal cortex involvement in acute mania. Am J Psychiatry. 2001;158:1605–1611. doi: 10.1176/appi.ajp.158.10.1605. [DOI] [PubMed] [Google Scholar]
  43. Cole TB, Giordano G, Co AL, Mohar I, Kavanagh TJ, Costa LG. Behavioral Characterization of GCLM-Knockout Mice, a Model for Enhanced Susceptibility to Oxidative Stress. J Toxicol. 2011;2011:157687. doi: 10.1155/2011/157687. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Cousins DA, Butts K, Young AH. The role of dopamine in bipolar disorder. Bipolar Disord. 2009;11:787–806. doi: 10.1111/j.1399-5618.2009.00760.x. [DOI] [PubMed] [Google Scholar]
  45. Crowley JJ, Kim Y, Lenarcic AB, Quackenbush CR, Barrick CJ, Adkins DE, Shaw GS, Miller DR, de Villena FP, Sullivan PF, Valdar W. Genetics of adverse reactions to haloperidol in a mouse diallel: a drug-placebo experiment and Bayesian causal analysis. Genetics. 2014;196:321–347. doi: 10.1534/genetics.113.156901. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. de Bartolomeis A, Buonaguro EF, Iasevoli F, Tomasetti C. The emerging role of dopamine-glutamate interaction and of the postsynaptic density in bipolar disorder pathophysiology: Implications for treatment. J Psychopharmacol. 2014;28:505–526. doi: 10.1177/0269881114523864. [DOI] [PubMed] [Google Scholar]
  47. De Sarno P, Li X, Jope RS. Regulation of Akt and glycogen synthase kinase-3 beta phosphorylation by sodium valproate and lithium. Neuropharmacology. 2002;43:1158–1164. doi: 10.1016/s0028-3908(02)00215-0. [DOI] [PubMed] [Google Scholar]
  48. de Visser L, Homberg JR, Mitsogiannis M, Zeeb FD, Rivalan M, Fitoussi A, Galhardo V, van den Bos R, Winstanley CA, Dellu-Hagedorn F. Rodent versions of the iowa gambling task: opportunities and challenges for the understanding of decision-making. Front Neurosci. 2011;5:109. doi: 10.3389/fnins.2011.00109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. De Wilde JE, Doogan DP. Fluvoxamine and chlorimipramine in endogenous depression. J Affect Disord. 1982;4:249–259. doi: 10.1016/0165-0327(82)90009-x. [DOI] [PubMed] [Google Scholar]
  50. Dean OM, van den Buuse M, Berk M, Copolov DL, Mavros C, Bush AI. N-acetyl cysteine restores brain glutathione loss in combined 2-cyclohexene-1-one and d-amphetamine-treated rats: relevance to schizophrenia and bipolar disorder. Neurosci Lett. 2011;499:149–153. doi: 10.1016/j.neulet.2011.05.027. [DOI] [PubMed] [Google Scholar]
  51. Decker S, Grider G, Cobb M, Li XP, Huff MO, El-Mallakh RS, Levy RS. Open field is more sensitive than automated activity monitor in documenting ouabain-induced hyperlocomotion in the development of an animal model for bipolar illness. Prog Neuropsychopharmacol Biol Psychiatry. 2000;24:455–462. doi: 10.1016/s0278-5846(99)00111-6. [DOI] [PubMed] [Google Scholar]
  52. Dencker D, Dias R, Pedersen ML, Husum H. Effect of the new antiepileptic drug retigabine in a rodent model of mania. Epilepsy Behav. 2008;12:49–53. doi: 10.1016/j.yebeh.2007.09.023. [DOI] [PubMed] [Google Scholar]
  53. Dencker D, Husum H. Antimanic efficacy of retigabine in a proposed mouse model of bipolar disorder. Behav Brain Res. 2010;207:78–83. doi: 10.1016/j.bbr.2009.09.040. [DOI] [PubMed] [Google Scholar]
  54. Dick DM, Foroud T, Flury L, Bowman ES, Miller MJ, Rau NL, Moe PR, Samavedy N, El-Mallakh R, Manji H, Glitz DA, Meyer ET, Smiley C, Hahn R, Widmark C, McKinney R, Sutton L, Ballas C, Grice D, Berrettini W, Byerley W, Coryell W, DePaulo R, MacKinnon DF, Gershon ES, Kelsoe JR, McMahon FJ, McInnis M, Murphy DL, Reich T, Scheftner W, Nurnberger JI., Jr Genomewide linkage analyses of bipolar disorder: a new sample of 250 pedigrees from the National Institute of Mental Health Genetics Initiative. Am J Hum Genet. 2003;73:107–114. doi: 10.1086/376562. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Dietz RM, Kiedrowski L, Shuttleworth CW. Contribution of Na(+)/Ca(2+) exchange to excessive Ca(2+) loading in dendrites and somata of CA1 neurons in acute slice. Hippocampus. 2007;17:1049–1059. doi: 10.1002/hipo.20336. [DOI] [PubMed] [Google Scholar]
  56. Dulcis D, Jamshidi P, Leutgeb S, Spitzer NC. Neurotransmitter switching in the adult brain regulates behavior. Science. 2013;340:449–453. doi: 10.1126/science.1234152. [DOI] [PubMed] [Google Scholar]
  57. Durak O, de Anda FC, Singh KK, Leussis MP, Petryshen TL, Sklar P, Tsai LH. Ankyrin-G regulates neurogenesis and Wnt signaling by altering the subcellular localization of beta-catenin. Mol Psychiatry. 2015;20:388–397. doi: 10.1038/mp.2014.42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Dwivedi T, Zhang H. Lithium-induced neuroprotection is associated with epigenetic modification of specific BDNF gene promoter and altered expression of apoptotic-regulatory proteins. Front Neurosci. 2014;8:457. doi: 10.3389/fnins.2014.00457. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Dzirasa K, Coque L, Sidor MM, Kumar S, Dancy EA, Takahashi JS, McClung CA, Nicolelis MA. Lithium ameliorates nucleus accumbens phase-signaling dysfunction in a genetic mouse model of mania. J Neurosci. 2010;30:16314–16323. doi: 10.1523/JNEUROSCI.4289-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Dzirasa K, McGarity DL, Bhattacharya A, Kumar S, Takahashi JS, Dunson D, McClung CA, Nicolelis MA. Impaired limbic gamma oscillatory synchrony during anxiety-related behavior in a genetic mouse model of bipolar mania. J Neurosci. 2011;31:6449–6456. doi: 10.1523/JNEUROSCI.6144-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Easton A, Arbuzova J, Turek FW. The circadian Clock mutation increases exploratory activity and escape-seeking behavior. Genes Brain Behav. 2003;2:11–19. doi: 10.1034/j.1601-183x.2003.00002.x. [DOI] [PubMed] [Google Scholar]
  62. Ehlers CL, Frank E, Kupfer DJ. Social zeitgebers and biological rhythms. A unified approach to understanding the etiology of depression. Arch Gen Psychiatry. 1988;45:948–952. doi: 10.1001/archpsyc.1988.01800340076012. [DOI] [PubMed] [Google Scholar]
  63. Eichelman B, Thoa NB, Perez-Cruet J. Alkali metal cations: effects on aggression and adrenal enzymes. Pharmacol Biochem Behav. 1973;1:121–123. doi: 10.1016/0091-3057(73)90066-x. [DOI] [PubMed] [Google Scholar]
  64. Einat H. Modelling facets of mania--new directions related to the notion of endophenotypes. J Psychopharmacol. 2006;20:714–722. doi: 10.1177/0269881106060241. [DOI] [PubMed] [Google Scholar]
  65. Einat H. Establishment of a battery of simple models for facets of bipolar disorder: a practical approach to achieve increased validity, better screening and possible insights into endophenotypes of disease. Behav Genet. 2007;37:244–255. doi: 10.1007/s10519-006-9093-4. [DOI] [PubMed] [Google Scholar]
  66. Einat H, Kronfeld-Schor N, Eilam D. Sand rats see the light: short photoperiod induces a depression-like response in a diurnal rodent. Behav Brain Res. 2006;173:153–157. doi: 10.1016/j.bbr.2006.06.006. [DOI] [PubMed] [Google Scholar]
  67. El-Mallakh RS, Decker S, Morris M, Li XP, Huff MO, El-Masri MA, Levy RS. Efficacy of olanzapine and haloperidol in an animal model of mania. Prog Neuropsychopharmacol Biol Psychiatry. 2006;30:1261–1264. doi: 10.1016/j.pnpbp.2006.04.003. [DOI] [PubMed] [Google Scholar]
  68. el-Mallakh RS, Harrison LT, Li R, Changaris DG, Levy RS. An animal model for mania: preliminary results. Prog Neuropsychopharmacol Biol Psychiatry. 1995;19:955–962. doi: 10.1016/0278-5846(95)00123-d. [DOI] [PubMed] [Google Scholar]
  69. Ellicott A, Hammen C, Gitlin M, Brown G, Jamison K. Life events and the course of bipolar disorder. Am J Psychiatry. 1990;147:1194–1198. doi: 10.1176/ajp.147.9.1194. [DOI] [PubMed] [Google Scholar]
  70. Ene HM, Kara NZ, Einat H. The effects of the atypical antipsychotic asenapine in a strain-specific battery of tests for mania-like behaviors. Behav Pharmacol. 2015 doi: 10.1097/FBP.0000000000000128. [DOI] [PubMed] [Google Scholar]
  71. Engel SR, Creson TK, Hao Y, Shen Y, Maeng S, Nekrasova T, Landreth GE, Manji HK, Chen G. The extracellular signal-regulated kinase pathway contributes to the control of behavioral excitement. Mol Psychiatry. 2009;14:448–461. doi: 10.1038/sj.mp.4002135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Etain B, Jamain S, Milhiet V, Lajnef M, Boudebesse C, Dumaine A, Mathieu F, Gombert A, Ledudal K, Gard S, Kahn JP, Henry C, Boland A, Zelenika D, Lechner D, Lathrop M, Leboyer M, Bellivier F. Association between circadian genes, bipolar disorders and chronotypes. Chronobiol Int. 2014;31:807–814. doi: 10.3109/07420528.2014.906445. [DOI] [PubMed] [Google Scholar]
  73. Evenden J. Impulsivity: a discussion of clinical and experimental findings. J Psychopharmacol. 1999;13:180–192. doi: 10.1177/026988119901300211. [DOI] [PubMed] [Google Scholar]
  74. Fessler RG, Sturgeon RD, London SF, Meltzer HY. Effects of lithium on behaviour induced by phencyclidine and amphetamine in rats. Psychopharmacology (Berl) 1982;78:373–376. doi: 10.1007/BF00433745. [DOI] [PubMed] [Google Scholar]
  75. Flaisher-Grinberg S, Einat H. Strain-specific battery of tests for domains of mania: effects of valproate, lithium and imipramine. Front Psychiatry. 2010;1:10. doi: 10.3389/fpsyt.2010.00010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Flaisher-Grinberg S, Kronfeld-Schor N, Einat H. Models of mania: from facets to domains and from animal models to model animals. J Psychopharmacol. 2010;24:437–438. doi: 10.1177/0269881108097905. [DOI] [PubMed] [Google Scholar]
  77. Frangou S, Kington J, Raymont V, Shergill SS. Examining ventral and dorsal prefrontal function in bipolar disorder: a functional magnetic resonance imaging study. Eur Psychiatry. 2008;23:300–308. doi: 10.1016/j.eurpsy.2007.05.002. [DOI] [PubMed] [Google Scholar]
  78. Frank E, Soreca I, Swartz HA, Fagiolini AM, Mallinger AG, Thase ME, Grochocinski VJ, Houck PR, Kupfer DJ. The role of interpersonal and social rhythm therapy in improving occupational functioning in patients with bipolar I disorder. Am J Psychiatry. 2008;165:1559–1565. doi: 10.1176/appi.ajp.2008.07121953. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Frank E, Swartz HA, Kupfer DJ. Interpersonal and social rhythm therapy: managing the chaos of bipolar disorder. Biological psychiatry. 2000;48:593–604. doi: 10.1016/s0006-3223(00)00969-0. [DOI] [PubMed] [Google Scholar]
  80. Fries GR, Valvassori SS, Bock H, Stertz L, Magalhaes PV, Mariot E, Varela RB, Kauer-Sant’Anna M, Quevedo J, Kapczinski F, Saraiva-Pereira ML. Memory and brain-derived neurotrophic factor after subchronic or chronic amphetamine treatment in an animal model of mania. J Psychiatr Res. 2015 doi: 10.1016/j.jpsychires.2015.05.006. [DOI] [PubMed] [Google Scholar]
  81. Gardner R., Jr Mechanisms in manic-depressive disorder: an evolutionary model. Arch Gen Psychiatry. 1982;39:1436–1441. doi: 10.1001/archpsyc.1982.04290120066013. [DOI] [PubMed] [Google Scholar]
  82. Gawryluk JW, Wang JF, Andreazza AC, Shao L, Yatham LN, Young LT. Prefrontal cortex glutathione S-transferase levels in patients with bipolar disorder, major depression and schizophrenia. Int J Neuropsychopharmacol. 2011a;14:1069–1074. doi: 10.1017/S1461145711000617. [DOI] [PubMed] [Google Scholar]
  83. Gawryluk JW, Wang JF, Andreazza AC, Shao L, Young LT. Decreased levels of glutathione, the major brain antioxidant, in post-mortem prefrontal cortex from patients with psychiatric disorders. Int J Neuropsychopharmacol. 2011b;14:123–130. doi: 10.1017/S1461145710000805. [DOI] [PubMed] [Google Scholar]
  84. Geoffroy PA, Bellivier F, Scott J, Boudebesse C, Lajnef M, Gard S, Kahn JP, Azorin JM, Henry C, Leboyer M, Etain B. Bipolar disorder with seasonal pattern: clinical characteristics and gender influences. Chronobiol Int. 2013;30:1101–1107. doi: 10.3109/07420528.2013.800091. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Gessa GL, Pani L, Fadda P, Fratta W. Sleep deprivation in the rat: an animal model of mania. Eur Neuropsychopharmacol. 1995;5(Suppl):89–93. doi: 10.1016/0924-977x(95)00023-i. [DOI] [PubMed] [Google Scholar]
  86. Geyer MA, Russo PV, Masten VL. Multivariate assessment of locomotor behavior: pharmacological and behavioral analyses. Pharmacol Biochem Behav. 1986;25:277–288. doi: 10.1016/0091-3057(86)90266-2. [DOI] [PubMed] [Google Scholar]
  87. Giakoumaki SG, Bitsios P, Frangou S, Roussos P, Aasen I, Galea A, Kumari V. Low baseline startle and deficient affective startle modulation in remitted bipolar disorder patients and their unaffected siblings. Psychophysiology. 2010;47:659–668. doi: 10.1111/j.1469-8986.2010.00977.x. [DOI] [PubMed] [Google Scholar]
  88. Gijsman HJ, Geddes JR, Rendell JM, Nolen WA, Goodwin GM. Antidepressants for bipolar depression: a systematic review of randomized, controlled trials. Am J Psychiatry. 2004;161:1537–1547. doi: 10.1176/appi.ajp.161.9.1537. [DOI] [PubMed] [Google Scholar]
  89. Gogos A, van den Buuse M, Rossell S. Gender differences in prepulse inhibition (PPI) in bipolar disorder: men have reduced PPI, women have increased PPI. Int J Neuropsychopharmacol. 2009;12:1249–1259. doi: 10.1017/S1461145709000480. [DOI] [PubMed] [Google Scholar]
  90. Goldberg JF, Burdick KE. Levetiracetam for acute mania. Am J Psychiatry. 2002;159:148. doi: 10.1176/appi.ajp.159.1.148. [DOI] [PubMed] [Google Scholar]
  91. Goldstein I, Lerer E, Laiba E, Mallet J, Mujaheed M, Laurent C, Rosen H, Ebstein RP, Lichtstein D. Association between sodium- and potassium-activated adenosine triphosphatase alpha isoforms and bipolar disorders. Biol Psychiatry. 2009;65:985–991. doi: 10.1016/j.biopsych.2008.10.033. [DOI] [PubMed] [Google Scholar]
  92. Goldstein TR, Fersch-Podrat R, Axelson DA, Gilbert A, Hlastala SA, Birmaher B, Frank E. Early intervention for adolescents at high risk for the development of bipolar disorder: pilot study of Interpersonal and Social Rhythm Therapy (IPSRT) Psychotherapy (Chic) 2014;51:180–189. doi: 10.1037/a0034396. [DOI] [PubMed] [Google Scholar]
  93. Gomez-Sintes R, Bortolozzi A, Artigas F, Lucas JJ. Reduced striatal dopamine DA D2 receptor function in dominant-negative GSK-3 transgenic mice. Eur Neuropsychopharmacol. 2014;24:1524–1533. doi: 10.1016/j.euroneuro.2014.07.004. [DOI] [PubMed] [Google Scholar]
  94. Gonzalez JM, Thompson PM, Moore TA. Review of the safety, efficacy, and side effect profile of asenapine in the treatment of bipolar 1 disorder. Patient Prefer Adherence. 2011;5:333–341. doi: 10.2147/PPA.S10968. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Gonzalez R, Bernardo C, Cruz D, Walss-Bass C, Thompson PM. The relationships between clinical characteristics, alcohol and psychotropic exposure, and circadian gene expression in human postmortem samples of affective disorder and control subjects. Psychiatry Res. 2014;218:359–362. doi: 10.1016/j.psychres.2014.04.048. [DOI] [PubMed] [Google Scholar]
  96. Gould TD, O’Donnell KC, Picchini AM, Manji HK. Strain differences in lithium attenuation of d-amphetamine-induced hyperlocomotion: a mouse model for the genetics of clinical response to lithium. Neuropsychopharmacology. 2007;32:1321–1333. doi: 10.1038/sj.npp.1301254. [DOI] [PubMed] [Google Scholar]
  97. Gould TJ, Keith RA, Bhat RV. Differential sensitivity to lithium’s reversal of amphetamine-induced open-field activity in two inbred strains of mice. Behav Brain Res. 2001;118:95–105. doi: 10.1016/s0166-4328(00)00318-1. [DOI] [PubMed] [Google Scholar]
  98. Gray CM, McCormick DA. Chattering cells: superficial pyramidal neurons contributing to the generation of synchronous oscillations in the visual cortex. Science. 1996;274:109–113. doi: 10.1126/science.274.5284.109. [DOI] [PubMed] [Google Scholar]
  99. Green MF. Cognitive impairment and functional outcome in schizophrenia and bipolar disorder. J Clin Psychiatry. 2006;67(Suppl 9):3–8. discussion 36–42. [PubMed] [Google Scholar]
  100. Greenwood TA, Schork NJ, Eskin E, Kelsoe JR. Identification of additional variants within the human dopamine transporter gene provides further evidence for an association with bipolar disorder in two independent samples. Mol Psychiatry. 2006;11:125–133. 115. doi: 10.1038/sj.mp.4001764. [DOI] [PubMed] [Google Scholar]
  101. Grisar T, Guillaume D, Delgado-Escueta AV. Contribution of Na+, K(+)-ATPase to focal epilepsy: a brief review. Epilepsy Res. 1992;12:141–149. doi: 10.1016/0920-1211(92)90034-q. [DOI] [PubMed] [Google Scholar]
  102. Grunze H, Langosch J, Born C, Schaub G, Walden J. Levetiracetam in the treatment of acute mania: an open add-on study with an on-off-on design. J Clin Psychiatry. 2003;64:781–784. doi: 10.4088/jcp.v64n0707. [DOI] [PubMed] [Google Scholar]
  103. Guilmatre A, Huguet G, Delorme R, Bourgeron T. The emerging role of SHANK genes in neuropsychiatric disorders. Dev Neurobiol. 2014;74:113–122. doi: 10.1002/dneu.22128. [DOI] [PubMed] [Google Scholar]
  104. Gustafson CL, Partch CL. Emerging models for the molecular basis of mammalian circadian timing. Biochemistry. 2015;54:134–149. doi: 10.1021/bi500731f. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Han K, Holder JL, Jr, Schaaf CP, Lu H, Chen H, Kang H, Tang J, Wu Z, Hao S, Cheung SW, Yu P, Sun H, Breman AM, Patel A, Lu HC, Zoghbi HY. SHANK3 overexpression causes manic-like behaviour with unique pharmacogenetic properties. Nature. 2013;503:72–77. doi: 10.1038/nature12630. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Hannah-Poquette C, Anderson GW, Flaisher-Grinberg S, Wang J, Meinerding TM, Einat H. Modeling mania: Further validation for Black Swiss mice as model animals. Behav Brain Res. 2011;223:222–226. doi: 10.1016/j.bbr.2011.04.047. [DOI] [PubMed] [Google Scholar]
  107. Hashimoto K. BDNF and proBDNF as biomarkers for bipolar disorder. Br J Psychiatry. 2014;205:410. doi: 10.1192/bjp.205.5.410. [DOI] [PubMed] [Google Scholar]
  108. Henry BL, Minassian A, Young JW, Paulus MP, Geyer MA, Perry W. Cross-species assessments of motor and exploratory behavior related to bipolar disorder. Neurosci Biobehav Rev. 2010;34:1296–1306. doi: 10.1016/j.neubiorev.2010.04.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Hicks RA, Moore JD. REM sleep deprivation diminishes fear in rats. Physiol Behav. 1979;22:689–692. doi: 10.1016/0031-9384(79)90232-4. [DOI] [PubMed] [Google Scholar]
  110. Hicks RA, Moore JD, Hayes C, Phillips N, Hawkins J. REM sleep deprivation increases aggressiveness in male rats. Physiol Behav. 1979;22:1097–1100. doi: 10.1016/0031-9384(79)90263-4. [DOI] [PubMed] [Google Scholar]
  111. Himmelhoch JM, Fuchs CZ, Symons BJ. A double-blind study of tranylcypromine treatment of major anergic depression. J Nerv Ment Dis. 1982;170:628–634. doi: 10.1097/00005053-198210000-00007. [DOI] [PubMed] [Google Scholar]
  112. Hlastala SA. Stress, social rhythms, and behavioral activation: psychosocial factors and the bipolar illness course. Curr Psychiatry Rep. 2003;5:477–483. doi: 10.1007/s11920-003-0087-0. [DOI] [PubMed] [Google Scholar]
  113. Homberg JR, van den Bos R, den Heijer E, Suer R, Cuppen E. Serotonin transporter dosage modulates long-term decision-making in rat and human. Neuropharmacology. 2008;55:80–84. doi: 10.1016/j.neuropharm.2008.04.016. [DOI] [PubMed] [Google Scholar]
  114. Horschitz S, Hummerich R, Lau T, Rietschel M, Schloss P. A dopamine transporter mutation associated with bipolar affective disorder causes inhibition of transporter cell surface expression. Mol Psychiatry. 2005;10:1104–1109. doi: 10.1038/sj.mp.4001730. [DOI] [PubMed] [Google Scholar]
  115. Houtepen LC, Boks MP, Kahn RS, Joels M, Vinkers CH. Antipsychotic use is associated with a blunted cortisol stress response: a study in euthymic bipolar disorder patients and their unaffected siblings. Eur Neuropsychopharmacol. 2015;25:77–84. doi: 10.1016/j.euroneuro.2014.10.005. [DOI] [PubMed] [Google Scholar]
  116. Jacobsen JP, Rodriguiz RM, Mork A, Wetsel WC. Monoaminergic dysregulation in glutathione-deficient mice: possible relevance to schizophrenia? Neuroscience. 2005;132:1055–1072. doi: 10.1016/j.neuroscience.2005.01.059. [DOI] [PubMed] [Google Scholar]
  117. Jiang WG, Li SX, Liu JF, Sun Y, Zhou SJ, Zhu WL, Shi J, Lu L. Hippocampal CLOCK protein participates in the persistence of depressive-like behavior induced by chronic unpredictable stress. Psychopharmacology (Berl) 2013;227:79–92. doi: 10.1007/s00213-012-2941-4. [DOI] [PubMed] [Google Scholar]
  118. Jin N, Kovacs AD, Sui Z, Dewhurst S, Maggirwar SB. Opposite effects of lithium and valproic acid on trophic factor deprivation-induced glycogen synthase kinase-3 activation, c-Jun expression and neuronal cell death. Neuropharmacology. 2005;48:576–583. doi: 10.1016/j.neuropharm.2004.11.010. [DOI] [PubMed] [Google Scholar]
  119. Jochum T, Bar KJ, Sauer H. Topiramate induced manic episode. J Neurol Neurosurg Psychiatry. 2002;73:208–209. doi: 10.1136/jnnp.73.2.208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Jope RS, Roh MS. Glycogen synthase kinase-3 (GSK3) in psychiatric diseases and therapeutic interventions. Curr Drug Targets. 2006;7:1421–1434. doi: 10.2174/1389450110607011421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Jornada LK, Valvassori SS, Arent CO, Leffa D, Damiani AA, Hainzenreder G, Ferreira CL, Moretti M, Andrade VM, Quevedo J. DNA damage after intracerebroventricular injection of ouabain in rats. Neurosci Lett. 2010;471:6–9. doi: 10.1016/j.neulet.2009.12.072. [DOI] [PubMed] [Google Scholar]
  122. Jornada LK, Valvassori SS, Steckert AV, Moretti M, Mina F, Ferreira CL, Arent CO, Dal-Pizzol F, Quevedo J. Lithium and valproate modulate antioxidant enzymes and prevent ouabain-induced oxidative damage in an animal model of mania. J Psychiatr Res. 2011;45:162–168. doi: 10.1016/j.jpsychires.2010.05.011. [DOI] [PubMed] [Google Scholar]
  123. Joyce PR, Fergusson DM, Woollard G, Abbott RM, Horwood LJ, Upton J. Urinary catecholamines and plasma hormones predict mood state in rapid cycling bipolar affective disorder. J Affect Disord. 1995;33:233–243. doi: 10.1016/0165-0327(94)00094-p. [DOI] [PubMed] [Google Scholar]
  124. Jung SH, Park JM, Moon E, Chung YI, Lee BD, Lee YM, Kim JH, Kim SY, Jeong HJ. Delay in the recovery of normal sleep-wake cycle after disruption of the light-dark cycle in mice: a bipolar disorder-prone animal model? Psychiatry Investig. 2014;11:487–491. doi: 10.4306/pi.2014.11.4.487. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Kaladchibachi SA, Doble B, Anthopoulos N, Woodgett JR, Manoukian AS. Glycogen synthase kinase 3, circadian rhythms, and bipolar disorder: a molecular link in the therapeutic action of lithium. J Circadian Rhythms. 2007;5:3. doi: 10.1186/1740-3391-5-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Kalinichev M, Dawson LA. Evidence for antimanic efficacy of glycogen synthase kinase-3 (GSK3) inhibitors in a strain-specific model of acute mania. Int J Neuropsychopharmacol. 2011;14:1051–1067. doi: 10.1017/S1461145710001495. [DOI] [PubMed] [Google Scholar]
  127. Kam JW, Bolbecker AR, O’Donnell BF, Hetrick WP, Brenner CA. Resting state EEG power and coherence abnormalities in bipolar disorder and schizophrenia. J Psychiatr Res. 2013;47:1893–1901. doi: 10.1016/j.jpsychires.2013.09.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Kelly MP, Logue SF, Dwyer JM, Beyer CE, Majchrowski H, Cai Z, Liu Z, Adedoyin A, Rosenzweig-Lipson S, Comery TA. The supra-additive hyperactivity caused by an amphetamine-chlordiazepoxide mixture exhibits an inverted-U dose response: negative implications for the use of a model in screening for mood stabilizers. Pharmacol Biochem Behav. 2009;92:649–654. doi: 10.1016/j.pbb.2009.03.003. [DOI] [PubMed] [Google Scholar]
  129. Kim MJ, Kwon JS, Suh SH, Suh JK, Jung J, Lee SN, Kim YH, Cho MC, Oh GT, Lee K. Transgenic overexpression of translationally controlled tumor protein induces systemic hypertension via repression of Na+, K+-ATPase. J Mol Cell Cardiol. 2008;44:151–159. doi: 10.1016/j.yjmcc.2007.09.017. [DOI] [PubMed] [Google Scholar]
  130. Kim YT, Lee SJ, Kim SH. Effects of the history of conduct disorder on the Iowa Gambling Tasks. Alcohol Clin Exp Res. 2006;30:466–472. doi: 10.1111/j.1530-0277.2006.00051.x. [DOI] [PubMed] [Google Scholar]
  131. King DP, Vitaterna MH, Chang AM, Dove WF, Pinto LH, Turek FW, Takahashi JS. The mouse Clock mutation behaves as an antimorph and maps within the W19H deletion, distal of Kit. Genetics. 1997a;146:1049–1060. doi: 10.1093/genetics/146.3.1049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. King DP, Zhao Y, Sangoram AM, Wilsbacher LD, Tanaka M, Antoch MP, Steeves TD, Vitaterna MH, Kornhauser JM, Lowrey PL, Turek FW, Takahashi JS. Positional cloning of the mouse circadian clock gene. Cell. 1997b;89:641–653. doi: 10.1016/s0092-8674(00)80245-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Kirshenbaum GS, Burgess CR, Dery N, Fahnestock M, Peever JH, Roder JC. Attenuation of mania-like behavior in Na(+), K(+)-ATPase alpha3 mutant mice by prospective therapies for bipolar disorder: melatonin and exercise. Neuroscience. 2014;260:195–204. doi: 10.1016/j.neuroscience.2013.12.011. [DOI] [PubMed] [Google Scholar]
  134. Kirshenbaum GS, Clapcote SJ, Petersen J, Vilsen B, Ralph MR, Roder JC. Genetic suppression of agrin reduces mania-like behavior in Na+, K+-ATPase alpha3 mutant mice. Genes Brain Behav. 2012;11:436–443. doi: 10.1111/j.1601-183X.2012.00800.x. [DOI] [PubMed] [Google Scholar]
  135. 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;10:542–550. doi: 10.1111/j.1601-183X.2011.00691.x. [DOI] [PubMed] [Google Scholar]
  136. Kishi T, Elmquist JK. Body weight is regulated by the brain: a link between feeding and emotion. Mol Psychiatry. 2005;10:132–146. doi: 10.1038/sj.mp.4001638. [DOI] [PubMed] [Google Scholar]
  137. Koszewska I, Rybakowski JK. Antidepressant-induced mood conversions in bipolar disorder: a retrospective study of tricyclic versus non-tricyclic antidepressant drugs. Neuropsychobiology. 2009;59:12–16. doi: 10.1159/000202824. [DOI] [PubMed] [Google Scholar]
  138. Kovacsics CE, Gould TD. Shock-induced aggression in mice is modified by lithium. Pharmacol Biochem Behav. 2010;94:380–386. doi: 10.1016/j.pbb.2009.09.020. [DOI] [PubMed] [Google Scholar]
  139. Kozikowski AP, Gaisina IN, Yuan H, Petukhov PA, Blond SY, Fedolak A, Caldarone B, McGonigle P. Structure-based design leads to the identification of lithium mimetics that block mania-like effects in rodents. possible new GSK-3beta therapies for bipolar disorders. J Am Chem Soc. 2007;129:8328–8332. doi: 10.1021/ja068969w. [DOI] [PubMed] [Google Scholar]
  140. Kripke DF, Mullaney DJ, Atkinson M, Wolf S. Circadian rhythm disorders in manic-depressives. Biological psychiatry. 1978;13:335–351. [PubMed] [Google Scholar]
  141. Kripke DF, Nievergelt CM, Joo E, Shekhtman T, Kelsoe JR. Circadian polymorphisms associated with affective disorders. J Circadian Rhythms. 2009;7:2. doi: 10.1186/1740-3391-7-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Krishnan KR. Psychiatric and medical comorbidities of bipolar disorder. Psychosom Med. 2005;67:1–8. doi: 10.1097/01.psy.0000151489.36347.18. [DOI] [PubMed] [Google Scholar]
  143. Kruger S, Sarkar R, Pietsch R, Hasenclever D, Braunig P. Levetiracetam as monotherapy or add-on to valproate in the treatment of acute mania-a randomized open-label study. Psychopharmacology (Berl) 2008;198:297–299. doi: 10.1007/s00213-008-1109-8. [DOI] [PubMed] [Google Scholar]
  144. Kucyi A, Alsuwaidan MT, Liauw SS, McIntyre RS. Aerobic physical exercise as a possible treatment for neurocognitive dysfunction in bipolar disorder. Postgrad Med. 2010;122:107–116. doi: 10.3810/pgm.2010.11.2228. [DOI] [PubMed] [Google Scholar]
  145. Kulak A, Cuenod M, Do KQ. Behavioral phenotyping of glutathione-deficient mice: relevance to schizophrenia and bipolar disorder. Behav Brain Res. 2012;226:563–570. doi: 10.1016/j.bbr.2011.10.020. [DOI] [PubMed] [Google Scholar]
  146. Kulak A, Steullet P, Cabungcal JH, Werge T, Ingason A, Cuenod M, Do KQ. Redox dysregulation in the pathophysiology of schizophrenia and bipolar disorder: insights from animal models. Antioxid Redox Signal. 2013;18:1428–1443. doi: 10.1089/ars.2012.4858. [DOI] [PubMed] [Google Scholar]
  147. Kushner SF, Khan A, Lane R, Olson WH. Topiramate monotherapy in the management of acute mania: results of four double-blind placebo-controlled trials. Bipolar Disord. 2006;8:15–27. doi: 10.1111/j.1399-5618.2006.00276.x. [DOI] [PubMed] [Google Scholar]
  148. Lamberty Y, Margineanu DG, Klitgaard H. Effect of the New Antiepileptic Drug Levetiracetam in an Animal Model of Mania. Epilepsy Behav. 2001;2:454–459. doi: 10.1006/ebeh.2001.0254. [DOI] [PubMed] [Google Scholar]
  149. Lamont EW, Coutu DL, Cermakian N, Boivin DB. Circadian rhythms and clock genes in psychotic disorders. Isr J Psychiatry Relat Sci. 2010;47:27–35. [PubMed] [Google Scholar]
  150. Le-Niculescu H, McFarland MJ, Ogden CA, Balaraman Y, Patel S, Tan J, Rodd ZA, Paulus M, Geyer MA, Edenberg HJ, Glatt SJ, Faraone SV, Nurnberger JI, Kuczenski R, Tsuang MT, Niculescu AB. Phenomic, convergent functional genomic, and biomarker studies in a stress-reactive genetic animal model of bipolar disorder and co-morbid alcoholism. Am J Med Genet B Neuropsychiatr Genet. 2008;147B:134–166. doi: 10.1002/ajmg.b.30707. [DOI] [PubMed] [Google Scholar]
  151. Lee KW, Woon PS, Teo YY, Sim K. Genome wide association studies (GWAS) and copy number variation (CNV) studies of the major psychoses: what have we learnt? Neurosci Biobehav Rev. 2012;36:556–571. doi: 10.1016/j.neubiorev.2011.09.001. [DOI] [PubMed] [Google Scholar]
  152. Lee SH, Ripke S, Neale BM, Faraone SV, Purcell SM, Perlis RH, et al. Genetic relationship between five psychiatric disorders estimated from genome-wide SNPs. Nat Genet. 2013;45:984–994. doi: 10.1038/ng.2711. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Legrand R, Fielder R. Role of dominance-submission relationships in shock-induced fighting of mice. J Comp Physiol Psychol. 1973;82:501–506. doi: 10.1037/h0034127. [DOI] [PubMed] [Google Scholar]
  154. Leussis MP, Berry-Scott EM, Saito M, Jhuang H, de Haan G, Alkan O, Luce CJ, Madison JM, Sklar P, Serre T, Root DE, Petryshen TL. The ANK3 bipolar disorder gene regulates psychiatric-related behaviors that are modulated by lithium and stress. Biol Psychiatry. 2013;73:683–690. doi: 10.1016/j.biopsych.2012.10.016. [DOI] [PubMed] [Google Scholar]
  155. Li J, Lu WQ, Beesley S, Loudon AS, Meng QJ. Lithium impacts on the amplitude and period of the molecular circadian clockwork. PLoS One. 2012;7:e33292. doi: 10.1371/journal.pone.0033292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Li R, el-Mallakh RS, Harrison L, Changaris DG, Levy RS. Lithium prevents ouabain-induced behavioral changes. Toward an animal model for manic depression. Mol Chem Neuropathol. 1997;31:65–72. doi: 10.1007/BF02815161. [DOI] [PubMed] [Google Scholar]
  157. Li W, Xiong Y, Shang C, Twu KY, Hang CT, Yang J, Han P, Lin CY, Lin CJ, Tsai FC, Stankunas K, Meyer T, Bernstein D, Pan M, Chang CP. Brg1 governs distinct pathways to direct multiple aspects of mammalian neural crest cell development. Proc Natl Acad Sci U S A. 2013;110:1738–1743. doi: 10.1073/pnas.1218072110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Lin YF, Huang MC, Liu HC. Glycogen synthase kinase 3beta gene polymorphisms may be associated with bipolar I disorder and the therapeutic response to lithium. J Affect Disord. 2013;147:401–406. doi: 10.1016/j.jad.2012.08.025. [DOI] [PubMed] [Google Scholar]
  159. Logan RW, Robledo RF, Recla JM, Philip VM, Bubier JA, Jay JJ, Harwood C, Wilcox T, Gatti DM, Bult CJ, Churchill GA, Chesler EJ. High-precision genetic mapping of behavioral traits in the diversity outbred mouse population. Genes Brain Behav. 2013;12:424–437. doi: 10.1111/gbb.12029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Logan RW, Williams WP, McClung CA. Circadian rhythms and addiction: Mechanistic insights and future directions. Behav Neurosci. 2014;128:387–412. doi: 10.1037/a0036268. [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Lopez-Molina L, Conquet F, Dubois-Dauphin M, Schibler U. The DBP gene is expressed according to a circadian rhythm in the suprachiasmatic nucleus and influences circadian behavior. EMBO J. 1997;16:6762–6771. doi: 10.1093/emboj/16.22.6762. [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Machado-Vieira R, Dietrich MO, Leke R, Cereser VH, Zanatto V, Kapczinski F, Souza DO, Portela LV, Gentil V. Decreased plasma brain derived neurotrophic factor levels in unmedicated bipolar patients during manic episode. Biol Psychiatry. 2007;61:142–144. doi: 10.1016/j.biopsych.2006.03.070. [DOI] [PubMed] [Google Scholar]
  163. Machado-Vieira R, Soeiro-De-Souza MG, Richards EM, Teixeira AL, Zarate CA., Jr Multiple levels of impaired neural plasticity and cellular resilience in bipolar disorder: developing treatments using an integrated translational approach. World J Biol Psychiatry. 2014;15:84–95. doi: 10.3109/15622975.2013.830775. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Malatynska E, Knapp RJ. Dominant-submissive behavior as models of mania and depression. Neurosci Biobehav Rev. 2005;29:715–737. doi: 10.1016/j.neubiorev.2005.03.014. [DOI] [PubMed] [Google Scholar]
  165. Malatynska E, Pinhasov A, Crooke JJ, Smith-Swintosky VL, Brenneman DE. Reduction of dominant or submissive behaviors as models for antimanic or antidepressant drug testing: technical considerations. J Neurosci Methods. 2007;165:175–182. doi: 10.1016/j.jneumeth.2007.05.035. [DOI] [PubMed] [Google Scholar]
  166. Malkesman O, Austin DR, Chen G, Manji HK. Reverse translational strategies for developing animal models of bipolar disorder. Dis Model Mech. 2009;2:238–245. doi: 10.1242/dmm.001628. [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. Malkoff-Schwartz S, Frank E, Anderson B, Sherrill JT, Siegel L, Patterson D, Kupfer DJ. Stressful life events and social rhythm disruption in the onset of manic and depressive bipolar episodes: a preliminary investigation. Arch Gen Psychiatry. 1998;55:702–707. doi: 10.1001/archpsyc.55.8.702. [DOI] [PubMed] [Google Scholar]
  168. Mamiya T, Kise M, Morikawa K. Ferulic acid attenuated cognitive deficits and increase in carbonyl proteins induced by buthionine-sulfoximine in mice. Neurosci Lett. 2008;430:115–118. doi: 10.1016/j.neulet.2007.10.029. [DOI] [PubMed] [Google Scholar]
  169. Marston HM, Young JW, Martin FD, Serpa KA, Moore CL, Wong EH, Gold L, Meltzer LT, Azar MR, Geyer MA, Shahid M. Asenapine effects in animal models of psychosis and cognitive function. Psychopharmacology (Berl) 2009;206:699–714. doi: 10.1007/s00213-009-1570-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. McCarthy MJ, Wei H, Marnoy Z, Darvish RM, McPhie DL, Cohen BM, Welsh DK. Genetic and clinical factors predict lithium’s effects on PER2 gene expression rhythms in cells from bipolar disorder patients. Transl Psychiatry. 2013;3:e318. doi: 10.1038/tp.2013.90. [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. McCarthy MJ, Welsh DK. Cellular circadian clocks in mood disorders. J Biol Rhythms. 2012;27:339–352. doi: 10.1177/0748730412456367. [DOI] [PubMed] [Google Scholar]
  172. McClung CA. Circadian genes, rhythms and the biology of mood disorders. Pharmacol Ther. 2007;114:222–232. doi: 10.1016/j.pharmthera.2007.02.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. McClung CA. How might circadian rhythms control mood? Let me count the ways. Biol Psychiatry. 2013;74:242–249. doi: 10.1016/j.biopsych.2013.02.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. McClung CA, Sidiropoulou K, Vitaterna M, Takahashi JS, White FJ, Cooper DC, Nestler EJ. Regulation of dopaminergic transmission and cocaine reward by the Clock gene. Proc Natl Acad Sci U S A. 2005;102:9377–9381. doi: 10.1073/pnas.0503584102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. McElroy SL, Weisler RH, Chang W, Olausson B, Paulsson B, Brecher M, Agambaram V, Merideth C, Nordenhem A, Young AH. A double-blind, placebo-controlled study of quetiapine and paroxetine as monotherapy in adults with bipolar depression (EMBOLDEN II) J Clin Psychiatry. 2010;71:163–174. doi: 10.4088/JCP.08m04942gre. [DOI] [PubMed] [Google Scholar]
  176. McInerney SJ, Kennedy SH. Review of evidence for use of antidepressants in bipolar depression. Prim Care Companion CNS Disord. 2014:16. doi: 10.4088/PCC.14r01653. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Meyendorff E, Lerer B, Moore NC, Bow J, Gershon S. Methylphenidate infusion in euthymic bipolars: effect of carbamazepine pretreatment. Psychiatry Res. 1985;16:303–308. doi: 10.1016/0165-1781(85)90121-0. [DOI] [PubMed] [Google Scholar]
  178. Miczek KA, Maxson SC, Fish EW, Faccidomo S. Aggressive behavioral phenotypes in mice. Behav Brain Res. 2001;125:167–181. doi: 10.1016/s0166-4328(01)00298-4. [DOI] [PubMed] [Google Scholar]
  179. Miczek KA, O’Donnell JM. Intruder-evoked aggression in isolated and nonisolated mice: effects of psychomotor stimulants and L-dopa. Psychopharmacology (Berl) 1978;57:47–55. doi: 10.1007/BF00426957. [DOI] [PubMed] [Google Scholar]
  180. Milhiet V, Etain B, Boudebesse C, Bellivier F. Circadian biomarkers, circadian genes and bipolar disorders. J Physiol Paris. 2011;105:183–189. doi: 10.1016/j.jphysparis.2011.07.002. [DOI] [PubMed] [Google Scholar]
  181. Minassian A, Henry BL, Geyer MA, Paulus MP, Young JW, Perry W. The quantitative assessment of motor activity in mania and schizophrenia. J Affect Disord. 2010;120:200–206. doi: 10.1016/j.jad.2009.04.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Minassian A, Henry BL, Young JW, Masten V, Geyer MA, Perry W. Repeated assessment of exploration and novelty seeking in the human behavioral pattern monitor in bipolar disorder patients and healthy individuals. PLoS One. 2011;6:e24185. doi: 10.1371/journal.pone.0024185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Mitchell PJ. Antidepressant treatment and rodent aggressive behaviour. Eur J Pharmacol. 2005;526:147–162. doi: 10.1016/j.ejphar.2005.09.029. [DOI] [PubMed] [Google Scholar]
  184. Moore JD, Hayes C, Hicks RA. REM sleep deprivation increases preference for novelty in rats. Physiol Behav. 1979;23:975–976. doi: 10.1016/0031-9384(79)90211-7. [DOI] [PubMed] [Google Scholar]
  185. Moore PJ, Little MA, McSharry PE, Goodwin GM, Geddes JR. Correlates of depression in bipolar disorder. Proc Biol Sci. 2014;281:20132320. doi: 10.1098/rspb.2013.2320. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Morden B, Mitchell G, Dement W. Selective REM sleep deprivation and compensation phenomena in the rat. Brain Res. 1967;5:339–349. doi: 10.1016/0006-8993(67)90042-x. [DOI] [PubMed] [Google Scholar]
  187. Morissette J, Villeneuve A, Bordeleau L, Rochette D, Laberge C, Gagne B, Laprise C, Bouchard G, Plante M, Gobeil L, Shink E, Weissenbach J, Barden N. Genome-wide search for linkage of bipolar affective disorders in a very large pedigree derived from a homogeneous population in quebec points to a locus of major effect on chromosome 12q23-q24. Am J Med Genet. 1999;88:567–587. doi: 10.1002/(sici)1096-8628(19991015)88:5<567::aid-ajmg24>3.0.co;2-8. [DOI] [PubMed] [Google Scholar]
  188. Mukherjee BP, Pradhan SN. Effects of lithium on foot shock-induced aggressive behavior in rats. Arch Int Pharmacodyn Ther. 1976;222:125–131. [PubMed] [Google Scholar]
  189. Mukherjee S, Coque L, Cao JL, Kumar J, Chakravarty S, Asaithamby A, Graham A, Gordon E, Enwright JF, 3rd, DiLeone RJ, Birnbaum SG, Cooper DC, McClung CA. Knockdown of Clock in the ventral tegmental area through RNA interference results in a mixed state of mania and depression-like behavior. Biol Psychiatry. 2010;68:503–511. doi: 10.1016/j.biopsych.2010.04.031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Mynett-Johnson L, Murphy V, McCormack J, Shields DC, Claffey E, Manley P, McKeon P. Evidence for an allelic association between bipolar disorder and a Na+, K+ adenosine triphosphatase alpha subunit gene (ATP1A3) Biol Psychiatry. 1998;44:47–51. doi: 10.1016/s0006-3223(97)00343-0. [DOI] [PubMed] [Google Scholar]
  191. Nejtek VA, Kaiser KA, Zhang B, Djokovic M. Iowa Gambling Task scores predict future drug use in bipolar disorder outpatients with stimulant dependence. Psychiatry Res. 2013;210:871–879. doi: 10.1016/j.psychres.2013.08.021. [DOI] [PubMed] [Google Scholar]
  192. Nestler EJ. Epigenetic mechanisms of depression. JAMA Psychiatry. 2014;71:454–456. doi: 10.1001/jamapsychiatry.2013.4291. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Ng F, Dodd S, Berk M. The effects of physical activity in the acute treatment of bipolar disorder: a pilot study. J Affect Disord. 2007;101:259–262. doi: 10.1016/j.jad.2006.11.014. [DOI] [PubMed] [Google Scholar]
  194. Nunez PL, Srinivasan R, Westdorp AF, Wijesinghe RS, Tucker DM, Silberstein RB, Cadusch PJ. EEG coherency. I: Statistics, reference electrode, volume conduction, Laplacians, cortical imaging, and interpretation at multiple scales. Electroencephalogr Clin Neurophysiol. 1997;103:499–515. doi: 10.1016/s0013-4694(97)00066-7. [DOI] [PubMed] [Google Scholar]
  195. O’Brien WT, Harper AD, Jove F, Woodgett JR, Maretto S, Piccolo S, Klein PS. Glycogen synthase kinase-3beta haploinsufficiency mimics the behavioral and molecular effects of lithium. J Neurosci. 2004;24:6791–6798. doi: 10.1523/JNEUROSCI.4753-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  196. O’Brien WT, Huang J, Buccafusca R, Garskof J, Valvezan AJ, Berry GT, Klein PS. Glycogen synthase kinase-3 is essential for beta-arrestin-2 complex formation and lithium-sensitive behaviors in mice. J Clin Invest. 2011;121:3756–3762. doi: 10.1172/JCI45194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  197. O’Donnell BF, Hetrick WP, Vohs JL, Krishnan GP, Carroll CA, Shekhar A. Neural synchronization deficits to auditory stimulation in bipolar disorder. Neuroreport. 2004;15:1369–1372. doi: 10.1097/01.wnr.0000127348.64681.b2. [DOI] [PubMed] [Google Scholar]
  198. O’Donnell KC, Gould TD. The behavioral actions of lithium in rodent models: leads to develop novel therapeutics. Neurosci Biobehav Rev. 2007;31:932–962. doi: 10.1016/j.neubiorev.2007.04.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  199. Ochs SM, Dorostkar MM, Aramuni G, Schon C, Filser S, Poschl J, Kremer A, Van Leuven F, Ovsepian SV, Herms J. Loss of neuronal GSK3beta reduces dendritic spine stability and attenuates excitatory synaptic transmission via beta-catenin. Mol Psychiatry. 2015;20:482–489. doi: 10.1038/mp.2014.55. [DOI] [PMC free article] [PubMed] [Google Scholar]
  200. Oehler J, Jahkel M, Schmidt J. Effect of lithium, carbamazepine, ca-valproate and diazepam on changes in social isolation-induced behavior in mice. Biomed Biochim Acta. 1985;44:1523–1530. [PubMed] [Google Scholar]
  201. Ongur D, Lundy M, Greenhouse I, Shinn AK, Menon V, Cohen BM, Renshaw PF. Default mode network abnormalities in bipolar disorder and schizophrenia. Psychiatry Res. 2010;183:59–68. doi: 10.1016/j.pscychresns.2010.04.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  202. Orford K, Crockett C, Jensen JP, Weissman AM, Byers SW. Serine phosphorylation-regulated ubiquitination and degradation of beta-catenin. J Biol Chem. 1997;272:24735–24738. doi: 10.1074/jbc.272.40.24735. [DOI] [PubMed] [Google Scholar]
  203. Ozburn AR, Falcon E, Mukherjee S, Gillman A, Arey R, Spencer S, McClung CA. The role of clock in ethanol-related behaviors. Neuropsychopharmacology. 2013;38:2393–2400. doi: 10.1038/npp.2013.138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Ozburn AR, Larson EB, Self DW, McClung CA. Cocaine self-administration behaviors in ClockDelta19 mice. Psychopharmacology (Berl) 2012;223:169–177. doi: 10.1007/s00213-012-2704-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  205. Ozerdem A, Guntekin B, Atagun I, Turp B, Basar E. Reduced long distance gamma (28–48 Hz) coherence in euthymic patients with bipolar disorder. J Affect Disord. 2011;132:325–332. doi: 10.1016/j.jad.2011.02.028. [DOI] [PubMed] [Google Scholar]
  206. Ozerdem A, Guntekin B, Saatci E, Tunca Z, Basar E. Disturbance in long distance gamma coherence in bipolar disorder. Prog Neuropsychopharmacol Biol Psychiatry. 2010;34:861–865. doi: 10.1016/j.pnpbp.2010.04.001. [DOI] [PubMed] [Google Scholar]
  207. Ozerdem A, Guntekin B, Tunca Z, Basar E. Brain oscillatory responses in patients with bipolar disorder manic episode before and after valproate treatment. Brain Res. 2008;1235:98–108. doi: 10.1016/j.brainres.2008.06.101. [DOI] [PubMed] [Google Scholar]
  208. Ozerdema A, Guntekind B, Atagune MI, Basar E. Brain oscillations in bipolar disorder in search of new biomarkers. Suppl Clin Neurophysiol. 2013;62:207–221. doi: 10.1016/b978-0-7020-5307-8.00014-4. [DOI] [PubMed] [Google Scholar]
  209. Pandey GN, Rizavi HS, Tripathi M, Ren X. Region-specific dysregulation of glycogen synthase kinase-3beta and beta-catenin in the postmortem brains of subjects with bipolar disorder and schizophrenia. Bipolar Disord. 2015;17:160–171. doi: 10.1111/bdi.12228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  210. Pedrosa E, Shah A, Tenore C, Capogna M, Villa C, Guo X, Zheng D, Lachman HM. beta-catenin promoter ChIP-chip reveals potential schizophrenia and bipolar disorder gene network. J Neurogenet. 2010;24:182–193. doi: 10.3109/01677063.2010.495182. [DOI] [PubMed] [Google Scholar]
  211. Peet M, Peters S. Drug-induced mania. Drug Saf. 1995;12:146–153. doi: 10.2165/00002018-199512020-00007. [DOI] [PubMed] [Google Scholar]
  212. Pena CJ, Bagot RC, Labonte B, Nestler EJ. Epigenetic signaling in psychiatric disorders. J Mol Biol. 2014;426:3389–3412. doi: 10.1016/j.jmb.2014.03.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  213. Perry W, Minassian A, Feifel D, Braff DL. Sensorimotor gating deficits in bipolar disorder patients with acute psychotic mania. Biol Psychiatry. 2001;50:418–424. doi: 10.1016/s0006-3223(01)01184-2. [DOI] [PubMed] [Google Scholar]
  214. Perry W, Minassian A, Henry B, Kincaid M, Young JW, Geyer MA. Quantifying over-activity in bipolar and schizophrenia patients in a human open field paradigm. Psychiatry Res. 2010;178:84–91. doi: 10.1016/j.psychres.2010.04.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  215. Perry W, Minassian A, Paulus MP, Young JW, Kincaid MJ, Ferguson EJ, Henry BL, Zhuang X, Masten VL, Sharp RF, Geyer MA. A reverse-translational study of dysfunctional exploration in psychiatric disorders: from mice to men. Arch Gen Psychiatry. 2009;66:1072–1080. doi: 10.1001/archgenpsychiatry.2009.58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  216. Pinsonneault JK, Han DD, Burdick KE, Kataki M, Bertolino A, Malhotra AK, Gu HH, Sadee W. Dopamine transporter gene variant affecting expression in human brain is associated with bipolar disorder. Neuropsychopharmacology. 2011;36:1644–1655. doi: 10.1038/npp.2011.45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  217. Poitou P, Boulu R, Bohuon C. Effect of lithium and other drugs on the amphetamine chlordiazeposice hyperactivity in mice. Experientia. 1975;31:99–101. doi: 10.1007/BF01924701. [DOI] [PubMed] [Google Scholar]
  218. Polter A, Beurel E, Yang S, Garner R, Song L, Miller CA, Sweatt JD, McMahon L, Bartolucci AA, Li X, Jope RS. Deficiency in the inhibitory serine-phosphorylation of glycogen synthase kinase-3 increases sensitivity to mood disturbances. Neuropsychopharmacology. 2010;35:1761–1774. doi: 10.1038/npp.2010.43. [DOI] [PMC free article] [PubMed] [Google Scholar]
  219. Polyakova M, Stuke K, Schuemberg K, Mueller K, Schoenknecht P, Schroeter ML. BDNF as a biomarker for successful treatment of mood disorders: a systematic & quantitative meta-analysis. J Affect Disord. 2015;174:432–440. doi: 10.1016/j.jad.2014.11.044. [DOI] [PubMed] [Google Scholar]
  220. Post RM. Transduction of psychosocial stress into the neurobiology of recurrent affective disorder. Am J Psychiatry. 1992;149:999–1010. doi: 10.1176/ajp.149.8.999. [DOI] [PubMed] [Google Scholar]
  221. Postolache TT, Oren DA. Circadian phase shifting, alerting, and antidepressant effects of bright light treatment. Clin Sports Med. 2005;24:381–413. xii. doi: 10.1016/j.csm.2004.12.005. [DOI] [PubMed] [Google Scholar]
  222. Powell SB, Young JW, Ong JC, Caron MG, Geyer MA. Atypical antipsychotics clozapine and quetiapine attenuate prepulse inhibition deficits in dopamine transporter knockout mice. Behav Pharmacol. 2008;19:562–565. doi: 10.1097/FBP.0b013e32830dc110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  223. Prasad V, Sheard MH. Effect of lithium upon desipramine enhanced shock-elicited fighting in rats. Pharmacol Biochem Behav. 1982;17:377–378. doi: 10.1016/0091-3057(82)90097-1. [DOI] [PubMed] [Google Scholar]
  224. Prickaerts J, Moechars D, Cryns K, Lenaerts I, van Craenendonck H, Goris I, Daneels G, Bouwknecht JA, Steckler T. Transgenic mice overexpressing glycogen synthase kinase 3beta: a putative model of hyperactivity and mania. The Journal of neuroscience: the official journal of the Society for Neuroscience. 2006;26:9022–9029. doi: 10.1523/JNEUROSCI.5216-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  225. Ralph-Williams RJ, Paulus MP, Zhuang X, Hen R, Geyer MA. Valproate attenuates hyperactive and perseverative behaviors in mutant mice with a dysregulated dopamine system. Biol Psychiatry. 2003;53:352–359. doi: 10.1016/s0006-3223(02)01489-0. [DOI] [PubMed] [Google Scholar]
  226. Redrobe JP, Nielsen AN. Effects of neuronal Kv7 potassium channel activators on hyperactivity in a rodent model of mania. Behav Brain Res. 2009;198:481–485. doi: 10.1016/j.bbr.2008.12.027. [DOI] [PubMed] [Google Scholar]
  227. Reinhart V, Bove SE, Volfson D, Lewis DA, Kleiman RJ, Lanz TA. Evaluation of TrkB and BDNF transcripts in prefrontal cortex, hippocampus, and striatum from subjects with schizophrenia, bipolar disorder, and major depressive disorder. Neurobiol Dis. 2015;77:220–227. doi: 10.1016/j.nbd.2015.03.011. [DOI] [PubMed] [Google Scholar]
  228. Riegel RE, Valvassori SS, Elias G, Reus GZ, Steckert AV, de Souza B, Petronilho F, Gavioli EC, Dal-Pizzol F, Quevedo J. Animal model of mania induced by ouabain: Evidence of oxidative stress in submitochondrial particles of the rat brain. Neurochem Int. 2009;55:491–495. doi: 10.1016/j.neuint.2009.05.003. [DOI] [PubMed] [Google Scholar]
  229. Ripke S, et al. Genome-wide association study identifies five new schizophrenia loci. Nat Genet. 2011;43:969–976. doi: 10.1038/ng.940. [DOI] [PMC free article] [PubMed] [Google Scholar]
  230. Ripperger JA, Shearman LP, Reppert SM, Schibler U. CLOCK, an essential pacemaker component, controls expression of the circadian transcription factor DBP. Genes Dev. 2000;14:679–689. [PMC free article] [PubMed] [Google Scholar]
  231. Risbrough V, Ji B, Hauger R, Zhou X. Generation and characterization of humanized mice carrying COMT158 Met/Val alleles. Neuropsychopharmacology. 2014;39:1823–1832. doi: 10.1038/npp.2014.29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  232. Rivalan M, Ahmed SH, Dellu-Hagedorn F. Risk-prone individuals prefer the wrong options on a rat version of the Iowa Gambling Task. Biol Psychiatry. 2009;66:743–749. doi: 10.1016/j.biopsych.2009.04.008. [DOI] [PubMed] [Google Scholar]
  233. Rock P, Goodwin G, Harmer C, Wulff K. Daily rest-activity patterns in the bipolar phenotype: A controlled actigraphy study. Chronobiol Int. 2014;31:290–296. doi: 10.3109/07420528.2013.843542. [DOI] [PMC free article] [PubMed] [Google Scholar]
  234. Rose AM, Mellett BJ, Valdes R, Jr, Kleinman JE, Herman MM, Li R, el-Mallakh RS. Alpha 2 isoform of the Na, K-adenosine triphosphatase is reduced in temporal cortex of bipolar individuals. Biol Psychiatry. 1998;44:892–897. doi: 10.1016/s0006-3223(97)00440-x. [DOI] [PubMed] [Google Scholar]
  235. Roussos P, Katsel P, Davis KL, Bitsios P, Giakoumaki SG, Jogia J, Rozsnyai K, Collier D, Frangou S, Siever LJ, Haroutunian V. Molecular and genetic evidence for abnormalities in the nodes of Ranvier in schizophrenia. Arch Gen Psychiatry. 2012;69:7–15. doi: 10.1001/archgenpsychiatry.2011.110. [DOI] [PubMed] [Google Scholar]
  236. Roybal K, Theobold D, Graham A, DiNieri JA, Russo SJ, Krishnan V, Chakravarty S, Peevey J, Oehrlein N, Birnbaum S, Vitaterna MH, Orsulak P, Takahashi JS, Nestler EJ, Carlezon WA, Jr, McClung CA. Mania-like behavior induced by disruption of CLOCK. Proc Natl Acad Sci U S A. 2007;104:6406–6411. doi: 10.1073/pnas.0609625104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  237. Sachs GS, Gardner-Schuster EE. Adjunctive treatment of acute mania: a clinical overview. Acta Psychiatr Scand Suppl. 2007:27–34. doi: 10.1111/j.1600-0447.2007.01056.x. [DOI] [PubMed] [Google Scholar]
  238. Sachs GS, Nierenberg AA, Calabrese JR, Marangell LB, Wisniewski SR, Gyulai L, Friedman ES, Bowden CL, Fossey MD, Ostacher MJ, Ketter TA, Patel J, Hauser P, Rapport D, Martinez JM, Allen MH, Miklowitz DJ, Otto MW, Dennehy EB, Thase ME. Effectiveness of adjunctive antidepressant treatment for bipolar depression. N Engl J Med. 2007;356:1711–1722. doi: 10.1056/NEJMoa064135. [DOI] [PubMed] [Google Scholar]
  239. Sarnthein J, Petsche H, Rappelsberger P, Shaw GL, von Stein A. Synchronization between prefrontal and posterior association cortex during human working memory. Proc Natl Acad Sci U S A. 1998;95:7092–7096. doi: 10.1073/pnas.95.12.7092. [DOI] [PMC free article] [PubMed] [Google Scholar]
  240. Saul MC, Gessay GM, Gammie SC. A new mouse model for mania shares genetic correlates with human bipolar disorder. PLoS One. 2012;7:e38128. doi: 10.1371/journal.pone.0038128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  241. Saul MC, Stevenson SA, Gammie SC. Sexually dimorphic, developmental, and chronobiological behavioral profiles of a mouse mania model. PLoS One. 2013;8:e72125. doi: 10.1371/journal.pone.0072125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  242. Scotti MA, Lee G, Stevenson SA, Ostromecki AM, Wied TJ, Kula DJ, Gessay GM, Gammie SC. Behavioral and pharmacological assessment of a potential new mouse model for mania. Physiol Behav. 2011;103:376–383. doi: 10.1016/j.physbeh.2011.03.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  243. Serretti A, Benedetti F, Mandelli L, Lorenzi C, Pirovano A, Colombo C, Smeraldi E. Genetic dissection of psychopathological symptoms: insomnia in mood disorders and CLOCK gene polymorphism. American journal of medical genetics Part B, Neuropsychiatric genetics: the official publication of the International Society of Psychiatric Genetics. 2003;121B:35–38. doi: 10.1002/ajmg.b.20053. [DOI] [PubMed] [Google Scholar]
  244. Shaldubina A, Einat H, Szechtman H, Shimon H, Belmaker RH. Preliminary evaluation of oral anticonvulsant treatment in the quinpirole model of bipolar disorder. J Neural Transm. 2002;109:433–440. doi: 10.1007/s007020200035. [DOI] [PubMed] [Google Scholar]
  245. Shaltiel G, Maeng S, Malkesman O, Pearson B, Schloesser RJ, Tragon T, Rogawski M, Gasior M, Luckenbaugh D, Chen G, Manji HK. Evidence for the involvement of the kainate receptor subunit GluR6 (GRIK2) in mediating behavioral displays related to behavioral symptoms of mania. Mol Psychiatry. 2008;13:858–872. doi: 10.1038/mp.2008.20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  246. Shinozaki G, Potash JB. New developments in the genetics of bipolar disorder. Curr Psychiatry Rep. 2014;16:493. doi: 10.1007/s11920-014-0493-5. [DOI] [PubMed] [Google Scholar]
  247. Sidor MM, Macqueen GM. Antidepressants for the acute treatment of bipolar depression: a systematic review and meta-analysis. J Clin Psychiatry. 2011;72:156–167. doi: 10.4088/JCP.09r05385gre. [DOI] [PubMed] [Google Scholar]
  248. Sidor MM, Spencer SM, Dzirasa K, Parekh PK, Tye KM, Warden MR, Arey RN, Enwright JF, 3rd, Jacobsen JP, Kumar S, Remillard EM, Caron MG, Deisseroth K, McClung CA. Daytime spikes in dopaminergic activity drive rapid mood-cycling in mice. Mol Psychiatry. 2015:5. doi: 10.1038/mp.2015.8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  249. Silverstone PH, Pukhovsky A, Rotzinger S. Lithium does not attenuate the effects of D-amphetamine in healthy volunteers. Psychiatry Res. 1998;79:219–226. doi: 10.1016/s0165-1781(98)00037-7. [DOI] [PubMed] [Google Scholar]
  250. Sintoni S, Kurtys E, Scandaglia M, Contestabile A, Monti B. Chronic valproic acid administration impairs contextual memory and dysregulates hippocampal GSK-3beta in rats. Pharmacol Biochem Behav. 2013;106:8–15. doi: 10.1016/j.pbb.2013.02.013. [DOI] [PubMed] [Google Scholar]
  251. Sjoholm LK, Kovanen L, Saarikoski ST, Schalling M, Lavebratt C, Partonen T. CLOCK is suggested to associate with comorbid alcohol use and depressive disorders. Journal of circadian rhythms. 2010;8:1. doi: 10.1186/1740-3391-8-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  252. Snead OC, 3rd, Clapcote SJ, Roder JC, Boulianne GL. Novel strategies for the development of animal models of refractory epilepsy. Adv Neurol. 2006;97:155–171. [PubMed] [Google Scholar]
  253. St Onge JR, Ahn S, Phillips AG, Floresco SB. Dynamic fluctuations in dopamine efflux in the prefrontal cortex and nucleus accumbens during risk-based decision making. J Neurosci. 2012;32:16880–16891. doi: 10.1523/JNEUROSCI.3807-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  254. Steullet P, Cabungcal JH, Kulak A, Kraftsik R, Chen Y, Dalton TP, Cuenod M, Do KQ. Redox dysregulation affects the ventral but not dorsal hippocampus: impairment of parvalbumin neurons, gamma oscillations, and related behaviors. J Neurosci. 2010;30:2547–2558. doi: 10.1523/JNEUROSCI.3857-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  255. Steullet P, Neijt HC, Cuenod M, Do KQ. Synaptic plasticity impairment and hypofunction of NMDA receptors induced by glutathione deficit: relevance to schizophrenia. Neuroscience. 2006;137:807–819. doi: 10.1016/j.neuroscience.2005.10.014. [DOI] [PubMed] [Google Scholar]
  256. Strakowski SM, Delbello MP, Adler CM. The functional neuroanatomy of bipolar disorder: a review of neuroimaging findings. Mol Psychiatry. 2005;10:105–116. doi: 10.1038/sj.mp.4001585. [DOI] [PubMed] [Google Scholar]
  257. Sudhof TC, Malenka RC. Understanding synapses: past, present, and future. Neuron. 2008;60:469–476. doi: 10.1016/j.neuron.2008.10.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  258. Svenson KL, Gatti DM, Valdar W, Welsh CE, Cheng R, Chesler EJ, Palmer AA, McMillan L, Churchill GA. High-resolution genetic mapping using the Mouse Diversity outbred population. Genetics. 2012;190:437–447. doi: 10.1534/genetics.111.132597. [DOI] [PMC free article] [PubMed] [Google Scholar]
  259. Sylvia LG, Dupuy JM, Ostacher MJ, Cowperthwait CM, Hay AC, Sachs GS, Nierenberg AA, Perlis RH. Sleep disturbance in euthymic bipolar patients. J Psychopharmacol. 2012;26:1108–1112. doi: 10.1177/0269881111421973. [DOI] [PMC free article] [PubMed] [Google Scholar]
  260. Szabo ST, Machado-Vieira R, Yuan P, Wang Y, Wei Y, Falke C, Cirelli C, Tononi G, Manji HK, Du J. Glutamate receptors as targets of protein kinase C in the pathophysiology and treatment of animal models of mania. Neuropharmacology. 2009;56:47–55. doi: 10.1016/j.neuropharm.2008.08.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  261. Tang H, Shen N, Jin H, Liu D, Miao X, Zhu LQ. GSK-3beta polymorphism discriminates bipolar disorder and schizophrenia: a systematic meta-analysis. Mol Neurobiol. 2013;48:404–411. doi: 10.1007/s12035-013-8414-x. [DOI] [PubMed] [Google Scholar]
  262. Threadgill DW, Churchill GA. Ten years of the Collaborative Cross. Genetics. 2012;190:291–294. doi: 10.1534/genetics.111.138032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  263. Tochigi M, Iwamoto K, Bundo M, Sasaki T, Kato N, Kato T. Gene expression profiling of major depression and suicide in the prefrontal cortex of postmortem brains. Neurosci Res. 2008;60:184–191. doi: 10.1016/j.neures.2007.10.010. [DOI] [PubMed] [Google Scholar]
  264. Torres IJ, DeFreitas CM, DeFreitas VG, Bond DJ, Kunz M, Honer WG, Lam RW, Yatham LN. Relationship between cognitive functioning and 6-month clinical and functional outcome in patients with first manic episode bipolar I disorder. Psychol Med. 2011;41:971–982. doi: 10.1017/S0033291710001613. [DOI] [PubMed] [Google Scholar]
  265. Trivedi MS, Deth R. Redox-based epigenetic status in drug addiction: a potential contributor to gene priming and a mechanistic rationale for metabolic intervention. Front Neurosci. 2014;8:444. doi: 10.3389/fnins.2014.00444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  266. Tunca Z, Ozerdem A, Ceylan D, Yalcin Y, Can G, Resmi H, Akan P, Ergor G, Aydemir O, Cengisiz C, Kerim D. Alterations in BDNF (brain derived neurotrophic factor) and GDNF (glial cell line-derived neurotrophic factor) serum levels in bipolar disorder: The role of lithium. J Affect Disord. 2014;166:193–200. doi: 10.1016/j.jad.2014.05.012. [DOI] [PubMed] [Google Scholar]
  267. Valvassori SS, Resende WR, Lopes-Borges J, Mariot E, Dal-Pont GC, Vitto MF, Luz G, de Souza CT, Quevedo J. Effects of mood stabilizers on oxidative stress-induced cell death signaling pathways in the brains of rats subjected to the ouabain-induced animal model of mania: Mood stabilizers exert protective effects against ouabain-induced activation of the cell death pathway. J Psychiatr Res. 2015;65:63–70. doi: 10.1016/j.jpsychires.2015.04.009. [DOI] [PubMed] [Google Scholar]
  268. van Enkhuizen J, Geyer MA, Halberstadt AL, Zhuang X, Young JW. Dopamine depletion attenuates some behavioral abnormalities in a hyperdopaminergic mouse model of bipolar disorder. J Affect Disord. 2014a;155:247–254. doi: 10.1016/j.jad.2013.08.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  269. van Enkhuizen J, Geyer MA, Young JW. Differential effects of dopamine transporter inhibitors in the rodent Iowa gambling task: relevance to mania. Psychopharmacology (Berl) 2013a;225:661–674. doi: 10.1007/s00213-012-2854-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  270. van Enkhuizen J, Henry BL, Minassian A, Perry W, Milienne-Petiot M, Higa KK, Geyer MA, Young JW. Reduced dopamine transporter functioning induces high-reward risk-preference consistent with bipolar disorder. Neuropsychopharmacology. 2014b;39:3112–3122. doi: 10.1038/npp.2014.170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  271. van Enkhuizen J, Minassian A, Young JW. Further evidence for ClockDelta19 mice as a model for bipolar disorder mania using cross-species tests of exploration and sensorimotor gating. Behav Brain Res. 2013b;249:44–54. doi: 10.1016/j.bbr.2013.04.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  272. Vassout A, Veenstra S, Hauser K, Ofner S, Brugger F, Schilling W, Gentsch C. NKP608: a selective NK-1 receptor antagonist with anxiolytic-like effects in the social interaction and social exploration test in rats. Regul Pept. 2000;96:7–16. doi: 10.1016/s0167-0115(00)00194-4. [DOI] [PubMed] [Google Scholar]
  273. Vasudev K, Macritchie K, Geddes J, Watson S, Young A. Topiramate for acute affective episodes in bipolar disorder. Cochrane Database Syst Rev. 2006:CD003384. doi: 10.1002/14651858.CD003384.pub2. [DOI] [PubMed] [Google Scholar]
  274. Vitaterna MH, King DP, Chang AM, Kornhauser JM, Lowrey PL, McDonald JD, Dove WF, Pinto LH, Turek FW, Takahashi JS. Mutagenesis and mapping of a mouse gene, Clock, essential for circadian behavior. Science. 1994;264:719–725. doi: 10.1126/science.8171325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  275. Vitaterna MH, Ko CH, Chang AM, Buhr ED, Fruechte EM, Schook A, Antoch MP, Turek FW, Takahashi JS. The mouse Clock mutation reduces circadian pacemaker amplitude and enhances efficacy of resetting stimuli and phase-response curve amplitude. Proc Natl Acad Sci U S A. 2006;103:9327–9332. doi: 10.1073/pnas.0603601103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  276. Vucurovic K, Landais E, Delahaigue C, Eutrope J, Schneider A, Leroy C, Kabbaj H, Motte J, Gaillard D, Rolland AC, Doco-Fenzy M. Bipolar affective disorder and early dementia onset in a male patient with SHANK3 deletion. Eur J Med Genet. 2012;55:625–629. doi: 10.1016/j.ejmg.2012.07.009. [DOI] [PubMed] [Google Scholar]
  277. Wang B, Chen D. Evidence for seasonal mania: a review. J Psychiatr Pract. 2013;19:301–308. doi: 10.1097/01.pra.0000432600.32384.c5. [DOI] [PubMed] [Google Scholar]
  278. Whittington MA, Traub RD, Kopell N, Ermentrout B, Buhl EH. Inhibition-based rhythms: experimental and mathematical observations on network dynamics. Int J Psychophysiol. 2000;38:315–336. doi: 10.1016/s0167-8760(00)00173-2. [DOI] [PubMed] [Google Scholar]
  279. Wirz-Justice A. Biological rhythm disturbances in mood disorders. Int Clin Psychopharmacol. 2006;21(Suppl 1):S11–15. doi: 10.1097/01.yic.0000195660.37267.cf. [DOI] [PubMed] [Google Scholar]
  280. Wu JC, Kelsoe JR, Schachat C, Bunney BG, DeModena A, Golshan S, Gillin JC, Potkin SG, Bunney WE. Rapid and sustained antidepressant response with sleep deprivation and chronotherapy in bipolar disorder. Biol Psychiatry. 2009;66:298–301. doi: 10.1016/j.biopsych.2009.02.018. [DOI] [PubMed] [Google Scholar]
  281. Yang Y, Dieter MZ, Chen Y, Shertzer HG, Nebert DW, Dalton TP. Initial characterization of the glutamate-cysteine ligase modifier subunit Gclm(−/−) knockout mouse. Novel model system for a severely compromised oxidative stress response. J Biol Chem. 2002;277:49446–49452. doi: 10.1074/jbc.M209372200. [DOI] [PubMed] [Google Scholar]
  282. Yechiam E, Hayden EP, Bodkins M, O’Donnell BF, Hetrick WP. Decision making in bipolar disorder: a cognitive modeling approach. Psychiatry Res. 2008;161:142–152. doi: 10.1016/j.psychres.2007.07.001. [DOI] [PubMed] [Google Scholar]
  283. Yost C, Torres M, Miller JR, Huang E, Kimelman D, Moon RT. The axis-inducing activity, stability, and subcellular distribution of beta-catenin is regulated in Xenopus embryos by glycogen synthase kinase 3. Genes Dev. 1996;10:1443–1454. doi: 10.1101/gad.10.12.1443. [DOI] [PubMed] [Google Scholar]
  284. Young JW, Dulcis D. Investigating the mechanism(s) underlying switching between states in bipolar disorder. Eur J Pharmacol. 2015 doi: 10.1016/j.ejphar.2015.03.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  285. Young JW, Goey AK, Minassian A, Perry W, Paulus MP, Geyer MA. The mania-like exploratory profile in genetic dopamine transporter mouse models is diminished in a familiar environment and reinstated by subthreshold psychostimulant administration. Pharmacol Biochem Behav. 2010;96:7–15. doi: 10.1016/j.pbb.2010.03.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  286. Young JW, Henry BL, Geyer MA. Predictive animal models of mania: hits, misses and future directions. Br J Pharmacol. 2011a;164:1263–1284. doi: 10.1111/j.1476-5381.2011.01318.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  287. Young JW, Minassian A, Paulus MP, Geyer MA, Perry W. A reverse-translational approach to bipolar disorder: rodent and human studies in the Behavioral Pattern Monitor. Neurosci Biobehav Rev. 2007;31:882–896. doi: 10.1016/j.neubiorev.2007.05.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  288. Young JW, van Enkhuizen J, Winstanley CA, Geyer MA. Increased risk-taking behavior in dopamine transporter knockdown mice: further support for a mouse model of mania. J Psychopharmacol. 2011b;25:934–943. doi: 10.1177/0269881111400646. [DOI] [PMC free article] [PubMed] [Google Scholar]
  289. Zhou D, Lambert S, Malen PL, Carpenter S, Boland LM, Bennett V. AnkyrinG is required for clustering of voltage-gated Na channels at axon initial segments and for normal action potential firing. J Cell Biol. 1998;143:1295–1304. doi: 10.1083/jcb.143.5.1295. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES