Skip to main content
British Journal of Pharmacology logoLink to British Journal of Pharmacology
. 2013 Mar 12;168(7):1531–1554. doi: 10.1111/bph.12052

Antidepressant therapy in epilepsy: can treating the comorbidities affect the underlying disorder?

L Cardamone 1, MR Salzberg 1,2,3, TJ O'Brien 1, NC Jones 1
PMCID: PMC3605864  PMID: 23146067

Abstract

There is a high incidence of psychiatric comorbidity in people with epilepsy (PWE), particularly depression. The manifold adverse consequences of comorbid depression have been more clearly mapped in recent years. Accordingly, considerable efforts have been made to improve detection and diagnosis, with the result that many PWE are treated with antidepressant drugs, medications with the potential to influence both epilepsy and depression. Exposure to older generations of antidepressants (notably tricyclic antidepressants and bupropion) can increase seizure frequency. However, a growing body of evidence suggests that newer (‘second generation’) antidepressants, such as selective serotonin reuptake inhibitors or serotonin-noradrenaline reuptake inhibitors, have markedly less effect on excitability and may lead to improvements in epilepsy severity. Although a great deal is known about how antidepressants affect excitability on short time scales in experimental models, little is known about the effects of chronic antidepressant exposure on the underlying processes subsumed under the term ‘epileptogenesis’: the progressive neurobiological processes by which the non-epileptic brain changes so that it generates spontaneous, recurrent seizures. This paper reviews the literature concerning the influences of antidepressants in PWE and in animal models. The second section describes neurobiological mechanisms implicated in both antidepressant actions and in epileptogenesis, highlighting potential substrates that may mediate any effects of antidepressants on the development and progression of epilepsy. Although much indirect evidence suggests the overall clinical effects of antidepressants on epilepsy itself are beneficial, there are reasons for caution and the need for further research, discussed in the concluding section.

Keywords: epilepsy, antidepressants, depression, anxiety, SSRI, epileptogenesis

Introduction

Epilepsy is a common group of neurological disorders whose hallmark is unprovoked seizures (Engel and Pedley, 2007; Berg and Scheffer, 2011). Seizures can be distressing, harmful and even fatal, but an additional and no less significant consequence is the associated comorbidities of epilepsy, which contribute greatly to disability and impaired quality of life. While these can include physical and cognitive impairments as well as adverse effects of antiepileptic drugs (AEDs), it is the psychiatric disorders of several kinds that constitute a large proportion of the burden of comorbidity in epilepsy (Gaitatzis et al., 2004; Hesdorffer and Krishnamoorthy, 2011). When these psychopathologies, such as depression, are recognized, they are generally treated with antidepressant pharmacotherapy – drugs that have the potential to interfere not only with their target syndromes but also to impact the epilepsy. The mechanisms underlying this interaction will have major implications for the treatment of both epilepsy and comorbid psychiatric disorders.

This review will focus on depressive and anxiety syndromes in epilepsy, the two main sets of disorders for which antidepressants are prescribed (for convenience, the term ‘antidepressant’ will be employed, although these medications are also used for anxiety and other disorders). Before examining the question that is the focus of this paper, it is useful to set the scene by briefly outlining the definitions and varieties of epilepsy, depression and anxiety and the epidemiological scale of the problem. We will also discuss the natural history and consequences of comorbid anxiety and depression in epilepsy and current theories about their causation. The management of depressive and anxiety disorders in epilepsy is then discussed broadly, before focusing on antidepressant medications – their classes and modes of action.

A key issue structuring this paper is that antidepressants may affect epilepsy in two general ways: via short-term effects on excitability (e.g. seizure threshold, seizure frequency) and via effects on the longer-term neurobiological processes that underlie the epileptic state. First-generation antidepressants, notably tricyclic antidepressants (TCAs), were quickly recognized to have the capacity to trigger seizures in non-epileptic patients (Preskorn and Fast, 1992) (Wroblewski et al., 1990) and to aggravate preexisting epilepsy (Pisani et al., 1999). This was corroborated by in vivo (reviewed in (Trimble, 1978) and in vitro (Luchins et al., 1984) experimental electrophysiological studies, and consequently, clinicians became reluctant to prescribe any antidepressant to people with epilepsy (PWE) (Cotterman-Hart, 2010). For these and other reasons, many depressed PWE remained – and still remain – either undertreated or completely untreated for their mood disorder. However, a growing body of evidence suggests that newer (‘second generation’) antidepressants have markedly less effect on excitability and indeed may potentially lead to improvements in epilepsy severity. Considerable effort has been expended improving the detection, diagnosis and treatment of depression (and anxiety) in PWE. Some have proposed that the evidence is now sufficient to conduct clinical trials of the efficacy of antidepressants in improving epilepsy endpoints (Favale et al., 1995; 2003), as opposed to solely measuring depression and anxiety endpoints. Whilst the evidence to date is promising, there remain grounds for caution and a need for further, focused research.

Epilepsy: definition and prevalence

The formal definitions of ‘seizure’ and ‘epilepsy’ are complex and controversial (Berg et al., 2010); however, a seizure is generally defined as ‘a transient occurrence of signs and/or symptoms due to abnormal excessive or synchronous neuronal activity in the brain’ (Fisher et al., 2005); and epilepsy is considered ‘a group of neurologic conditions, the fundamental characteristics of which are recurrent, usually unprovoked, epileptic seizures’ (Engel and Pedley, 2008). An alternative definition of epilepsy, encompassing aspects of the associated comorbidities has been proposed: ‘a chronic condition of the brain characterized by an enduring propensity to generate epileptic seizures, and by the neurobiological, cognitive, psychological, and social consequences of this condition’ (Fisher et al., 2005). Current terminology classifies epilepsies as focal or generalized with genetic, structural/metabolic or unknown origins (Berg and Scheffer, 2011), with specific clinical manifestations of seizures determined by the brain structures affected by the abnormal neuronal firing pattern. Collectively, the epilepsies constitute a common group of disorders, with an incidence in developed countries of about 80 per 100 000 persons year−1 and a point prevalence of 1–10 cases per 1000 persons (Banerjee et al., 2009; Shorvon, 2010), with the highest incidences observed in children and the elderly.

Epilepsy and its comorbidities

Although recurrent unprovoked seizures are the hallmark of the disorder and can be distressing, harmful and even fatal, the associated comorbidities contribute greatly to disability and impaired quality of life. This can include somatic comorbidities (e.g. cardiac and respiratory disorders) (Gaitatzis et al., 2012), physical impairments such as injury (e.g. burns, fractures), adverse effects of AEDs and impaired fertility in addition to cognitive impairments. Mortality is elevated due to injury, suicide, sudden unexpected death in epilepsy (SUDEP) and underlying diseases that may also give rise to seizures and some epilepsies (e.g. brain tumours). A large proportion of the burden of comorbidity is related to psychiatric disorders (Gaitatzis et al., 2004; Hesdorffer and Krishnamoorthy, 2011; Ottman et al., 2011). The importance of these problems was highlighted by the National Institute of Health Epilepsy Research Benchmarks, which nominated the comorbidities of epilepsy, including prominently psychiatric comorbidities, as priority research areas (Kelley et al., 2009). An important development is depression screening of epilepsy patients. The Neurological Disorders Depression Inventory for Epilepsy (NDDI-E) is a six-item scale for use in epilepsy populations (Gilliam et al., 2006; Friedman et al., 2009). Validated in tertiary centres, its properties are as yet unknown in community based epilepsy samples with lower depression prevalence.

Descriptive epidemiology of depressive and anxiety disorders in epilepsy

The most common psychiatric comorbidities in epilepsy include depression, anxiety, attention-deficit hyperactivity disorder and psychoses (Gaitatzis et al., 2004; Hesdorffer and Krishnamoorthy, 2011). For anxiety and depression, rates of these disorders in most community and clinic-based studies are clearly elevated above the general population and often similar to or greater than rates for anxiety and depression associated with other chronic illnesses (for depression, see McLaughlin et al., 2008; Fuller-Thomson and Brennenstuhl, 2009; for anxiety disorders, see Mensah et al., 2007; Kanner, 2011a; for an omnibus study of psychiatric comorbidity, see Tellez-Zenteno et al., 2007). Depression is common in newly presenting patients with epilepsy (e.g. (Panelli et al., 2007; Velissaris et al., 2009) and indeed may precede epilepsy onset and constitute a risk factor for it (Hesdorffer et al., 2012, and see discussion below).

Depression can take several forms: major depressive disorder (MDD), dysthymic disorder, adjustment disorder and depressive phases of bipolar disorder. Debate continues about the validity of ‘interictal dysphoric disorder’, a depressive-like disorder suggested to be specific to epilepsy but argued by others to be a form of dysthymic disorder (Mula et al., 2010). In addition, depressive symptoms may occur before, during or after a seizure. Many studies have shown associations of comorbid depression with impaired quality of life, greater cognitive deficits and greater health care utilization (Lacey et al., 2009). Indeed, in some studies, depression is a greater predictor of impaired quality of life than epilepsy-related variables, such as illness duration or seizure frequency (Kanner, 2009; Kanner et al., 2010). Depression is a strong risk factor for suicide and a possible risk factor for SUDEP (Ridsdale et al., 2011). Compared with depression, anxiety disorders have until recently been somewhat neglected in epilepsy. However, they have shown to be highly prevalent, often more so than depression, and potentially linked to many of the same adverse consequences as depression (Kanner, 2011a). Anxiety disorder can take several forms, including generalized anxiety or panic, and probably constitute a risk factor for subsequent depression as they do in non-epileptic persons. Given their high prevalence and adverse consequences, management of depressive and anxiety symptoms and syndromes in PWE should be an essential part of epilepsy management and routinely available in services that treat PWE (Barry et al., 2008).

Causation of depressive and anxiety disorders in epilepsy

These affective disturbances tend to be considered either as understandable psychological reaction to the stresses and challenges of living with epilepsy; or as neurobiological epiphenomena of the epileptic brain (Salzberg, 2011). In addition, research interest has focused on two other ‘causal arrows’: not only may psychiatric disturbance result from epilepsy (via psychological and/or neurobiological pathways), but (i) a psychiatric disorder, notably depression, may also contribute to the causation of epilepsy, and (ii) shared causal factors (e.g. genes, traumatic brain injury, early life stress) may give rise to both psychiatric illness and epilepsy.

Much depression/anxiety research in the epilepsy field is ‘either/or’ – either psychosocial or neurobiological – and lacks a longitudinal perspective. Present-day general (non-epilepsy associated) depression research is biopsychosocial in nature, integrating insights from genetics, epidemiology, psychology and neurobiology, operating within a diathesis–stress framework (Hoppe and Elger, 2011), and adopts a lifespan perspective (Colman and Ataullahjan, 2010). The same pertains to anxiety disorders. The causation of depression and anxiety disorders is a multistage process with origins in early life; and, after onset of the disorder, there is a very high chance of recurrence throughout the lifespan (Colman and Ataullahjan, 2010). Recently, there have been calls for these perspectives to be adopted more fully in the epilepsy field (Hermann et al., 2008; Hermann and Jacoby, 2009).

Although some good-quality studies have shown associations between factors such as seizure frequency and rates of anxiety and depression, in general, neurological features of epilepsy have been somewhat inconsistent predictors of psychopathology: factors such as age of onset; type, frequency and severity of seizures; epilepsy syndrome (focal or generalized); anatomical location and laterality of focus have all been examined in many studies with often contradictory results (Adams et al., 2008; Filho et al., 2008; Asmussen et al., 2009; Babu et al., 2009; Desai et al., 2010) (for detailed reviews of this extensive and complex literature, see Hoppe and Elger, 2011; Lin et al., 2012). Duration of epilepsy is associated with severity of depression irrespective of epilepsy variables such as seizure type or frequency or EEG alterations (Robertson et al., 1987). The concept of a particular association of psychiatric comorbidity with limbic forms of epilepsy has progressively been eroded by evidence of elevated rates of such disorder in generalized and extra-temporal focal epilepsies (Adams et al., 2008). In part, the concept of a special link with temporal lobe epilepsy arose due to a preponderance of studies of patients in tertiary epilepsy centres, often those being assessed for epilepsy surgery. In contrast to neurological factors, psychosocial factors such as life stress, coping style, social support, perceived stigma and personality have been more consistent predictors (Hermann et al., 2000).

Neurobiologically oriented studies have been more informative, employing structural (Briellmann et al., 2007; Paparrigopoulos et al., 2008; Elst et al., 2009; Salgado et al., 2010; Finegersh et al., 2011; Labate et al., 2011) or functional imaging (Gilliam et al., 2007; Hasler et al., 2007; Bonelli et al., 2009; Assem-Hilger et al., 2010) or both (Richardson et al., 2007; Theodore et al., 2007; Lothe et al., 2008) and, where available, histopathological or molecular pathological study of excised temporal lobe tissue (Frisch et al., 2009). The most consistent findings have been in mesial temporal lobe epilepsy, with suggestions of enlarged amygdalae, diminished hippocampal and neocortical volumes, the latter in both temporal and extra-temporal cortex; diminished 5HT1A receptor binding in the hippocampus, and possibly raphe nuclei, insula and cingulate gyrus (Hasler et al., 2007); and a correlation between depression and degree of hippocampal abnormality on 1H-magnetic resonance spectroscopy imaging (Gilliam et al., 2007). Most such studies are cross-sectional, thus unable to determine causality, but the association of depression with diminished extra-temporal cortical thickness (Salgado et al., 2010) is important, as such thinning is found in both ordinary (non-epilepsy-associated) depression and in mesial temporal lobe epilepsy (Labate et al., 2011).

Antiepileptic drugs and psychiatric comorbidities

AEDs have a range of beneficial and adverse psychotropic effects and some are effective as treatments for mania and bipolar depression and as mood stabilizers in bipolar and schizoaffective disorder (Schmitz, 2011). Patients commonly believe that their AED treatment contributes to their mood-related symptomatology, however, establishing the causal role of any particular AED with regard to an affective symptom is complex, as it may be confounded by type of epilepsy, other co-occurring medication exposures, pre-existing psychopathology, cognitive deficits, the impact on seizure activity (e.g. a marked reduction in seizures may have either beneficial or adverse effects on mood) and pharmacokinetic effects. Also, the characterization of psychological symptoms is often poor: for example, many studies record ‘agitation’ and ‘irritability’, ill-defined symptoms that may stem from anxiety or other causes; similarly ‘apathy’ or ‘tiredness’ may stem from depression or other causes. However, AEDs thought to cause depressive symptoms and syndromes include carbamazepine, benzodiazepines (clobazam, clonazepam), lamotrigine, phenobarbital, phenytoin, pregabalin, primidone, tiagabine, topiramate, vigabatrin and zonisamide (Mula and Monaco, 2009a; Schmitz, 2011); while levetiracetam, lamotrigine, pregabalin and tiagabine have been linked to anxiety symptoms (Mula and Monaco, 2009a; Schmitz, 2011). Furthermore, a recent meta-analysis has linked AED use with increased suicide risk (FDA, 2008). Despite the aforementioned caveats for linking AED use to any psychopathology, and the methodological issues associated with the meta-analysis (see discussion in Fountoulakis et al., 2012), this report prompted the US Food and Drugs Administration to publish a warning encompassing all AEDs pertaining to this increased risk. This serves to highlight the seriousness and complexities of psychiatric problems associated with AED use.

A bidirectional relationship between epilepsy and psychiatric disorders

The ‘causal arrows’ considered so far run from epilepsy to psychopathology via various neurobiological pathways including AED exposure and psychosocial effects. However, several lines of evidence suggest a bidirectional relationship between epilepsy and psychiatric disorders and the possibility of shared factors giving rise to both epilepsy and psychiatric disorders (Kanner, 2011b).

First, several AEDs are efficacious for both epilepsy and mood disorders: valproate, carbamazepine and lamotrigine are effective mood stabilizers in bipolar and schizoaffective disorders, with some evidence that topiramate and levetiracetam may also have a role; and lamotrigine has efficacy in treating bipolar depression. Second, epidemiological evidence has emerged that depression is a risk factor for subsequent onset of epilepsy (Forsgren and Nystrom, 1990; Hesdorffer et al., 2000; Nilsson et al., 2003; Hesdorffer et al., 2006; Hesdorffer et al., 2012); and that maternal mood disorder is a risk factor for epilepsy in offspring (Morgan et al., 2012). Third, in most animal epilepsy models, affective disturbance is present, measured by standard tests of anxiety- and depressive-like behaviour. This applies to models of acquired epilepsy, such as electrical kindling (Adamec and Shallow, 2000; Kalynchuk, 2000; Post, 2002; Mazarati et al., 2007), post-status epilepticus (Groticke et al., 2007; 2008; Koh et al., 2007; Mazarati et al., 2008; Muller et al., 2009a,b) and febrile seizure (Mesquita et al., 2006) models; models of genetically determined epilepsy such as Genetically Epilepsy Prone Rats (GEPR) (Jobe and Browning, 2007), Genetic Absence Epilepsy Rats from Strasbourg (GAERS) (Jones et al., 2008b; 2010; Bouilleret et al., 2009), Wistar Albino Glaxo rats from Rijswijk (WAG-Rij) (Sarkisova and van Luijtelaar, 2011) and posttraumatic epilepsy models (Jones et al., 2008a) amongst others. Early life psychosocial exposures have also been shown to contribute to the development of epilepsy, for example maternal separation (Salzberg et al., 2007; Jones et al., 2009; Kumar et al., 2011) and cross-fostering (Gilby et al., 2009) models, as well as leading to alterations in cellular electrophysiology (Ali et al., 2011). Fourth, disturbances in several key neurotransmitter systems are implicated in epilepsies and depression, notably serotonergic, noradrenergic, GABAergic and glutamatergic systems (Jobe et al., 1999; Olsen et al., 1999; Cotter et al., 2002; Kanner and Balabanov, 2002; Jobe, 2003; Bagdy et al., 2007). Finally, in several studies, prior depression (or psychopathology more generally) is a predictor of treatment resistance, both to AEDs (Hitiris et al., 2007; Petrovski et al., 2010) or to epilepsy surgery (Kanner et al., 2009; Metternich et al., 2009), although a recent study of a temporal lobectomy series failed to find this association (Adams et al., 2012). Other commonalities in the neurobiology of epilepsy and depression/anxiety are discussed below, as they relate to the mechanisms of action of antidepressants.

Current treatment of anxiety and depression in epilepsy

For both depressive and anxiety disorders, the main effective treatments available at present are various psychotherapies, pharmacotherapies and their combination (other treatments employed for very severe or treatment-resistant illness will not be considered here, e.g. electroconvulsive therapy, psychosurgery, deep brain stimulation and transcranial magnetic stimulation). There is good evidence for the efficacy of psychotherapies in depression, with probable superiority to antidepressants in reducing relapse; there is also growing evidence specifically for depression in epilepsy (Ciechanowski et al., 2010; Thompson et al., 2010; Walker et al., 2010). Interestingly, vagal nerve stimulation has shown efficacy for both treatment-resistant epilepsy and treatment-resistant depression, suggesting shared neurobiological factors in their causation and/or treatment (Furmaga et al., 2012).

The various points at which antidepressant medication may be introduced during the course of epileptogenesis are shown in Figure 1. The principles of treatment selection are complex and beyond the scope of this review: many factors need to be taken into account such as the specific syndrome, past episodes of illness and past treatment response, physical comorbidities, other medications and risk of drug interactions, risk of suicide, suitability for psychotherapy, as well as cost and patient preferences and attitudes and beliefs regarding treatment options. However, an important principle is the considerable evidence for additive or synergistic effects of combined psychotherapy and medication for treatment of several disorders, both for initial treatment and prevention of relapse (Busch and Sandberg, 2012). This has been demonstrated recently in a mouse model (Karpova et al., 2011) where a combination of fear-extinction training and fluoxetine was effective in erasing conditioned fear, where each treatment separately was ineffective.

Figure 1.

Figure 1

The time course of epileptogenesis and the typical stages at which antidepressants may be clinically introduced. Neurobiological mechanisms involved in epileptogenesis are indicated in the grey box, a period that includes the time after the precipitating event to the latent period and the emergence of spontaneous, recurrent seizures. While this process is typical of epileptogenesis following an acquired injury, similar alterations following this time course can also occur with genetic epilepsies. At points before or during epileptogenesis, antidepressant medication may be initiated, indicated by the dashed vertical lines: green, prior to epilepsy onset; purple, the period immediately after a precipitating event; orange, the period around when seizures emerge; blue, during active epilepsy. At each of the points, antidepressants may impact on the different neurobiological alterations occurring in epileptogenesis, potentially influencing disease course. Figure modified from Scharfman (2007).

As the evidence base for anxiety and depression treatments in PWE is exceptionally small, treatment is largely informed by evidence from non-epilepsy patients (for reviews, including treatment guidelines see (Barry et al., 2008; Mula et al., 2008; Mula and Schmitz, 2009b; Noe et al., 2011; Perr and Ettinger, 2011; Kanner et al., 2012). Psychological interventions for depression and/or anxiety in PWE have been trialled in recent years (Ciechanowski et al., 2010; Walker et al., 2010; Macrodimitris et al., 2011), but combined therapies – psychotherapy with medication – have not yet been reported.

Antidepressants

The focus of this review will remain on the antidepressants most commonly prescribed to PWE suffering from depression and/or anxiety; principally selective serotonin reuptake inhibitors (SSRI), serotonin-noradrenaline reuptake inhibitors (SNRI) and related medications. These compounds are also useful for the treatment of other disorders, including chronic pain, sleep disorders, attention-deficit hyperactivity disorder and eating disorders. Table 1 summarizes the main antidepressants available for use in patients with depression and anxiety disorders, using the conventional classification based on synaptic actions. Whilst various other drugs, such as antipsychotics, ketamine and mood stabilizers (e.g. lithium, valproate, carbamazepine), have been shown to have antidepressant and sometimes anti-anxiety properties, they will not be discussed further, nor will the scores of novel medications at various stages of development (Stahl, 2008); the focus here is on established medications in common clinical usage.

Table 1.

Antidepressants in current clinical use

Drug class Examples in clinical use Proposed mechanism of action References to studies of use in patients or epilepsy models
Selective serotonin reuptake inhibitor (SSRIs) Citalopram, Escitalopram, Fluoxetine, Paroxetine, Sertraline Blockade of SERT, plus some with additional pharmacological actions Favale et al. (1995); Kanner et al. (2000); Specchio et al. (2004)
Serotonin and noradrenaline reuptake inhibitors (SNRIs) Venlafaxine, Desvenlafaxine, Duloxetine Selective SERT and NET blockade Santos et al. (2002); Ahern et al. (2006); Borowicz et al. (2011)
Tricyclics (TCAs) Imipramine, Desipramine, Doxepin, Dothiepin, Amoxapine Blockade of SERT and NET, also acting on histaminergic and cholinergic receptors. Negligible DAT affinity Dessain et al. (1986); Preskorn and Fast (1992)
Monoamine oxidase inhibitor (MAOIs), reversible Moclobemide Reversible inhibition of MAO-A (and others MAO-B also) Trimble (1978); Bonnet (2003); Krishnan (2007)
Monoamine oxidase inhibitor (MAOIs), irreversible Tranylcypromine, Phenelzine Irreversible inhibition of MAO-A and MAO-B Pisani et al. (2002)
Noradrenaline reuptake inhibitors (NRIs) Reboxetine Blockade of NET Kuhn et al. (2003)
Noradrenaline dopamine reuptake inhibitor (NDRI) Bupropion Not well defined, but shown to be a noradrenaline and dopamine reuptake inhibitor Settle et al. (1999) Mainie et al. (2001)
Serotonin-2 receptor antagonist and reuptake inhibitors (SARIs) Trazodone, Nefazodone Antagonist of 5-HT2A receptor and blockade of SERT Vanpee et al. (1999)
Noradrenaline and specific serotonergic antidepressant (NaSSA) Mirtazapine Blockade of 5-HT2 receptors, 5-HT3, H1 and α2 adrenergic receptor antagonist Kuhn et al. (2003)
Others Agomelatine Melatonergic (MT1 and MT2) receptor agonist, 5HT2C antagonist No reported use in PWE

DAT, dopamine transporter; NET, norepinephrine (noradrenaline) transporter; SERT, serotonin transporter.

Second-generation antidepressants: SSRIs and SNRIs

Due to better tolerability, with reduced side effects and relative safety in overdose compared with TCAs, SSRIs and SNRIs are the first-line drugs for the treatment of depression, especially in epilepsy. These drugs selectively inhibit serotonin reuptake at the neuronal presynaptic membrane by blocking the serotonin reuptake transporter, or in the case of SNRIs, the noradrenaline transporter as well, increasing serotonin and/or noradrenaline levels in the synapse. While being mostly selective in this action, there are also some effects on muscarinic and α-adrenergic receptors and on dopamine reuptake.

While considered safe for use in epilepsy, a dose-dependent increase in seizures has also been shown in overdose (Isbister et al., 2004). Furthermore, consideration of the pharmacokinetic interaction with AEDs is also important when prescribing these drugs to PWE. Some AEDs have been shown to increase clearance of antidepressants (e.g. carbamazepine), while some antidepressants can inhibit clearance of AEDs through interaction with the CYP-450 hepatic enzyme system (summarized in Barry et al., 2008).

Classifying antidepressants according to their synaptic targets is clinically useful, giving better understanding of the potential side effects and toxicity. However, direct activity at these pharmacological targets has long been recognized to not adequately explain their actions (discussed in Nestler et al., 2001). The acute increases in synaptic monoamine levels occurring as a consequence of reuptake inhibition are at odds with the timeframe of 2–3 weeks required for clinical antidepressant action and even longer for efficacy in anxiety disorders. Longer-term adaptations to antidepressants, mediated via second-messenger systems and entailing changes in gene expression and protein translation lead to downstream effects on neurogenesis and other forms of neuroplasticity, the neuroendocrine system [notably the hypothalamo–pituitary–adrenal (HPA) axis], other neurotransmitter systems and inflammatory pathways. Much evidence also implicates both neurotrophins, such as BDNF, and epigenetic mechanisms in these diverse actions. These various pathways and mechanisms become relevant when attempting to identify the mechanisms by which these compounds may also influence epilepsy and are discussed later. First, we summarize the human and experimental literature documenting the effects of antidepressants on seizure and epilepsy outcomes.

Studies of antidepressants in epilepsy: humans

Since the early reports of seizures as side effects of first generational antidepressants (such as TCAs – (Preskorn and Fast, 1992; Salzberg and Vajda, 2001), more recent reports have revisited this issue, focusing on newer generation drugs. These clinical studies have largely investigated effects of antidepressants in PWE on seizure frequency, but have not addressed whether antidepressants have influence on the processes associated with epileptogenesis. Studies investigating the effects of SSRI or SNRI treatment in PWE are summarized in Table 2. In these 10 studies, all patients received antidepressants adjunct to AED therapy, whether to treat the comorbid depressive symptoms or to assess the antiepileptic potential of the antidepressant. In all but one of the studies (see Kanner et al., 2000), patients were monitored for the occurrence of seizures in the months before and then during antidepressant treatment.

Table 2.

Effects of antidepressants in patients with epilepsy

Study group Antidepressants Follow up Seizure frequency in treatment group Reference
9 PWE FLX 3 months 4 Px increase in seizure frequency by >50%; 5 Px not increased Gigli et al. (1994)
36 PWE and MDD SRT, FLX 1 year 2 Px increase in seizure frequency; 34 Px not increased (number of Px with reduced seizures not mentioned) Thome-Souza et al. (2007)
100 PWE and depression or OCD SRT 1 year 6 Px increase in seizure frequency; 94 Px not increased (number of Px with reduced seizures not mentioned) Kanner et al. (2000)
39 PWE and depression CIT 4 months 2 Px had 50% increase in seizures; 37 Px with 50% decrease in seizures Specchio et al. (2004)
11 PWE CIT 8–10 months All Px showed improvements: 64.1% mean reduction in seizure frequency Favale et al. (2003)
17 PWE FLX 14 months Seizures disappeared in 6 Px; in others, seizure frequency reduced by 30% Favale et al. (1995)
43 PWE and HAMD score >15 CIT 2 months No change in seizure frequency Hovorka et al. (2000)
75 PWE and HAMD score >15 MIR, CIT, REB 7.5 months No change in seizure frequency Kuhn et al. (2003)
28 PWE FLV 29–444 days No change in seizure frequency Harmant et al. (1990)
121 PWE Various 1st and 2nd gen. drugs 1 year No change in seizure frequency compared to non-treated Px Okazaki et al. (2011)

Patient drop outs not included in sample sizes.

CIT, citalopram; FLV, fluvoxamine; FLX, fluoxetine; HAMD, hamilton depression rating scale; MDD, major depressive disorder; MIR, mirtazepine; OCD, obsessive compulsive disorder; PWE, people with epilepsy; Px, patients; REB, reboxetine; SRT, sertraline.

With long-term SSRI/SNRI antidepressant treatment (1 month to approximately 15 months), the overall group effects in these studies suggest that there is no worsening in seizure frequency (Harmant et al., 1990; Gigli et al., 1994; Hovorka et al., 2000; Kanner et al., 2000; Kuhn et al., 2003; Thome-Souza et al., 2007; Okazaki et al., 2011). Moreover, in several of the studies, improvements in seizure outcomes are seen, with some patients experiencing dramatic and complete seizure freedom during AD treatment (Favale et al., 1995; 2003; Specchio et al., 2004). It should be noted that some individuals in these studies experienced a worsening of seizure frequency (Gigli et al., 1994; Kanner et al., 2000; Specchio et al., 2004), but these were generally rare occurrences, and could be reversed by either increasing the AED medication, or removing the antidepressant. Nevertheless, this should be taken into account when considering these studies. It should also be noted that in most studies that assessed features of depression, these improved in a large proportion of the patients, which should also be considered as an important outcome.

As with many clinical studies, considerable limitations exist when interpreting these effects of antidepressants in PWE. These limitations are particularly pertinent when regarding the studies in which SSRIs either did not adversely affect seizures (Gigli et al., 1994), or reduced seizure frequency (Favale et al., 1995; 2003), which were very small studies, open-labelled, and contained short time periods of treatment and seizure analysis. Overall, there have been no double-blind, randomized controlled studies; most of the studies have been small and on highly selected patient populations from epilepsy clinics or following epilepsy surgery, and with few longitudinal, follow-up studies. Furthermore, assessment of antidepressants on seizure outcomes may be hindered by other factors, such as AED (or antidepressant) compliance, stress or insomnia. Also, all but one study assessed seizures before and after antidepressant treatment, which could be confounded by fluctuations in seizure frequency – common throughout the course of the illness. Case-control study designs should be employed, where a proportion of patients receive no active treatment (see Kanner et al., 2000).

Of note, an influential study by Alper et al. (2007) reviewed the effects of antidepressant drugs on seizure incidence in ∼75 000 non-epileptic patients in phase II and III clinical trials of depression treatment. The study found a significantly decreased incidence of seizures occurring in depressed patients treated with antidepressants compared to those treated with placebo (Alper et al., 2007). Despite the fact that patients with known epilepsy would have been excluded from such clinical trials, it is interesting to note that the rate of spontaneous seizures in the depressed patients in these trials was considerably above known population rates. Although not performed in PWE, this report strongly suggests an overall anti-seizure effect of antidepressants at therapeutic doses, consistent with the findings from studies in PWE (Table 2).

To date, there have been no clinical studies that have investigated whether antidepressant treatment has disease-modifying effects on the epilepsy itself (as opposed to anti-seizure effects). Undertaking such a study would be logistically difficult, requiring long-term treatment and follow-up, appropriately matched placebo-treated controls and the need to address various manifestation of epilepsy progression, such as seizure frequency, AED resistance, neuropsychiatric and neurocognitive deficits and structural brain changes. This can be more readily addressed in appropriate animal models, where long-term follow-up and assessment of epileptic outcomes, in addition to behavioural, structural and functional changes can be achieved in a shorter time frame and with greater control of experimental settings.

Studies of antidepressants in epilepsy: animal models

Both genetic and acquired models of epilepsy are valuable tools for assessing the biological processes and modulators of epilepsy development and progression. Many of these models accurately recapitulate the human disease and are appropriately used to assess drug response and efficacy on seizures and comorbidities. As mentioned above, a wide variety of animal models of epilepsy, both with acquired and genetic aetiology, also exhibit behaviours relevant to the psychiatric comorbidities present in PWE. The presence of the psychiatric comorbidities in both humans and in animal models of epilepsy suggests shared neurobiological mechanisms may be at play. However, how antidepressants affect these mechanisms and the disorders themselves is unclear and needs to be resolved.

There have been several studies investigating the effects of antidepressants on seizure occurrence in animal models. A recent review focusing on SSRIs concluded that, overall, treatment with this form of antidepressant exerts anticonvulsant effects in animal models of seizures/epilepsy (Igelstrom, 2012). Previous studies in animal models have shown that TCAs (Koella et al., 1979; Preskorn and Fast, 1992; Ago et al., 2006, 2007) increase seizure susceptibility, in line with the data from human studies showing a similar increase in seizure occurrence during TCA treatment (Preskorn and Fast, 1992). Similar pro-seizure effects have been reported for bupropion, a noradrenaline-dopamine reuptake inhibitor (Settle et al., 1999). The studies of all of the newer generation antidepressants on seizure outcomes in animal models of epilepsy are summarized in Table 3. Although there are seven studies reporting proconvulsant activity of SSRIs exclusively tested using acute seizure assessments, the majority of studies, which include models of both acute seizures and chronic epilepsy, indicate that overall, SSRIs and SNRIs exert either beneficial effects (25 studies) or are without influence (19 studies) on seizure outcomes. This suggests that the effects wrought by antidepressants on seizures are dependant upon the model used: these drugs are generally beneficial in the more clinically relevant models of chronic partial epilepsy, such as the post-status epilepticus, GEPRs, El mice and limbic kindling models, but can be detrimental in models of acute seizures, including pentylenetetrazol (PTZ)-, lidocaine- or fluorothyl-induced seizures.

Table 3.

Studies of antidepressant administration in animal models of seizures and epilepsy

Model Species Drug Dose (mg kg–1) Study protocol Seizure outcome measures Reference
Antidepressants were found to be proconvulsant
PTZ Mouse CIT 25, 50 Injection 30 min prior to sz test Behavioural threshold Payandemehr et al. (2012)*
PTZ Rat FLX 10 Injection 1 h prior to sz test Severity Zienowicz et al. (2005)
PTZ Rat FLX 10 Sz tested after 21 days admin. Behavioural threshold Ferrero et al. (2005)
PTZ Rat VEN 75–100 Injection 30 min prior to sz test Severity and latency Santos et al. (2002)
Lidocaine Mouse DES, MAP 5–20 Sz tested 2–5 after 5 days admin. Severity, threshold, frequency Arai et al. (2003)*
Lidocaine Mouse CIT 10 Injection 30 min prior to sz test Frequency Arai et al. (2003)*
Pilocarpine Mouse CIT, REB, BUP 15,30,20–40 Injection 30 min prior to sz test Behavioural threshold Vermoesen et al. (2011)*
Flurothyl Mouse SRT, BUP 40,40 Injection 30 min prior to sz test Latency Ahern et al. (2006)*
Flurothyl Mouse REB 20 Sz tested after 21 days admin. Latency Ahern et al. (2006)*
Antidepressants were found to be anticonvulsant
MES Mouse BUP 15–30 Injection 30 min prior to sz Behavioural threshold Tutka et al. (2004)*
PTZ Zebrafish CIT, REB, BUP 10–300 μM Injection 1 min prior to sz Behavioural threshold Vermoesen et al. (2011)*
PTZ Mouse REB, BUP 5–10, 10–40 Injection 30 min prior to sz test Behavioural threshold Vermoesen et al. (2011)*
PTZ Mouse FLX 20 Injection 60 min prior to sz test Survival Kecskemeti et al. (2005)*
PTZ Mouse FLX 1–20 Injection 30 min prior to sz test Behavioural threshold, latency Ugale et al. (2004)
PTZ Mouse CIT 0.5, 1 Injection 30 min prior to sz test Behavioural threshold Payandemehr et al. (2012)*
PTZ Rat VEN 25–50 Injection 30 min prior to sz test Severity and latency Santos et al. (2002)
Pilocarpine Zebrafish CIT 10–300 μM Injection 1 min prior to sz test Behavioural threshold Vermoesen et al. (2011)*
Pilocarpine Rat FLX 20 Once epileptic, drug admin for 5 days, sz assessed for 72 h SRS Hernandez et al. (2002)
Pilocarpine Rat FLX 20 Once epileptic, drug admin for 10 days, sz assessed for 7 days Behavioural threshold Mazarati et al. (2008)*
Flurothyl Mouse REB 20 Sz tested after 21 days admin. Latency Ahern et al. (2006)*
Focal pilocarpine Rat CIT 1 μM Sz tested over 4 h drug infusion Severity Clinckers et al. (2004)
Focal bicuculline Rat FLX 5–20 Injection 1 h prior to sz test Frequency, severity Prendiville and Gale (1993)
Focal bicuculline Rat FLX 1.75–7 nM Injection 15 min prior to sz test Behavioural threshold Pasini et al. (1992)
Picrotoxin Mouse FLX 20 Injection 65 min prior to sz test Latency Pericic et al. (2005)*
Picrotoxin Mouse FLX 20 Sz tested after 5 days admin. Latency Pericic et al. (2005)*
KA Mouse CIT 10 Sz tested after 7 days admin. Severity Jaako et al. (2011)
KA Rat CIT, REB 15, 20–30 Once epileptic, drug admin for 4 days and sz assessed Frequency, cumulative duration Vermoesen et al. (2012)*
MES Mouse VEN 12.5, 25 Injection 30 min prior to sz test Convulsive threshold Borowicz et al. (2011)*
MES Mouse VEN 12.5, 25 Sz tested after 14 days admin. Convulsive threshold Borowicz et al. (2011)*
MES Mouse FLX 10 Injection 60 min prior to sz test Hind limb extension threshold Raju et al. (1999)*
Ear shock Mouse FLX 15–25 Injection 30 min prior to sz test Hind limb extension threshold Borowicz et al. (2006)*
Hippocampal kindling Rat FLX 10 nmol Injection 15 min prior to sz test Electrical threshold Wada et al. (1993)
Hippocampal kindling Rat FLX 10 Sz tested after 21 days admin., 7 days wait, one injection Electrical threshold Wada et al. (1995)*
Amygdala kindling Cat FLX 2–10 Injection 1–49 h prior to sz test Electrical threshold Siegal and Murphy (1979)
GEPR-3 and GEPR-9 Rat FLX 30 Injection 1–5 h prior to sz test Severity Dailey et al. (1992)*
GEPR-3 and GEPR-9 Rat FLX 7–20 Sz tested every 7 days over 28 days admin. Behavioural threshold Dailey et al. (1992)*
GEPR-9 Rat FLX 15 Injection 1 h prior to sz test Severity and latency, behavioural Yan et al. (1994)
El mice Mouse CIT 0.01, 0.02, 0.04%# Sz tested after 14 days admin. Behavioural threshold Kabuto et al. (1994b)
El mice Mouse FLX 10 Sz tested after 3 or 7 days admin. Behavioural threshold Richman and Heinrichs (2007)
Antidepressants resulted in no change
PTZ, KA Mouse BUP 5–50 Injection 30 min prior to sz test Number of convulsions Tutka et al. (2004)*
PTZ, LH Rat FLX 10 Injection 1 h prior to sz test Behavioural threshold Ferrero et al. (2005)*
PTZ, pilocarpine Mouse CIT, REB, BUP various Injection 30 min prior to sz test Behavioural threshold Vermoesen et al. (2011)*
PTZ Mouse FLX 20 Injection 60 min prior to sz test Latency Kecskemeti et al. (2005)*
PTZ Rat FLX 2.5–20 Injection 30 min prior to sz test Latency and severity Ceyhan et al. (2005)
Lidocaine Mouse DES, MAP 20 Sz tested 5–9 after 5 days admin. Frequency Arai et al. (2003)*
Lidocaine Mouse CIT 10–20 Sz tested after 5 days admin. Frequency, threshold Arai et al. (2003)*
Pilocarpine Zebrafish REB, BUP 10–100 Injection 1 min prior to sz test Behavioural threshold Vermoesen et al. (2011)*
Pilocarpine Rat FLX 20 Once epileptic, drug admin for 10 days, sz assessed for 7 days Seizure frequency Mazarati et al. (2008)*
Flurothyl Mouse VEN, REB 20–40,20 Injection 30 min prior to sz test Latency Ahern et al. (2006)*
Flurothyl Mouse VEN, BUP, SRT 20–40,40,40 Sz tested after 21 days admin. Latency to first seizure Ahern et al. (2006)*
KA Mouse FLX 10–50 Injection 30 min prior to sz test Severity and duration Jin et al. (2009)
KA Rat CIT, REB 5–10, 10 Once epileptic, drug admin for 4 days and sz assessed Severity and duration Vermoesen et al. (2012)*
MES Mouse FLX 10 Sz tested after 21 days admin. Hind limb extension threshold Raju et al. (1999)*
Ear shock Mouse FLX 10 Injection 30 min prior to sz test Hind limb extension threshold Borowicz et al. (2006)*
Ear shock Mouse FLX 7.5–20 Sz tested after 14 days admin. Hind limb extension threshold Borowicz et al. (2007)
Hippocampal kindling Rat FLX 10 Injection 2 h prior to sz test Threshold and duration Wada et al. (1999)
Hippocampal kindling Rat FLX 1,10 Injection 1 h prior to sz test Electrical threshold Wada et al. (1995)*
EEG Rat FLV, CLV 10–30 Drugs infused and seizures assessed at 10 min intervals Epileptic activity on EEG Krijzer et al. (1984)
GEPR-3 Rat FLX 3.5, 7.2, 14.1 nmol Injection 15 min prior to sz test Severity and latency Statnick et al. (1996)
GEPR-9 Rat FLX 15 Injection 2 h prior to sz test Severity and latency Browning et al. (1997)
El mice Mouse CIT 80 Sz tested after 14 days admin. Behavioural threshold Kabuto et al. (1994a)
*

Study is repeated within the table; #dietary supplement; no difference after 4 weeks.

ADT, afterdischarge threshold; BUP, bupropion; CIT, citalopram; CLV, clovoxamine; DES, desipramine; FLV, fluvoxamine; FLX, fluoxetine; GEPR, genetically epilepsy prone rats; h, hour; KA, kainic acid; LH, learned helplessness; MAP, maprotiline; min, minute; PAR, paroxetine; SRS, spontaneous recurrent seizures (as recorded on EEG); sz, seizure; VEN, venlafaxine.

Although more labour-intensive, models of chronic epilepsy are the most appropriate to study effects of drugs. The post-status epilepticus model is one of the most widely used and well-validated chronic models of mTLE (Morimoto et al., 2004). Using this model, Hernandez et al. (2002) found that five days fluoxetine treatment inhibited spontaneous recurrent seizures following pilocarpine-induced status epilepticus (Hernandez et al., 2002), while in the same model, Mazarati et al. (2008) found no differences in spontaneous recurrent seizures following 10 days of fluoxetine treatment (Mazarati et al., 2008). In the post-kainic acid-induced status epilepticus model, Vermoesen et al. (2012) investigated the effects of 4 days of citalopram treatment, finding that citalopram reduced seizure frequency and cumulative seizure duration, without affecting seizure severity (Vermoesen et al., 2012). These strong studies suggest SSRIs are able to improve seizure outcomes when used in chronic epilepsy models, although the period of SSRI treatment, from 3–10 days, is relatively brief.

The strength of the collective data suggesting that antidepressants have no effect on seizures and may even be anticonvulsant is that the evidence is derived from a wide variety of models including both chemoconvulsant and electrically induced seizures, genetic models of epilepsy and also the use of different animal species. However, limitations to the majority of studies also exist: (i) many utilize acute seizure tests, such as the PTZ and maximal electroshock (MES) models, in otherwise non-epileptic animals. This is very different situation to the pathological circuits associated with truly epileptic brains, and so drugs may well have contrasting effects in such conditions. Only three studies have investigated long-term seizure outcomes in chronically epileptic rats following antidepressant treatment (Hernandez et al., 2002; Mazarati et al., 2008; Vermoesen et al., 2012) – see description above. (ii) Many of the models do not exhibit behavioural abnormalities indicative of psychopathology in humans, and so an examination of effects on depression-related outcomes is tenuous. (iii) There is large variety in experimental design relating to the dose and timing of injection of the antidepressant relative to seizure susceptibility testing, making collective interpretation difficult. (iv) In many studies, antidepressants were administered as a single dose, and this does not mimic the clinical situation where several weeks of treatment are required for the beneficial therapeutic effect of antidepressants to manifest in patients. Indeed, only about half of all the studies (16 of 36) reviewed in Table 3 incorporate a chronic treatment arm (defined as more than 2 days of ongoing treatment). Similarly absent in the experimental literature are any studies examining epileptogenesis. Future studies need to address the effects of chronic antidepressant treatment in chronic epilepsy models, incorporating measures of disease progression, as well as effects on seizure propensity. At the current state of play, the evidence suggests that seizure susceptibility can be influenced by antidepressant exposure. Next, we discuss several potential mechanisms that are candidates to mediate these effects.

Shared neurobiological substrates of antidepressant action and epileptogenesis

Understanding the mechanisms by which antidepressants influence epilepsy has not been thoroughly investigated, despite the broad literature on the neurobiological effects of antidepressants, and also of the pathophysiological processes associated with epileptogenesis. Analysing these two bodies of literature reveals several intermediary substrates that have been prominently linked to antidepressant activity and icto/epileptogenic mechanisms. The following section describes in detail these relevant intermediaries, the most prominent being neurogenesis and other forms of neuroplasticity; the HPA axis and glucocorticoids; monoamines and other neurotransmitters; BDNF; and inflammatory processes. Although these candidates are discussed separately, clearly there is an overlap – for example, the role of antidepressants on neurogenesis may be driven by BDNF signalling (Li et al., 2008), and glucocorticoids exert powerful effects on hippocampal neurogenesis (Mirescu and Gould, 2006) and inflammatory processes (Sorrells and Sapolsky, 2007). This list is by no means exhaustive, and other processes, including genetic/genomic and epigenetic mechanisms may equally be important. For instance, ketamine has recently shown promise as an effective antidepressant (Zarate et al., 2006), actions that may be linked to the mTOR pathway (Li et al., 2010), a pathway recently been implicated in a broad range of models of epileptogenesis (Wong, 2010).

Neurogenesis/neuroplasticity

The hippocampus retains significant capacity for neuroplasticity in adult life and is associated with many mood and neurological disorders (Small et al., 2011). Indeed, one of the key features of acquired epilepsy is aberrant hippocampal plasticity, which may result in the generation of hyperexcitable circuitry. Studies in both patients and animal models of temporal lobe epilepsy display a selective pattern of pyramidal cell loss in CA1 and CA3 regions (Corsellis, 1957; Sloviter, 1994a,b), gliosis (Niquet et al., 1994), axonal sprouting (which results in formation of recurrent synapses) (Tauck and Nadler, 1985; Mathern et al., 1998; Buckmaster et al., 2002; Kron et al., 2010) and aberrant migration of granule cells (McCloskey et al., 2006; Scharfman et al., 2007). The hippocampus is also frequently the site of seizure initiation (Babb, 2001). Aberrant hippocampal neuroplasticity is therefore considered to be a prevalent feature of acquired epilepsy and represents a strong candidate mechanism by which antidepressant drugs might interact with disease processes.

In recent years, one of the most widely studied forms of hippocampal neuroplasticity associated with epilepsy has been neurogenesis – the birth and integration of new neurons into the existing hippocampal circuitry. Neurogenesis persists into adulthood in discrete brain regions in both animals (Altman and Das, 1965) and humans (Eriksson et al., 1998), and alterations in neurogenesis are a key feature in both humans and animal models of epilepsy (Parent et al., 2006). Acute seizures stimulate the production of new neurons (Bengzon et al., 1997), whereas dramatically reduced rates in neuronal differentiation have been observed in chronically epileptic rats (Hattiangady et al., 2004; Hattiangady and Shetty, 2010). Following status epilepticus – an extended period of sustained seizures that can result in the development of epilepsy – newly generated neurons migrate ectopically into the dentate hilus region, creating aberrant dendritic connections which may contribute to the generation of epileptic circuitry (Parent et al., 1997; Scharfman et al., 2007; Kron et al., 2010). However, other reports suggest that some of these newly generated neurons play roles that would serve to protect the injured hippocampus from excessive excitability (Jakubs et al., 2006; Murphy et al., 2011), so it is not yet clear how the neurogenic response to seizures impact the development of epilepsy.

Antidepressants, particularly SSRIs, are also reported to increase hippocampal neurogenesis in both humans (Boldrini et al., 2009) and animals (Malberg et al., 2000; Santarelli et al., 2003). Indeed, early reports suggest that this property is critical for antidepressant activity in animal models (Santarelli et al., 2003), although this is still subject to debate (Miller et al., 2008; Zhao et al., 2008; Bessa et al., 2009). Exploration of the components of the ‘neuronal differentiation cascade’ suggests that antidepressants specifically increase the division of early neural progenitor cells, but not stem cells or neuroblasts (Encinas et al., 2006). Another intriguing property of SSRIs is their ability to reverse the maturation of a significant proportion of dentate granule cells into cells with immature neuronal properties (Kobayashi et al., 2010; Karpova et al., 2011), changes that could result in reduced mossy fibre synaptic facilitation from the dentate gyrus to CA3 region. This could have relevance for the spread of epileptic activity through the hippocampal circuitry. Together, these data describe significant effects of antidepressants on hippocampal neuroplasticity, which may be relevant for the reorganization of this structure during epileptogenesis.

Two studies to date have specifically examined neuroplasticity associated with chemoconvulsant-induced seizures during chronic treatment with the SSRI citalopram. These found that antidepressant treatment reduced levels of reactive gliosis and aberrant cell proliferation (Jaako et al., 2009) and reduced cell death and axonal sprouting (Jaako et al., 2011) induced by the seizures. This suggests that chronic exposure to SSRIs mitigates against the pathological hippocampal circuit remodelling induced by seizures, potentially preventing the onset of epilepsy which occurs after this insult (although this was not explored). However, much remains to be examined with respect to the influence of antidepressants on hippocampal neuroplasticity in the context of epilepsy/epileptogenesis. For instance, how does antidepressant exposure alter the migration, and the electrophysiological phenotype, of newly born neurons following brain insults? How do existing granule cells exposed to antidepressants respond when challenged with a brain injury? By increasing the number of newborn cells, are antidepressants accelerating or impeding the creation of hyperexcitable circuits following injury? Long-term studies determining the effect of antidepressants on neurogenesis and associated neuroplastic rearrangements in epilepsy are crucial to understand how these drugs may influence epileptogenesis.

Hypothalamo–pituitary–adrenal axis and glucocorticoids

Glucocorticoids are steroid hormones released in circadian and pulsatile rhythms, and also in times of stress. These hormones have many homeostatic functions in the body and also perform specialized actions in the brain related to the function of the amygdala and hippocampus (Joels and Baram, 2009), key structures implicated in both epilepsy and depression. Circulating and stress-induced levels of glucocorticoids are tightly regulated by the HPA axis, which releases glucocorticoids into the circulation. Negative and positive feedback loops incorporating the hippocampus and amygdala further regulate the activity of the HPA axis (Lupien et al., 2009). While HPA axis hyperactivity is one of the most consistent attributes in depressed patients (reviewed in (Pariante and Lightman, 2008), this system has also been shown to be dysfunctional in epilepsy: PWE (temporal and extra-temporal foci) without comorbid depression or anxiety had greater elevation in cortisol following the dexamethasone/corticotrophin-releasing hormone (Dex/CRH) test compared to controls, indicative of a hyperreactive HPA axis (Zobel et al., 2004). This is consistent with rat models of epilepsy (Mazarati et al., 2009). Elevations in cortisol and adrenocorticotrophic hormone (ACTH) have also been reported after seizures (Gallagher et al., 1984; Pritchard et al., 1985; Takeshita et al., 1986; Gallagher, 1987; Rao et al., 1989; Kumar et al., 2011). Animal studies provide insight on the cause/effect relationship between HPA axis dysfunction and epilepsy. Corticosterone supplementation enhances epileptogenesis in the amygdala kindling model (Karst et al., 1999; Taher et al., 2005; Kumar et al., 2007), and conversely, retarded kindling rates have been demonstrated in surgically adrenalectomized rats compared with sham rats (Cottrell et al., 1984; Weiss et al., 1993; Edwards et al., 1999). Additionally, early life stress, which is recognized to up-regulate HPA axis responsivity, imparts an enduring vulnerability to experimental limbic epileptogenesis (reviewed in Koe et al., 2009). Together, this evidence suggests a faulty regulation of the HPA axis in epilepsy, and implicates a modulatory role for this system in epilepsy pathophysiology.

As mentioned above, increases in glucocorticoids are observed in ∼50% of depressed patients (Holsboer, 2001), presumed to be due to dysfunctional negative feedback of the HPA axis. Interestingly, successful treatment in depressed patients with antidepressants, particularly SSRIs, can improve this impaired HPA axis regulation (Pariante, 2006; Himmerich et al., 2007). Animal studies also demonstrate suppression of HPA axis function following antidepressant treatment (Jensen et al., 1999; Jongsma et al., 2005), an effect that may be due to observed increases in glucocorticoid and mineralocorticoid receptors (Pepin et al., 1989; Reul et al., 1993; 1994; Barden, 1999).

The ability of antidepressants to influence HPA axis regulation and the evidence suggestive of a detrimental role of glucocorticoids in epilepsy development promotes the HPA axis as a potential site of interaction whereby antidepressants could influence epileptogenesis and seizure susceptibility. Although a large number of studies have investigated HPA axis changes in epilepsy, and following antidepressant treatment individually, to date no studies have investigated neuroendocrine function in epilepsy specifically during antidepressant exposure.

Monoamines and other neurotransmitters

A broad literature describes monoamine abnormalities in the pathophysiology of both depression and epilepsy. Notably, PET studies in PWE show reduced hippocampal 5HT1A receptor binding, reduced further with comorbid depression (Hasler et al., 2007) and an inverse correlation between 5HT1A receptor binding and severity of depressive symptoms (Theodore et al., 2007). Since many of the new generation antidepressants target monoaminergic neurotransmission, these systems present prime candidates to mediate the influence of antidepressants on seizures and epilepsy development. As described above, the pharmacological target of SSRIs and SNRIs are the serotonin and/or noradrenaline transporters, and inhibition of these targets increases synaptic levels of these neurotransmitters. It appears unlikely that this is the sole mechanism of antidepressant activity since these drugs usually need to be taken for several weeks to become clinically effective. However, the acute increases in synaptic neurotransmitter levels may set in train downstream mechanisms critical to the actions of these drugs, so it is important to consider these acute effects as potentially relevant to any chronically observed effects. A broad range of studies has demonstrated that increases in monoamines induce antiepileptic effects in seizure models (for reviews, see (Bagdy et al., 2007; Lu and Gean, 1998).

In addition to acute effects on seizure susceptibility, antidepressants could also affect epilepsy disease processes through monoaminergic mechanisms. In animals, lesioning noradrenergic inputs from the locus coeruleus accelerated amygdala kindling, implying an antiepileptogenic role for noradrenaline. Furthermore, the GEPR-9 rats display seizure susceptibility concurrent with noradrenergic deficits (Dailey and Jobe, 1986). Reversing these neurotransmitter deficits reduces the seizure susceptibility (Yan et al., 1993), implying a causal relationship. Other studies have investigated the effects on serotonergic (Wada et al., 1997) or noradrenergic agonists and antagonists (Shin and McIntyre, 2007) on seizures and epileptogenesis. For example, Wada et al. (1997) investigated disease progression using the kindling model following administration of serotonin receptor agonists and found that a 5HT1A receptor agonist, 8-hydroxy-2-(di-n-propylamino)tetralin (8-OH-DPAT) inhibited kindling rate while a 5HT2 receptor agonist, 2,5-dimethoxy-4-iodoamphetamine (DOI) facilitated kindling (Wada et al., 1997). The effect of DOI was blocked by the addition of a 5HT2 receptor antagonist, ketanserin. Activation of 5HT1A receptors may also be relevant to acute effects of SSRIs on seizures: 5HT1A receptor knockout mice demonstrate reduced seizure threshold (Sarnyai et al., 2000) and increased mortality resulting from a seizure (Parsons et al., 2001). Similarly, 5HT1A receptor antagonists have been shown to increase seizure severity and duration (Watanabe et al., 2000; Pericic et al., 2005), while agonists are protective against induced seizures (Hagan et al., 1995; Gariboldi et al., 1996). Moreover, activation of post-synaptic 5HT1A receptors elicits powerful inhibitory effects on hippocampal electrophysiology (e.g. Blier and de Montigny, 1987). These studies suggest that 5HT1A receptors may also be potential mediators of SSRI effects on seizures.

Preclinical studies have directly examined the influence of the serotonergic system in the effects of fluoxetine in epilepsy. Again, using the GEPRs, fluoxetine reduced the severity of sound-induced convulsions, and the time course of this effect paralleled the concurrent fluoxetine-induced rise in extracellular serotonin (Dailey et al., 1992). Also, using the pilocarpine-induced post-status epilepticus model of epilepsy, chronic fluoxetine treatment reduced brain excitability and also restored the epilepsy-induced serotonergic deficits (Mazarati et al., 2008). Intriguingly, the depression-like behavioural deficits associate with the model were not ameliorated by fluoxetine treatment, suggesting that depression in epilepsy may have distinct underlying mechanisms not related to serotonergic dysfunction. These initial results of preclinical studies provide promising support for a role of monoamines in mediating the effects of antidepressant on seizures and epileptogenesis.

Other neurotransmitter systems, including GABA and glutamate, may also be potential substrates of interaction for antidepressants on epileptogenesis. Imbalances in glutamate and GABA have long been implicated as the cause of convulsive seizures (Olsen et al., 1999), being an important area of investigation for the management of seizures and potentially epileptogenesis. Some reports suggest that antidepressants can influence the GABAergic system (Krystal et al., 2002), and while this suggests a future area of research, to date this evidence appears sparse. Similarly, effects of SSRIs on sodium (Pancrazio et al., 1998; Wang et al., 2008; Igelstrom and Heyward, 2012), potassium (Choi et al., 1999, 2001; Yeung et al., 1999; Lee and Kim, 2010; Lee et al., 2010) and calcium (Deak et al., 2000; Traboulsie et al., 2006) channels have been reported, which may be a mechanism by which SSRIs exert an anticonvulsant effect. However, there have been few studies to investigate this but suggest that this also represent areas for future investigations.

Brain-derived neurotrophic factor (BDNF)

BDNF is a key regulator of neuronal plasticity in both health and disease. It has been heavily implicated in epilepsy development, not least because of its modulatory roles on excitatory and inhibitory neurotransmission (Elmariah et al., 2004). BDNF is elevated in hippocampal tissue in PWE (Mathern et al., 1997; Takahashi et al., 1999; Murray et al., 2000), and its synthesis is increased by acute seizures (Ernfors et al., 1991; Rudge et al., 1998).

Several studies have demonstrated pro-epileptogenic effects of BDNF, including observations of spontaneous seizures after intra-hippocampal infusion (Scharfman et al., 2002) or transgenic overexpression of BDNF (Croll et al., 1999). Also, genetic deletion of TrkB, the primary signalling target of BDNF, prevents kindling epileptogenesis (He et al., 2004). Furthermore, BDNF elicits hyper-excitability in dentate granule cells in rodent (Asztely et al., 2000; Koyama et al., 2004) and human brain slices (Zhu and Roper, 2001). Conversely, others have demonstrated antiepileptic effects: chronic infusion of BDNF delays the development of electrical kindling (Larmet et al., 1995; Osehobo et al., 1999; Reibel et al., 2000a; 2000b), effects that may be due to desensitization of the signalling pathway (Reibel et al., 2000a), or to effects on neuropeptide Y and GABAergic inhibition (Koyama and Ikegaya, 2005).

BDNF has also been implicated in the aetiology of depression (Castren et al., 2007) and many studies suggest that BDNF and its receptor TrkB are involved in the mechanisms of antidepressant action (D'Sa and Duman, 2002; Nestler et al., 2002; Popoli et al., 2002). Several clinical studies have reported increases in serum BDNF levels following antidepressant treatment in depressed patients, which correlates with improvements in mood (Karege et al., 2002; Aydemir et al., 2005; 2006; Gervasoni et al., 2005; Gonul et al., 2005). In animal models, increases in BDNF mRNA (Nibuya et al., 1995; Russo-Neustadt et al., 2000; Dias et al., 2003) and protein (Chen et al., 2001; Altar et al., 2003; Xu et al., 2003), as well as TrkB expression (Nibuya et al., 1995) and activation (Saarelainen et al., 2003) in the hippocampus and prefrontal cortex have been shown following antidepressant treatment; however, others have failed to demonstrate such elevations (Miro et al., 2002; Coppell et al., 2003; Altieri et al., 2004). Functionally, these antidepressant-driven increases in BDNF and TrkB signalling may be essential for therapeutic activity (Duman et al., 1997; Castren et al., 2007; Adachi et al., 2008; Li et al., 2008). Despite the clear rationale for BDNF as an intermediary, to our knowledge, no studies have examined the role of neurotrophin signalling in antidepressant actions in seizures or epilepsy.

Inflammation

Inflammation is another biological process implicated in the pathogenesis of epilepsy (Vezzani et al., 2012), which is also influenced by antidepressants (Janssen et al., 2010). With respect to epilepsy, increases in inflammatory mediators have been identified in resected brain tissue from patients with temporal lobe epilepsy (Crespel et al., 2002; Zattoni et al., 2011; Hirvonen et al., 2012) and in experimental seizure models (Minami et al., 1991; Vezzani et al., 2000; Turrin and Rivest, 2004; Gorter et al., 2006). In some cases, such as following status epilepticus, these elevations can persist for weeks (Foresti et al., 2009; Xu et al., 2009). Moreover, various inflammatory mediators are involved in generating and exacerbating seizures (summarized in Vezzani et al., 2011). For example, genetic and pharmacological blockade of Toll-like receptor-4 (Maroso et al., 2010) or IL-1β (Vezzani et al., 2000; Dube et al., 2005) induces antiepileptic effects in animal models. Furthermore, patients with autoimmune disorders such as vasculitis or Crohn's disease may exhibit recurrent seizures (Najjar et al., 2008), and animal models of infection exhibit increased seizure susceptibility (Sayyah et al., 2003; Galic et al., 2008; 2009; Auvin et al., 2009), which may be due to the associated inflammation. Together this collective evidence promotes a causal role for inflammatory mechanisms in epileptogenesis and seizures.

Increases in various inflammatory mediators have also been reported in cases of depression. Most frequently reported is elevation of the cytokine, IL-6 (Maes et al., 1995; Tiemeier et al., 2003; Alesci et al., 2005; Brietzke et al., 2009) with similar reports for IL-1β and TNF-α (Lanquillon et al., 2000; Mikova et al., 2001; Owen et al., 2001). Treatment with SSRIs and SNRIs have anti-inflammatory effects, decreasing production of pro-inflammatory cytokines IL-1β, TNF-α, IFN-γ and increasing anti-inflammatory cytokines such as IL-10 (Kenis and Maes, 2002; Sutcigil et al., 2007), although other studies have failed to replicate this (Hannestad et al., 2011), or demonstrated pro-inflammatory effects of antidepressants (Piletz et al., 2009; Fornaro et al., 2011). Similar anti-inflammatory effects have been observed in animal studies where antidepressants attenuate cytokine production and concurrently reduce depressive-like symptoms (Castanon et al., 2001; Yirmiya et al., 2001; Song et al., 2009).

These data postulate a role for inflammation in epilepsy, which could be modulated by concurrent antidepressant treatment. This is further supported by demonstration of an interaction between fluoxetine and IL-1β in the pilocarpine-induced post-status epilepticus model. In a series of studies, Mazarati and colleagues first showed that this model induces depressive-like behaviours, dysregulation of the HPA axis and abnormal serotonin signalling. They then demonstrated that two weeks of intrahippocampal infusion of IL-1 receptor antagonist (IL-1ra) improved the depression-like phenotype in these animals (but had no effect on spontaneous seizures) (Mazarati et al., 2010). A subsequent study investigated the combination of fluoxetine and IL-1ra treatment in the same model, and found that while fluoxetine treatment in isolation had no effect, the combination treatment improved the depression-like behaviours and abnormalities in serotonin signalling, but only partially improving HPA axis hyper-reactivity (Pineda et al., 2012). This indicates that co-treatment of antidepressants with IL-1 receptor blockers, or other anti-inflammatory agents, may be an effective therapy for treatment-resistant depression and may reverse some of the pathological processes associated with epilepsy development.

Conclusions

With psychiatric disorders contributing a large proportion of the comorbidity in epilepsy, management of these disorders has become of great importance. In this paper, we have highlighted some of the potential implications for treating these disorders with antidepressants in PWE, particularly the effects of these drugs on seizure susceptibility and on the potential interaction with processes that are associated with epilepsy development and progression. While much of the literature to date has focused on the effects of antidepressants on short term end points, such as effects on seizure thresholds, future studies should focus upon the effects of antidepressants on longer term end points, such as seizure frequency over time, and on the neurobiological alterations that may influence epileptogenesis. In this paper, several lines of indirect evidence regarding some common potential substrates of interaction have been highlighted, areas that may be considered in future investigations of the effects of antidepressants in epilepsy. Primary among these are investigations on the production and functioning of newborn granule cells, in particular how newborn cells are affected by antidepressant treatment in epilepsy, as well as assessment of serotonergic receptor alterations (function/expression), particularly the 5HT1A receptor. Further exploration of the effects of antidepressants on ion channels and the influence of this on seizures and epilepsy are also warranted, given the emerging evidence in the area. Drugs targeting the serotonergic system, including SSRIs, should also be explored as potential therapies in treating epilepsy itself.

Overall, the majority of clinical and experimental data suggest that the effects of antidepressants on epilepsy itself are beneficial, indicating that antidepressants, in particular the newer generation compounds, SSRIs and SNRIs, should be considered safe for use in epilepsy. However, it is wise to be cautious of such data. While many studies have investigated how SSRIs affect seizures, few have investigated the underlying processes that may lead to the effects of SSRIs on seizures and in particular epileptogenesis. Future studies will need to investigate the effects of chronic SSRI exposure, investigating the effects on seizures at clinically relevant time points including prior to seizure onset, at onset and at the chronic stages, as well as investigating the effects on mediators of epileptogenesis and how these are altered with antidepressant treatment.

Glossary

8-OH-DPAT

8-hydroxy-2-(di-n-propylamino)tetralin

ACTH

adrenocorticotrophin releasing hormone

AED

antiepileptic drug

DOI

2,5-dimethoxy-4-iodoamphetamine

GAERS

Genetic Absence Epilepsy Rats from Strasbourg

GEPR

Genetically Epilepsy Prone Rats

HPA axis

hypothalamo-pituitary-adrenal axis

MDD

major depressive disorder

MES

maximal electroshock

NaSSA

noradrenaline and specific serotonergic antidepressant

NDRI

noradrenaline and dopamine reuptake inhibitor

NRI

noradrenaline reuptake inhibitor

PTZ

pentylenetetrazol

PWE

people with epilepsy

SARI

serotonin-2 receptor antagonist and reuptake inhibitor

SNRI

serotonin and noradrenaline reuptake inhibitor

SNRI

serotonin-noradrenaline reuptake inhibitors

SSRI

selective serotonin reuptake inhibitors

sz

seizure

TCA

tricyclic antidepressant

TeCA

tetracyclic antidepressant

WAG-Rij

Wistar Albino Glaxo rats from Rijswijk

Conflict of interest

The authors have no conflicts of interest, financial or otherwise, associated with this work to disclose.

References

  1. Adachi M, Barrot M, Autry AE, Theobald D, Monteggia LM. Selective loss of brain-derived neurotrophic factor in the dentate gyrus attenuates antidepressant efficacy. Biol Psychiatry. 2008;63:642–649. doi: 10.1016/j.biopsych.2007.09.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Adamec R, Shallow T. Rodent anxiety and kindling of the central amygdala and nucleus basalis. Physiol Behav. 2000;70:177–187. doi: 10.1016/s0031-9384(00)00250-x. [DOI] [PubMed] [Google Scholar]
  3. Adams SJ, O'Brien TJ, Lloyd J, Kilpatrick CJ, Salzberg MR, Velakoulis D. Neuropsychiatric morbidity in focal epilepsy. Br J Psychiatry. 2008;192:464–469. doi: 10.1192/bjp.bp.107.046664. [DOI] [PubMed] [Google Scholar]
  4. Adams SJ, Velakoulis D, Kaye AH, Corcoran NM, O'Brien TJ. Psychiatric history does not predict seizure outcome following temporal lobectomy for mesial temporal sclerosis. Epilepsia. 2012;53:1700–1704. doi: 10.1111/j.1528-1167.2012.03569.x. [DOI] [PubMed] [Google Scholar]
  5. Ago J, Ishikawa T, Matsumoto N, Ashequr Rahman M, Kamei C. Mechanism of imipramine-induced seizures in amygdala-kindled rats. Epilepsy Res. 2006;72:1–9. doi: 10.1016/j.eplepsyres.2006.06.022. [DOI] [PubMed] [Google Scholar]
  6. Ago J, Ishikawa T, Matsumoto N, Rahman MA, Kamei C. Epileptiformic activity induced by antidepressants in amygdala-kindled rats. Eur J Pharmacol. 2007;560:23–28. doi: 10.1016/j.ejphar.2006.11.050. [DOI] [PubMed] [Google Scholar]
  7. Ahern TH, Javors MA, Eagles DA, Martillotti J, Mitchell HA, Liles LC, et al. The effects of chronic norepinephrine transporter inactivation on seizure susceptibility in mice. Neuropsychopharmacology. 2006;31:730–738. doi: 10.1038/sj.npp.1300847. [DOI] [PubMed] [Google Scholar]
  8. Alesci S, Martinez PE, Kelkar S, Ilias I, Ronsaville DS, Listwak SJ, et al. Major depression is associated with significant diurnal elevations in plasma interleukin-6 levels, a shift of its circadian rhythm, and loss of physiological complexity in its secretion: clinical implications. J Clin Endocrinol Metab. 2005;90:2522–2530. doi: 10.1210/jc.2004-1667. [DOI] [PubMed] [Google Scholar]
  9. Ali I, Salzberg MR, French C, Jones NC. Electrophysiological insights into the enduring effects of early life stress on the brain. Psychopharmacology (Berl) 2011;214:155–173. doi: 10.1007/s00213-010-2125-z. [DOI] [PubMed] [Google Scholar]
  10. Alper K, Schwartz KA, Kolts RL, Khan A. Seizure incidence in psychopharmacological clinical trials: an analysis of Food and Drug Administration (FDA) summary basis of approval reports. Biol Psychiatry. 2007;62:345–354. doi: 10.1016/j.biopsych.2006.09.023. [DOI] [PubMed] [Google Scholar]
  11. Altar CA, Whitehead RE, Chen R, Wortwein G, Madsen TM. Effects of electroconvulsive seizures and antidepressant drugs on brain-derived neurotrophic factor protein in rat brain. Biol Psychiatry. 2003;54:703–709. doi: 10.1016/s0006-3223(03)00073-8. [DOI] [PubMed] [Google Scholar]
  12. Altieri M, Marini F, Arban R, Vitulli G, Jansson BO. Expression analysis of brain-derived neurotrophic factor (BDNF) mRNA isoforms after chronic and acute antidepressant treatment. Brain Res. 2004;1000:148–155. doi: 10.1016/j.brainres.2003.12.028. [DOI] [PubMed] [Google Scholar]
  13. Altman J, Das GD. Autoradiographic and histological evidence of postnatal hippocampal neurogenesis in rats. J Comp Neurol. 1965;124:319–335. doi: 10.1002/cne.901240303. [DOI] [PubMed] [Google Scholar]
  14. Arai S, Morita K, Kitayama S, Kumagai K, Kumagai M, Kihira K, et al. Chronic inhibition of the norepinephrine transporter in the brain participates in seizure sensitization to cocaine and local anesthetics. Brain Res. 2003;964:83–90. doi: 10.1016/s0006-8993(02)04068-4. [DOI] [PubMed] [Google Scholar]
  15. Asmussen SB, Kirlin KA, Gale SD, Chung SS. Differences in self-reported depressive symptoms between patients with epileptic and psychogenic nonepileptic seizures. Seizure. 2009;18:564–566. doi: 10.1016/j.seizure.2009.05.006. [DOI] [PubMed] [Google Scholar]
  16. Assem-Hilger E, Lanzenberger R, Savli M, Wadsak W, Mitterhauser M, Mien LK, et al. Central serotonin 1A receptor binding in temporal lobe epilepsy: a [carbonyl-(11)C]WAY-100635 PET study. Epilepsy Behav. 2010;19:467–473. doi: 10.1016/j.yebeh.2010.07.030. [DOI] [PubMed] [Google Scholar]
  17. Asztely F, Kokaia M, Olofsdotter K, Ortegren U, Lindvall O. Afferent-specific modulation of short-term synaptic plasticity by neurotrophins in dentate gyrus. Eur J Neurosci. 2000;12:662–669. doi: 10.1046/j.1460-9568.2000.00956.x. [DOI] [PubMed] [Google Scholar]
  18. Auvin S, Porta N, Nehlig A, Lecointe C, Vallee L, Bordet R. Inflammation in rat pups subjected to short hyperthermic seizures enhances brain long-term excitability. Epilepsy Res. 2009;86:124–130. doi: 10.1016/j.eplepsyres.2009.05.010. [DOI] [PubMed] [Google Scholar]
  19. Aydemir C, Yalcin ES, Aksaray S, Kisa C, Yildirim SG, Uzbay T, et al. Brain-derived neurotrophic factor (BDNF) changes in the serum of depressed women. Prog Neuropsychopharmacol Biol Psychiatry. 2006;30:1256–1260. doi: 10.1016/j.pnpbp.2006.03.025. [DOI] [PubMed] [Google Scholar]
  20. Aydemir O, Deveci A, Taneli F. The effect of chronic antidepressant treatment on serum brain-derived neurotrophic factor levels in depressed patients: a preliminary study. Prog Neuropsychopharmacol Biol Psychiatry. 2005;29:261–265. doi: 10.1016/j.pnpbp.2004.11.009. [DOI] [PubMed] [Google Scholar]
  21. Babb TL. Pathology of the temporal lobe: hippocampal sclerosis. In: Luders HO, Comair YG, editors. Epilepsy Surgery. 2nd edn. Philadelphia, PA: Lippincott Williams & Wilkins; 2001. pp. 901–906. [Google Scholar]
  22. Babu CS, Satishchandra P, Sinha S, Subbakrishna DK. Co-morbidities in people living with epilepsy: hospital based case-control study from a resource-poor setting. Epilepsy Res. 2009;86:146–152. doi: 10.1016/j.eplepsyres.2009.05.015. [DOI] [PubMed] [Google Scholar]
  23. Bagdy G, Kecskemeti V, Riba P, Jakus R. Serotonin and epilepsy. J Neurochem. 2007;100:857–873. doi: 10.1111/j.1471-4159.2006.04277.x. [DOI] [PubMed] [Google Scholar]
  24. Banerjee PN, Filippi D, Allen Hauser W. The descriptive epidemiology of epilepsy-a review. Epilepsy Res. 2009;85:31–45. doi: 10.1016/j.eplepsyres.2009.03.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Barden N. Regulation of corticosteroid receptor gene expression in depression and antidepressant action. J Psychiatry Neurosci. 1999;24:25–39. [PMC free article] [PubMed] [Google Scholar]
  26. Barry JJ, Ettinger AB, Friel P, Gilliam FG, Harden CL, Hermann B, et al. Consensus statement: the evaluation and treatment of people with epilepsy and affective disorders. Epilepsy Behav. 2008;13(Suppl. 1):S1–S29. doi: 10.1016/j.yebeh.2008.04.005. [DOI] [PubMed] [Google Scholar]
  27. Bengzon J, Kokaia Z, Elmer E, Nanobashvili A, Kokaia M, Lindvall O. Apoptosis and proliferation of dentate gyrus neurons after single and intermittent limbic seizures. Proc Natl Acad Sci U S A. 1997;94:10432–10437. doi: 10.1073/pnas.94.19.10432. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Berg AT, Scheffer IE. New concepts in classification of the epilepsies: entering the 21st century. Epilepsia. 2011;52:1058–1062. doi: 10.1111/j.1528-1167.2011.03101.x. [DOI] [PubMed] [Google Scholar]
  29. Berg AT, Berkovic SF, Brodie MJ, Buchhalter J, Cross JH, van Emde Boas W, et al. Revised terminology and concepts for organization of seizures and epilepsies: report of the ILAE Commission on Classification and Terminology, 2005–2009. Epilepsia. 2010;51:676–685. doi: 10.1111/j.1528-1167.2010.02522.x. [DOI] [PubMed] [Google Scholar]
  30. Bessa JM, Ferreira D, Melo I, Marques F, Cerqueira JJ, Palha JA, et al. The mood-improving actions of antidepressants do not depend on neurogenesis but are associated with neuronal remodeling. Mol Psychiatry. 2009;14:764–773. doi: 10.1038/mp.2008.119. 739. [DOI] [PubMed] [Google Scholar]
  31. Blier P, de Montigny C. Modification of 5-HT neuron properties by sustained administration of the 5-HT1A agonist gepirone: electrophysiological studies in the rat brain. Synapse. 1987;1:470–480. doi: 10.1002/syn.890010511. [DOI] [PubMed] [Google Scholar]
  32. Boldrini M, Underwood MD, Hen R, Rosoklija GB, Dwork AJ, John Mann J, et al. Antidepressants increase neural progenitor cells in the human hippocampus. Neuropsychopharmacology. 2009;34:2376–2389. doi: 10.1038/npp.2009.75. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Bonelli SB, Powell R, Yogarajah M, Thompson PJ, Symms MR, Koepp MJ, et al. Preoperative amygdala fMRI in temporal lobe epilepsy. Epilepsia. 2009;50:217–227. doi: 10.1111/j.1528-1167.2008.01739.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Bonnet U. Moclobemide: therapeutic use and clinical studies. CNS Drug Rev. 2003;9:97–140. doi: 10.1111/j.1527-3458.2003.tb00245.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Borowicz KK, Stepien K, Czuczwar SJ. Fluoxetine enhances the anticonvulsant effects of conventional antiepileptic drugs in maximal electroshock seizures in mice. Pharmacol Rep. 2006;58:83–90. [PubMed] [Google Scholar]
  36. Borowicz KK, Furmanek-Karwowska K, Sawicka K, Luszczki JJ, Czuczwar SJ. Chronically administered fluoxetine enhances the anticonvulsant activity of conventional antiepileptic drugs in the mouse maximal electroshock model. Eur J Pharmacol. 2007;567:77–82. doi: 10.1016/j.ejphar.2007.03.015. [DOI] [PubMed] [Google Scholar]
  37. Borowicz KK, Golyska D, Luszczki JJ, Czuczwar SJ. Effect of acutely and chronically administered venlafaxine on the anticonvulsant action of classical antiepileptic drugs in the mouse maximal electroshock model. Eur J Pharmacol. 2011;670:114–120. doi: 10.1016/j.ejphar.2011.08.042. [DOI] [PubMed] [Google Scholar]
  38. Bouilleret V, Hogan RE, Velakoulis D, Salzberg MR, Wang L, Egan GF, et al. Morphometric abnormalities and hyperanxiety in genetically epileptic rats: a model of psychiatric comorbidity? Neuroimage. 2009;45:267–274. doi: 10.1016/j.neuroimage.2008.12.019. [DOI] [PubMed] [Google Scholar]
  39. Briellmann RS, Hopwood MJ, Jackson GD. Major depression in temporal lobe epilepsy with hippocampal sclerosis: clinical and imaging correlates. J Neurol Neurosurg Psychiatry. 2007;78:1226–1230. doi: 10.1136/jnnp.2006.104521. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Brietzke E, Stertz L, Fernandes BS, Kauer-Sant'anna M, Mascarenhas M, Escosteguy Vargas A, et al. Comparison of cytokine levels in depressed, manic and euthymic patients with bipolar disorder. J Affect Disord. 2009;116:214–217. doi: 10.1016/j.jad.2008.12.001. [DOI] [PubMed] [Google Scholar]
  41. Browning RA, Wood AV, Merrill MA, Dailey JW, Jobe PC. Enhancement of the anticonvulsant effect of fluoxetine following blockade of 5-HT1A receptors. Eur J Pharmacol. 1997;336:1–6. doi: 10.1016/s0014-2999(97)01215-6. [DOI] [PubMed] [Google Scholar]
  42. Buckmaster PS, Zhang GF, Yamawaki R. Axon sprouting in a model of temporal lobe epilepsy creates a predominantly excitatory feedback circuit. J Neurosci. 2002;22:6650–6658. doi: 10.1523/JNEUROSCI.22-15-06650.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Busch FN, Sandberg LS. Combined treatment of depression. Psychiatr Clin North Am. 2012;35:165–179. doi: 10.1016/j.psc.2011.10.002. [DOI] [PubMed] [Google Scholar]
  44. Castanon N, Bluthe RM, Dantzer R. Chronic treatment with the atypical antidepressant tianeptine attenuates sickness behavior induced by peripheral but not central lipopolysaccharide and interleukin-1beta in the rat. Psychopharmacology (Berl) 2001;154:50–60. doi: 10.1007/s002130000595. [DOI] [PubMed] [Google Scholar]
  45. Castren E, Voikar V, Rantamaki T. Role of neurotrophic factors in depression. Curr Opin Pharmacol. 2007;7:18–21. doi: 10.1016/j.coph.2006.08.009. [DOI] [PubMed] [Google Scholar]
  46. Ceyhan M, Kayir H, Uzbay IT. Investigation of the effects of tianeptine and fluoxetine on pentylenetetrazole-induced seizures in rats. J Psychiatr Res. 2005;39:191–196. doi: 10.1016/j.jpsychires.2004.06.002. [DOI] [PubMed] [Google Scholar]
  47. Chen B, Dowlatshahi D, MacQueen GM, Wang JF, Young LT. Increased hippocampal BDNF immunoreactivity in subjects treated with antidepressant medication. Biol Psychiatry. 2001;50:260–265. doi: 10.1016/s0006-3223(01)01083-6. [DOI] [PubMed] [Google Scholar]
  48. Choi JS, Hahn SJ, Rhie DJ, Yoon SH, Jo YH, Kim MS. Mechanism of fluoxetine block of cloned voltage-activated potassium channel Kv1.3. J Pharmacol Exp Ther. 1999;291:1–6. [PubMed] [Google Scholar]
  49. Choi BH, Choi JS, Yoon SH, Rhie DJ, Min DS, Jo YH, et al. Effects of norfluoxetine, the major metabolite of fluoxetine, on the cloned neuronal potassium channel Kv3.1. Neuropharmacology. 2001;41:443–453. doi: 10.1016/s0028-3908(01)00088-0. [DOI] [PubMed] [Google Scholar]
  50. Ciechanowski P, Chaytor N, Miller J, Fraser R, Russo J, Unutzer J, et al. PEARLS depression treatment for individuals with epilepsy: a randomized controlled trial. Epilepsy Behav. 2010;19:225–231. doi: 10.1016/j.yebeh.2010.06.003. [DOI] [PubMed] [Google Scholar]
  51. Clinckers R, Smolders I, Meurs A, Ebinger G, Michotte Y. Anticonvulsant action of GBR-12909 and citalopram against acute experimentally induced limbic seizures. Neuropharmacology. 2004;47:1053–1061. doi: 10.1016/j.neuropharm.2004.07.032. [DOI] [PubMed] [Google Scholar]
  52. Colman I, Ataullahjan A. Life course perspectives on the epidemiology of depression. Can J Psychiatry. 2010;55:622–632. doi: 10.1177/070674371005501002. [DOI] [PubMed] [Google Scholar]
  53. Coppell AL, Pei Q, Zetterstrom TS. Bi-phasic change in BDNF gene expression following antidepressant drug treatment. Neuropharmacology. 2003;44:903–910. doi: 10.1016/s0028-3908(03)00077-7. [DOI] [PubMed] [Google Scholar]
  54. Corsellis JA. The incidence of Ammon's horn sclerosis. Brain. 1957;80:193–208. doi: 10.1093/brain/80.2.193. [DOI] [PubMed] [Google Scholar]
  55. Cotter D, Landau S, Beasley C, Stevenson R, Chana G, MacMillan L, et al. The density and spatial distribution of GABAergic neurons, labelled using calcium binding proteins, in the anterior cingulate cortex in major depressive disorder, bipolar disorder, and schizophrenia. Biol Psychiatry. 2002;51:377–386. doi: 10.1016/s0006-3223(01)01243-4. [DOI] [PubMed] [Google Scholar]
  56. Cotterman-Hart S. Depression in epilepsy: why aren't we treating? Epilepsy Behav. 2010;19:419–421. doi: 10.1016/j.yebeh.2010.08.018. [DOI] [PubMed] [Google Scholar]
  57. Cottrell GA, Nyakas C, de Kloet ER, Bohus B. Hippocampal kindling: corticosterone modulation of induced seizures. Brain Res. 1984;309:377–381. doi: 10.1016/0006-8993(84)90608-5. [DOI] [PubMed] [Google Scholar]
  58. Crespel A, Coubes P, Rousset MC, Brana C, Rougier A, Rondouin G, et al. Inflammatory reactions in human medial temporal lobe epilepsy with hippocampal sclerosis. Brain Res. 2002;952:159–169. doi: 10.1016/s0006-8993(02)03050-0. [DOI] [PubMed] [Google Scholar]
  59. Croll SD, Suri C, Compton DL, Simmons MV, Yancopoulos GD, Lindsay RM, et al. Brain-derived neurotrophic factor transgenic mice exhibit passive avoidance deficits, increased seizure severity and in vitro hyperexcitability in the hippocampus and entorhinal cortex. Neuroscience. 1999;93:1491–1506. doi: 10.1016/s0306-4522(99)00296-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. D'Sa C, Duman RS. Antidepressants and neuroplasticity. Bipolar Disord. 2002;4:183–194. doi: 10.1034/j.1399-5618.2002.01203.x. [DOI] [PubMed] [Google Scholar]
  61. Dailey JW, Jobe PC. Indices of noradrenergic function in the central nervous system of seizure-naive genetically epilepsy-prone rats. Epilepsia. 1986;27:665–670. doi: 10.1111/j.1528-1157.1986.tb03593.x. [DOI] [PubMed] [Google Scholar]
  62. Dailey JW, Yan QS, Mishra PK, Burger RL, Jobe PC. Effects of fluoxetine on convulsions and on brain serotonin as detected by microdialysis in genetically epilepsy-prone rats. J Pharmacol Exp Ther. 1992;260:533–540. [PubMed] [Google Scholar]
  63. Deak F, Lasztoczi B, Pacher P, Petheo GL, Valeria K, Spat A. Inhibition of voltage-gated calcium channels by fluoxetine in rat hippocampal pyramidal cells. Neuropharmacology. 2000;39:1029–1036. doi: 10.1016/s0028-3908(99)00206-3. [DOI] [PubMed] [Google Scholar]
  64. Desai SD, Shukla G, Goyal V, Singh S, Padma MV, Tripathi M, et al. Study of DSM-IV Axis I psychiatric disorders in patients with refractory complex partial seizures using a short structured clinical interview. Epilepsy Behav. 2010;19:301–305. doi: 10.1016/j.yebeh.2010.07.005. [DOI] [PubMed] [Google Scholar]
  65. Dessain EC, Schatzberg AF, Woods BT, Cole JO. Maprotiline treatment in depression. A perspective on seizures. Arch Gen Psychiatry. 1986;43:86–90. doi: 10.1001/archpsyc.1986.01800010088011. [DOI] [PubMed] [Google Scholar]
  66. Dias BG, Banerjee SB, Duman RS, Vaidya VA. Differential regulation of brain derived neurotrophic factor transcripts by antidepressant treatments in the adult rat brain. Neuropharmacology. 2003;45:553–563. doi: 10.1016/s0028-3908(03)00198-9. [DOI] [PubMed] [Google Scholar]
  67. Dube C, Vezzani A, Behrens M, Bartfai T, Baram TZ. Interleukin-1beta contributes to the generation of experimental febrile seizures. Ann Neurol. 2005;57:152–155. doi: 10.1002/ana.20358. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Duman RS, Heninger GR, Nestler EJ. A molecular and cellular theory of depression. Arch Gen Psychiatry. 1997;54:597–606. doi: 10.1001/archpsyc.1997.01830190015002. [DOI] [PubMed] [Google Scholar]
  69. Edwards HE, Burnham WM, Mendonca A, Bowlby DA, MacLusky NJ. Steroid hormones affect limbic afterdischarge thresholds and kindling rates in adult female rats. Brain Res. 1999;838:136–150. doi: 10.1016/s0006-8993(99)01619-4. [DOI] [PubMed] [Google Scholar]
  70. Elmariah SB, Crumling MA, Parsons TD, Balice-Gordon RJ. Postsynaptic TrkB-mediated signaling modulates excitatory and inhibitory neurotransmitter receptor clustering at hippocampal synapses. J Neurosci. 2004;24:2380–2393. doi: 10.1523/JNEUROSCI.4112-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Elst LT, Groffmann M, Ebert D, Schulze-Bonhage A. Amygdala volume loss in patients with dysphoric disorder of epilepsy. Epilepsy Behav. 2009;16:105–112. doi: 10.1016/j.yebeh.2009.06.009. [DOI] [PubMed] [Google Scholar]
  72. Encinas JM, Vaahtokari A, Enikolopov G. Fluoxetine targets early progenitor cells in the adult brain. Proc Natl Acad Sci U S A. 2006;103:8233–8238. doi: 10.1073/pnas.0601992103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Engel J, Pedley T, editors. Epilepsy: a Comprehensive Textbook. Philadelphia, PA: Wolters Kluwer Health/Lippincott Williams & Wilkins; 2007. [Google Scholar]
  74. Engel J, Pedley T. What Is Epilepsy? Philadelphia, PA: Lippincott Williams & Wilkins; 2008. [Google Scholar]
  75. Eriksson PS, Perfilieva E, Bjork-Eriksson T, Alborn AM, Nordborg C, Peterson DA, et al. Neurogenesis in the adult human hippocampus. Nat Med. 1998;4:1313–1317. doi: 10.1038/3305. [DOI] [PubMed] [Google Scholar]
  76. Ernfors P, Bengzon J, Kokaia Z, Persson H, Lindvall O. Increased levels of messenger RNAs for neurotrophic factors in the brain during kindling epileptogenesis. Neuron. 1991;7:165–176. doi: 10.1016/0896-6273(91)90084-d. [DOI] [PubMed] [Google Scholar]
  77. Favale E, Rubino V, Mainardi P, Lunardi G, Albano C. Anticonvulsant effect of fluoxetine in humans. Neurology. 1995;45:1926–1927. doi: 10.1212/wnl.45.10.1926. [DOI] [PubMed] [Google Scholar]
  78. Favale E, Audenino D, Cocito L, Albano C. The anticonvulsant effect of citalopram as an indirect evidence of serotonergic impairment in human epileptogenesis. Seizure. 2003;12:316–318. doi: 10.1016/s1059-1311(02)00315-1. [DOI] [PubMed] [Google Scholar]
  79. FDA. 2008. Statistical review and evaluation: antiepileptic drugs and suicidality. Available at: http://www.fda.gov/ohrms/dockets/ac/08/briefing/2008-4372b1-01-FDA.pdf (accessed 2/25/2013)
  80. Ferrero AJ, Cereseto M, Reines A, Bonavita CD, Sifonios LL, Rubio MC, et al. Chronic treatment with fluoxetine decreases seizure threshold in naive but not in rats exposed to the learned helplessness paradigm: correlation with the hippocampal glutamate release. Prog Neuropsychopharmacol Biol Psychiatry. 2005;29:678–686. doi: 10.1016/j.pnpbp.2005.04.013. [DOI] [PubMed] [Google Scholar]
  81. Filho GM, Rosa VP, Lin K, Caboclo LO, Sakamoto AC, Yacubian EM. Psychiatric comorbidity in epilepsy: a study comparing patients with mesial temporal sclerosis and juvenile myoclonic epilepsy. Epilepsy Behav. 2008;13:196–201. doi: 10.1016/j.yebeh.2008.01.008. [DOI] [PubMed] [Google Scholar]
  82. Finegersh A, Avedissian C, Shamim S, Dustin I, Thompson PM, Theodore WH. Bilateral hippocampal atrophy in temporal lobe epilepsy: effect of depressive symptoms and febrile seizures. Epilepsia. 2011;52:689–697. doi: 10.1111/j.1528-1167.2010.02928.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Fisher RS, van Emde Boas W, Blume W, Elger C, Genton P, Lee P, et al. Epileptic seizures and epilepsy: definitions proposed by the International League Against Epilepsy (ILAE) and the International Bureau for Epilepsy (IBE) Epilepsia. 2005;46:470–472. doi: 10.1111/j.0013-9580.2005.66104.x. [DOI] [PubMed] [Google Scholar]
  84. Foresti ML, Arisi GM, Katki K, Montanez A, Sanchez RM, Shapiro LA. Chemokine CCL2 and its receptor CCR2 are increased in the hippocampus following pilocarpine-induced status epilepticus. J Neuroinflammation. 2009;6:40. doi: 10.1186/1742-2094-6-40. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Fornaro M, Martino M, Battaglia F, Colicchio S, Perugi G. Increase in IL-6 levels among major depressive disorder patients after a 6-week treatment with duloxetine 60 mg/day: a preliminary observation. Neuropsychiatr Dis Treat. 2011;7:51–56. doi: 10.2147/NDT.S16382. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Forsgren L, Nystrom L. An incident case-referent study of epileptic seizures in adults. Epilepsy Res. 1990;6:66–81. doi: 10.1016/0920-1211(90)90010-s. [DOI] [PubMed] [Google Scholar]
  87. Fountoulakis KN, Gonda X, Samara M, Siapera M, Karavelas V, Ristic DI, et al. Antiepileptic drugs and suicidality. J Psychopharmacol. 2012;26:1401–1407. doi: 10.1177/0269881112440514. [DOI] [PubMed] [Google Scholar]
  88. Friedman DE, Kung DH, Laowattana S, Kass JS, Hrachovy RA, Levin HS. Identifying depression in epilepsy in a busy clinical setting is enhanced with systematic screening. Seizure. 2009;18:429–433. doi: 10.1016/j.seizure.2009.03.001. [DOI] [PubMed] [Google Scholar]
  89. Frisch C, Hanke J, Kleineruschkamp S, Roske S, Kaaden S, Elger CE, et al. Positive correlation between the density of neuropeptide y positive neurons in the amygdala and parameters of self-reported anxiety and depression in mesiotemporal lobe epilepsy patients. Biol Psychiatry. 2009;66:433–440. doi: 10.1016/j.biopsych.2009.03.025. [DOI] [PubMed] [Google Scholar]
  90. Fuller-Thomson E, Brennenstuhl S. The association between depression and epilepsy in a nationally representative sample. Epilepsia. 2009;50:1051–1058. doi: 10.1111/j.1528-1167.2008.01803.x. [DOI] [PubMed] [Google Scholar]
  91. Furmaga H, Carreno FR, Frazer A. Vagal nerve stimulation rapidly activates brain-derived neurotrophic factor receptor TrkB in rat brain. PLoS One. 2012;7:e34844. doi: 10.1371/journal.pone.0034844. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Gaitatzis A, Carroll K, Majeed A, Sander JW. The epidemiology of the comorbidity of epilepsy in the general population. Epilepsia. 2004;45:1613–1622. doi: 10.1111/j.0013-9580.2004.17504.x. [DOI] [PubMed] [Google Scholar]
  93. Gaitatzis A, Sisodiya SM, Sander JW. The somatic comorbidity of epilepsy: a weighty but often unrecognized burden. Epilepsia. 2012;53:1282–1293. doi: 10.1111/j.1528-1167.2012.03528.x. [DOI] [PubMed] [Google Scholar]
  94. Galic MA, Riazi K, Heida JG, Mouihate A, Fournier NM, Spencer SJ, et al. Postnatal inflammation increases seizure susceptibility in adult rats. J Neurosci. 2008;28:6904–6913. doi: 10.1523/JNEUROSCI.1901-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Galic MA, Riazi K, Henderson AK, Tsutsui S, Pittman QJ. Viral-like brain inflammation during development causes increased seizure susceptibility in adult rats. Neurobiol Dis. 2009;36:343–351. doi: 10.1016/j.nbd.2009.07.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Gallagher BB. Endocrine abnormalities in human temporal lobe epilepsy. Yale J Biol Med. 1987;60:93–97. [PMC free article] [PubMed] [Google Scholar]
  97. Gallagher BB, Murvin A, Flanigin HF, King DW, Luney D. Pituitary and adrenal function in epileptic patients. Epilepsia. 1984;25:683–689. doi: 10.1111/j.1528-1157.1984.tb03477.x. [DOI] [PubMed] [Google Scholar]
  98. Gariboldi M, Tutka P, Samanin R, Vezzani A. Stimulation of 5-HT1A receptors in the dorsal hippocampus and inhibition of limbic seizures induced by kainic acid in rats. Br J Pharmacol. 1996;119:813–818. doi: 10.1111/j.1476-5381.1996.tb15745.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Gervasoni N, Aubry JM, Bondolfi G, Osiek C, Schwald M, Bertschy G, et al. Partial normalization of serum brain-derived neurotrophic factor in remitted patients after a major depressive episode. Neuropsychobiology. 2005;51:234–238. doi: 10.1159/000085725. [DOI] [PubMed] [Google Scholar]
  100. Gigli GL, Diomedi M, Troisi A, Baldinetti F, Marciani MG, Girolami E, et al. Lack of potentiation of anticonvulsant effect by fluoxetine in drug-resistant epilepsy. Seizure. 1994;3:221–224. doi: 10.1016/s1059-1311(05)80192-x. [DOI] [PubMed] [Google Scholar]
  101. Gilby KL, Sydserff S, Patey AM, Thorne V, St-Onge V, Jans J, et al. Postnatal epigenetic influences on seizure susceptibility in seizure-prone versus seizure-resistant rat strains. Behav Neurosci. 2009;123:337–346. doi: 10.1037/a0014730. [DOI] [PubMed] [Google Scholar]
  102. Gilliam FG, Barry JJ, Hermann BP, Meador KJ, Vahle V, Kanner AM. Rapid detection of major depression in epilepsy: a multicentre study. Lancet Neurol. 2006;5:399–405. doi: 10.1016/S1474-4422(06)70415-X. [DOI] [PubMed] [Google Scholar]
  103. Gilliam FG, Maton BM, Martin RC, Sawrie SM, Faught RE, Hugg JW, et al. Hippocampal 1H-MRSI correlates with severity of depression symptoms in temporal lobe epilepsy. Neurology. 2007;68:364–368. doi: 10.1212/01.wnl.0000252813.86812.81. [DOI] [PubMed] [Google Scholar]
  104. Gonul AS, Akdeniz F, Taneli F, Donat O, Eker C, Vahip S. Effect of treatment on serum brain-derived neurotrophic factor levels in depressed patients. Eur Arch Psychiatry Clin Neurosci. 2005;255:381–386. doi: 10.1007/s00406-005-0578-6. [DOI] [PubMed] [Google Scholar]
  105. Gorter JA, van Vliet EA, Aronica E, Breit T, Rauwerda H, Lopes da Silva FH, et al. Potential new antiepileptogenic targets indicated by microarray analysis in a rat model for temporal lobe epilepsy. J Neurosci. 2006;26:11083–11110. doi: 10.1523/JNEUROSCI.2766-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Groticke I, Hoffmann K, Loscher W. Behavioral alterations in the pilocarpine model of temporal lobe epilepsy in mice. Exp Neurol. 2007;207:329–349. doi: 10.1016/j.expneurol.2007.06.021. [DOI] [PubMed] [Google Scholar]
  107. Groticke I, Hoffmann K, Loscher W. Behavioral alterations in a mouse model of temporal lobe epilepsy induced by intrahippocampal injection of kainate. Exp Neurol. 2008;213:71–83. doi: 10.1016/j.expneurol.2008.04.036. [DOI] [PubMed] [Google Scholar]
  108. Hagan JJ, Hatcher JP, Slade PD. The role of 5-HT1D and 5-HT1A receptors in mediating 5-hydroxytryptophan induced myoclonic jerks in guinea pigs. Eur J Pharmacol. 1995;294:743–751. doi: 10.1016/0014-2999(95)00627-3. [DOI] [PubMed] [Google Scholar]
  109. Hannestad J, DellaGioia N, Bloch M. The effect of antidepressant medication treatment on serum levels of inflammatory cytokines: a meta-analysis. Neuropsychopharmacology. 2011;36:2452–2459. doi: 10.1038/npp.2011.132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Harmant J, van Rijckevorsel-Harmant K, de Barsy T, Hendrickx B. Fluvoxamine: an antidepressant with low (or no) epileptogenic effect. Lancet. 1990;336:386. doi: 10.1016/0140-6736(90)91938-7. [DOI] [PubMed] [Google Scholar]
  111. Hasler G, Bonwetsch R, Giovacchini G, Toczek MT, Bagic A, Luckenbaugh DA, et al. 5-HT1A receptor binding in temporal lobe epilepsy patients with and without major depression. Biol Psychiatry. 2007;62:1258–1264. doi: 10.1016/j.biopsych.2007.02.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Hattiangady B, Shetty AK. Decreased neuronal differentiation of newly generated cells underlies reduced hippocampal neurogenesis in chronic temporal lobe epilepsy. Hippocampus. 2010;20:97–112. doi: 10.1002/hipo.20594. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Hattiangady B, Rao MS, Shetty AK. Chronic temporal lobe epilepsy is associated with severely declined dentate neurogenesis in the adult hippocampus. Neurobiol Dis. 2004;17:473–490. doi: 10.1016/j.nbd.2004.08.008. [DOI] [PubMed] [Google Scholar]
  114. He XP, Kotloski R, Nef S, Luikart BW, Parada LF, McNamara JO. Conditional deletion of TrkB but not BDNF prevents epileptogenesis in the kindling model. Neuron. 2004;43:31–42. doi: 10.1016/j.neuron.2004.06.019. [DOI] [PubMed] [Google Scholar]
  115. Hermann B, Jacoby A. The psychosocial impact of epilepsy in adults. Epilepsy Behav. 2009;15(Suppl. 1):S11–S16. doi: 10.1016/j.yebeh.2009.03.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Hermann B, Seidenberg M, Jones J. The neurobehavioural comorbidities of epilepsy: can a natural history be developed? Lancet Neurol. 2008;7:151–160. doi: 10.1016/S1474-4422(08)70018-8. [DOI] [PubMed] [Google Scholar]
  117. Hermann BP, Seidenberg M, Bell B. Psychiatric comorbidity in chronic epilepsy: identification, consequences, and treatment of major depression. Epilepsia. 2000;41(Suppl. 2):S31–S41. doi: 10.1111/j.1528-1157.2000.tb01522.x. [DOI] [PubMed] [Google Scholar]
  118. Hernandez EJ, Williams PA, Dudek FE. Effects of fluoxetine and TFMPP on spontaneous seizures in rats with pilocarpine-induced epilepsy. Epilepsia. 2002;43:1337–1345. doi: 10.1046/j.1528-1157.2002.48701.x. [DOI] [PubMed] [Google Scholar]
  119. Hesdorffer D, Krishnamoorthy E. Neuropsychiatric disorders in epilepsy: epidemiology and classification. In: Trimble M, Schmitz B, editors. Neuropsychiatric Disorders in Epilepsy: Epidemiology and Classification. Cambridge: Cambridge University Press; 2011. pp. 3–13. [Google Scholar]
  120. Hesdorffer DC, Hauser WA, Annegers JF, Cascino G. Major depression is a risk factor for seizures in older adults. Ann Neurol. 2000;47:246–249. [PubMed] [Google Scholar]
  121. Hesdorffer DC, Hauser WA, Olafsson E, Ludvigsson P, Kjartansson O. Depression and suicide attempt as risk factors for incident unprovoked seizures. Ann Neurol. 2006;59:35–41. doi: 10.1002/ana.20685. [DOI] [PubMed] [Google Scholar]
  122. Hesdorffer DC, Ishihara L, Mynepalli L, Webb DJ, Weil J, Hauser WA. Epilepsy, suicidality, and psychiatric disorders: a bidirectional association. Ann Neurol. 2012;72:184–191. doi: 10.1002/ana.23601. [DOI] [PubMed] [Google Scholar]
  123. Himmerich H, Zimmermann P, Ising M, Kloiber S, Lucae S, Kunzel HE, et al. Changes in the hypothalamic-pituitary-adrenal axis and leptin levels during antidepressant treatment. Neuropsychobiology. 2007;55:28–35. doi: 10.1159/000103573. [DOI] [PubMed] [Google Scholar]
  124. Hirvonen J, Kreisl WC, Fujita M, Dustin I, Khan O, Appel S, et al. Increased in vivo expression of an inflammatory marker in temporal lobe epilepsy. J Nucl Med. 2012;53:234–240. doi: 10.2967/jnumed.111.091694. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Hitiris N, Mohanraj R, Norrie J, Sills GJ, Brodie MJ. Predictors of pharmacoresistant epilepsy. Epilepsy Res. 2007;75:192–196. doi: 10.1016/j.eplepsyres.2007.06.003. [DOI] [PubMed] [Google Scholar]
  126. Holsboer F. Stress, hypercortisolism and corticosteroid receptors in depression: implications for therapy. J Affect Disord. 2001;62:77–91. doi: 10.1016/s0165-0327(00)00352-9. [DOI] [PubMed] [Google Scholar]
  127. Hoppe C, Elger CE. Depression in epilepsy: a critical review from a clinical perspective. Nat Rev Neurol. 2011;7:462–472. doi: 10.1038/nrneurol.2011.104. [DOI] [PubMed] [Google Scholar]
  128. Hovorka J, Herman E, Nemcova II. Treatment of interictal depression with citalopram in patients with epilepsy. Epilepsy Behav. 2000;1:444–447. doi: 10.1006/ebeh.2000.0123. [DOI] [PubMed] [Google Scholar]
  129. Igelstrom KM. Preclinical antiepileptic actions of selective serotonin reuptake inhibitors – implications for clinical trial design. Epilepsia. 2012;53:596–605. doi: 10.1111/j.1528-1167.2012.03427.x. [DOI] [PubMed] [Google Scholar]
  130. Igelstrom KM, Heyward PM. The antidepressant drug fluoxetine inhibits persistent sodium currents and seizure-like events. Epilepsy Res. 2012;101:174–181. doi: 10.1016/j.eplepsyres.2012.03.019. [DOI] [PubMed] [Google Scholar]
  131. Isbister GK, Bowe SJ, Dawson A, Whyte IM. Relative toxicity of selective serotonin reuptake inhibitors (SSRIs) in overdose. J Toxicol Clin Toxicol. 2004;42:277–285. doi: 10.1081/clt-120037428. [DOI] [PubMed] [Google Scholar]
  132. Jaako K, Zharkovsky T, Zharkovsky A. Effects of repeated citalopram treatment on kainic acid-induced neurogenesis in adult mouse hippocampus. Brain Res. 2009;1288:18–28. doi: 10.1016/j.brainres.2009.06.089. [DOI] [PubMed] [Google Scholar]
  133. Jaako K, Aonurm-Helm A, Kalda A, Anier K, Zharkovsky T, Shastin D, et al. Repeated citalopram administration counteracts kainic acid-induced spreading of PSA-NCAM-immunoreactive cells and loss of reelin in the adult mouse hippocampus. Eur J Pharmacol. 2011;666:61–71. doi: 10.1016/j.ejphar.2011.05.008. [DOI] [PubMed] [Google Scholar]
  134. Jakubs K, Nanobashvili A, Bonde S, Ekdahl CT, Kokaia Z, Kokaia M, et al. Environment matters: synaptic properties of neurons born in the epileptic adult brain develop to reduce excitability. Neuron. 2006;52:1047–1059. doi: 10.1016/j.neuron.2006.11.004. [DOI] [PubMed] [Google Scholar]
  135. Janssen DG, Caniato RN, Verster JC, Baune BT. A psychoneuroimmunological review on cytokines involved in antidepressant treatment response. Hum Psychopharmacol. 2010;25:201–215. doi: 10.1002/hup.1103. [DOI] [PubMed] [Google Scholar]
  136. Jensen JB, Jessop DS, Harbuz MS, Mork A, Sanchez C, Mikkelsen JD. Acute and long-term treatments with the selective serotonin reuptake inhibitor citalopram modulate the HPA axis activity at different levels in male rats. J Neuroendocrinol. 1999;11:465–471. doi: 10.1046/j.1365-2826.1999.00362.x. [DOI] [PubMed] [Google Scholar]
  137. Jin Y, Lim CM, Kim SW, Park JY, Seo JS, Han PL, et al. Fluoxetine attenuates kainic acid-induced neuronal cell death in the mouse hippocampus. Brain Res. 2009;1281:108–116. doi: 10.1016/j.brainres.2009.04.053. [DOI] [PubMed] [Google Scholar]
  138. Jobe PC. Common pathogenic mechanisms between depression and epilepsy: an experimental perspective. Epilepsy Behav. 2003;4(Suppl. 3):S14–S24. doi: 10.1016/j.yebeh.2003.08.020. [DOI] [PubMed] [Google Scholar]
  139. Jobe PC, Browning RA. Animal models of depression and epilepsy: the genetically epilepsy-prone rat. In: Ettinger AB, Kanner AM, editors. Psychiatric Issues in Epilepsy. Philadelphia, PA: Lippincott W&W; 2007. pp. 38–48. [Google Scholar]
  140. Jobe PC, Dailey JW, Wernicke JF. A noradrenergic and serotonergic hypothesis of the linkage between epilepsy and affective disorders. Crit Rev Neurobiol. 1999;13:317–356. doi: 10.1615/critrevneurobiol.v13.i4.10. [DOI] [PubMed] [Google Scholar]
  141. Joels M, Baram TZ. The neuro-symphony of stress. Nat Rev Neurosci. 2009;10:459–466. doi: 10.1038/nrn2632. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Jones NC, Cardamone L, Williams JP, Salzberg MR, Myers D, O'Brien TJ. Experimental traumatic brain injury induces a pervasive hyperanxious phenotype in rats. J Neurotrauma. 2008a;25:1367–1374. doi: 10.1089/neu.2008.0641. [DOI] [PubMed] [Google Scholar]
  143. Jones NC, Salzberg MR, Kumar G, Couper A, Morris MJ, O'Brien TJ. Elevated anxiety and depressive-like behavior in a rat model of genetic generalized epilepsy suggesting common causation. Exp Neurol. 2008b;209:254–260. doi: 10.1016/j.expneurol.2007.09.026. [DOI] [PubMed] [Google Scholar]
  144. Jones NC, Kumar G, O'Brien TJ, Morris MJ, Rees SM, Salzberg MR. Anxiolytic effects of rapid amygdala kindling, and the influence of early life experience in rats. Behav Brain Res. 2009;203:81–87. doi: 10.1016/j.bbr.2009.04.023. [DOI] [PubMed] [Google Scholar]
  145. Jones NC, Martin S, Megatia I, Hakami T, Salzberg MR, Pinault D, et al. A genetic epilepsy rat model displays endophenotypes of psychosis. Neurobiol Dis. 2010;39:116–125. doi: 10.1016/j.nbd.2010.02.001. [DOI] [PubMed] [Google Scholar]
  146. Jongsma ME, Bosker FJ, Cremers TI, Westerink BH, den Boer JA. The effect of chronic selective serotonin reuptake inhibitor treatment on serotonin 1B receptor sensitivity and HPA axis activity. Prog Neuropsychopharmacol Biol Psychiatry. 2005;29:738–744. doi: 10.1016/j.pnpbp.2005.04.026. [DOI] [PubMed] [Google Scholar]
  147. Kabuto H, Yokoi I, Endo A, Takei M, Kurimoto T, Mori A. Chronic administration of citalopram inhibited El mouse convulsions and decreased monoamine oxidase-A activity. Acta Med Okayama. 1994a;48:311–316. doi: 10.18926/AMO/31102. [DOI] [PubMed] [Google Scholar]
  148. Kabuto H, Yokoi I, Takei M, Kurimoto T, Mori A. The anticonvulsant effect of citalopram on El mice, and the levels of tryptophan and tyrosine and their metabolites in the brain. Neurochem Res. 1994b;19:463–467. doi: 10.1007/BF00967325. [DOI] [PubMed] [Google Scholar]
  149. Kalynchuk LE. Long-term amygdala kindling in rats as a model for the study of interictal emotionality in temporal lobe epilepsy. Neurosci Biobehav Rev. 2000;24:691–704. doi: 10.1016/s0149-7634(00)00031-2. [DOI] [PubMed] [Google Scholar]
  150. Kanner AM. Psychiatric issues in epilepsy: the complex relation of mood, anxiety disorders, and epilepsy. Epilepsy Behav. 2009;15:83–87. doi: 10.1016/j.yebeh.2009.02.034. [DOI] [PubMed] [Google Scholar]
  151. Kanner AM. Anxiety disorders in epilepsy: the forgotten psychiatric comorbidity. Epilepsy Curr. 2011a;11:90–91. doi: 10.5698/1535-7511-11.3.90. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Kanner AM. Depression and epilepsy: a bidirectional relation? Epilepsia. 2011b;52(Suppl. 1):21–27. doi: 10.1111/j.1528-1167.2010.02907.x. [DOI] [PubMed] [Google Scholar]
  153. Kanner AM, Balabanov A. Depression and epilepsy: how closely related are they? Neurology. 2002;58(8 Suppl. 5):S27–S39. doi: 10.1212/wnl.58.8_suppl_5.s27. [DOI] [PubMed] [Google Scholar]
  154. Kanner AM, Kozak AM, Frey M. The use of sertraline in patients with epilepsy: is it safe? Epilepsy Behav. 2000;1:100–105. doi: 10.1006/ebeh.2000.0050. [DOI] [PubMed] [Google Scholar]
  155. Kanner AM, Byrne R, Chicharro A, Wuu J, Frey M. A lifetime psychiatric history predicts a worse seizure outcome following temporal lobectomy. Neurology. 2009;72:793–799. doi: 10.1212/01.wnl.0000343850.85763.9c. [DOI] [PubMed] [Google Scholar]
  156. Kanner AM, Barry JJ, Gilliam F, Hermann B, Meador KJ. Anxiety disorders, subsyndromic depressive episodes, and major depressive episodes: do they differ on their impact on the quality of life of patients with epilepsy? Epilepsia. 2010;51:1152–1158. doi: 10.1111/j.1528-1167.2010.02582.x. [DOI] [PubMed] [Google Scholar]
  157. Kanner AM, Barry JJ, Gilliam F, Hermann B, Meador KJ. Depressive and anxiety disorders in epilepsy: do they differ in their potential to worsen common antiepileptic drug-related adverse events? Epilepsia. 2012;53:1104–1108. doi: 10.1111/j.1528-1167.2012.03488.x. [DOI] [PubMed] [Google Scholar]
  158. Karege F, Perret G, Bondolfi G, Schwald M, Bertschy G, Aubry JM. Decreased serum brain-derived neurotrophic factor levels in major depressed patients. Psychiatry Res. 2002;109:143–148. doi: 10.1016/s0165-1781(02)00005-7. [DOI] [PubMed] [Google Scholar]
  159. Karpova NN, Pickenhagen A, Lindholm J, Tiraboschi E, Kulesskaya N, Agustsdottir A, et al. Fear erasure in mice requires synergy between antidepressant drugs and extinction training. Science. 2011;334:1731–1734. doi: 10.1126/science.1214592. [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Karst H, de Kloet ER, Joels M. Episodic corticosterone treatment accelerates kindling epileptogenesis and triggers long-term changes in hippocampal CA1 cells, in the fully kindled state. Eur J Neurosci. 1999;11:889–898. doi: 10.1046/j.1460-9568.1999.00495.x. [DOI] [PubMed] [Google Scholar]
  161. Kecskemeti V, Rusznak Z, Riba P, Pal B, Wagner R, Harasztosi C, et al. Norfluoxetine and fluoxetine have similar anticonvulsant and Ca2+ channel blocking potencies. Brain Res Bull. 2005;67:126–132. doi: 10.1016/j.brainresbull.2005.06.027. [DOI] [PubMed] [Google Scholar]
  162. Kelley MS, Jacobs MP, Lowenstein DH. The NINDS epilepsy research benchmarks. Epilepsia. 2009;50:579–582. doi: 10.1111/j.1528-1167.2008.01813.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Kenis G, Maes M. Effects of antidepressants on the production of cytokines. Int J Neuropsychopharmacol. 2002;5:401–412. doi: 10.1017/S1461145702003164. [DOI] [PubMed] [Google Scholar]
  164. Kobayashi K, Ikeda Y, Sakai A, Yamasaki N, Haneda E, Miyakawa T, et al. Reversal of hippocampal neuronal maturation by serotonergic antidepressants. Proc Natl Acad Sci U S A. 2010;107:8434–8439. doi: 10.1073/pnas.0912690107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Koe AS, Jones NC, Salzberg MR. Early life stress as an influence on limbic epilepsy: an hypothesis whose time has come? Front Behav Neurosci. 2009;3:24. doi: 10.3389/neuro.08.024.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Koella WP, Glatt A, Klebs K, Durst T. Epileptic phenomena induced in the cat by the antidepressants maprotiline, imipramine, clomipramine, and amitriptyline. Biol Psychiatry. 1979;14:485–497. [PubMed] [Google Scholar]
  167. Koh S, Magid R, Chung H, Stine CD, Wilson DN. Depressive behavior and selective down-regulation of serotonin receptor expression after early-life seizures: reversal by environmental enrichment. Epilepsy Behav. 2007;10:26–31. doi: 10.1016/j.yebeh.2006.11.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Koyama R, Ikegaya Y. To BDNF or not to BDNF: that is the epileptic hippocampus. Neuroscientist. 2005;11:282–287. doi: 10.1177/1073858405278266. [DOI] [PubMed] [Google Scholar]
  169. Koyama R, Yamada MK, Fujisawa S, Katoh-Semba R, Matsuki N, Ikegaya Y. Brain-derived neurotrophic factor induces hyperexcitable reentrant circuits in the dentate gyrus. J Neurosci. 2004;24:7215–7224. doi: 10.1523/JNEUROSCI.2045-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. Krijzer F, Snelder M, Bradford D. Comparison of the (pro)convulsive properties of fluvoxamine and clovoxamine with eight other antidepressants in an animal model. Neuropsychobiology. 1984;12:249–254. doi: 10.1159/000118147. [DOI] [PubMed] [Google Scholar]
  171. Krishnan KR. Revisiting monoamine oxidase inhibitors. J Clin Psychiatry. 2007;68(Suppl. 8):35–41. [PubMed] [Google Scholar]
  172. Kron MM, Zhang H, Parent JM. The developmental stage of dentate granule cells dictates their contribution to seizure-induced plasticity. J Neurosci. 2010;30:2051–2059. doi: 10.1523/JNEUROSCI.5655-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Krystal JH, Sanacora G, Blumberg H, Anand A, Charney DS, Marek G, et al. Glutamate and GABA systems as targets for novel antidepressant and mood-stabilizing treatments. Mol Psychiatry. 2002;7(Suppl. 1):S71–S80. doi: 10.1038/sj.mp.4001021. [DOI] [PubMed] [Google Scholar]
  174. Kuhn KU, Quednow BB, Thiel M, Falkai P, Maier W, Elger CE. Antidepressive treatment in patients with temporal lobe epilepsy and major depression: a prospective study with three different antidepressants. Epilepsy Behav. 2003;4:674–679. doi: 10.1016/j.yebeh.2003.08.009. [DOI] [PubMed] [Google Scholar]
  175. Kumar G, Couper A, O'Brien TJ, Salzberg MR, Jones NC, Rees SM, et al. The acceleration of amygdala kindling epileptogenesis by chronic low-dose corticosterone involves both mineralocorticoid and glucocorticoid receptors. Psychoneuroendocrinology. 2007;32:834–842. doi: 10.1016/j.psyneuen.2007.05.011. [DOI] [PubMed] [Google Scholar]
  176. Kumar G, Jones NC, Morris MJ, Rees S, O'Brien TJ, Salzberg MR. Early life stress enhancement of limbic epileptogenesis in adult rats: mechanistic insights. Plos One. 2011;6:e24033. doi: 10.1371/journal.pone.0024033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Labate A, Cerasa A, Aguglia U, Mumoli L, Quattrone A, Gambardella A. Neocortical thinning in ‘benign’ mesial temporal lobe epilepsy. Epilepsia. 2011;52:712–717. doi: 10.1111/j.1528-1167.2011.03038.x. [DOI] [PubMed] [Google Scholar]
  178. Lacey CJ, Salzberg MR, Roberts H, Trauer T, D'Souza WJ. Psychiatric comorbidity and impact on health service utilization in a community sample of patients with epilepsy. Epilepsia. 2009;50:1991–1994. doi: 10.1111/j.1528-1167.2009.02165.x. [DOI] [PubMed] [Google Scholar]
  179. Lanquillon S, Krieg JC, Bening-Abu-Shach U, Vedder H. Cytokine production and treatment response in major depressive disorder. Neuropsychopharmacology. 2000;22:370–379. doi: 10.1016/S0893-133X(99)00134-7. [DOI] [PubMed] [Google Scholar]
  180. Larmet Y, Reibel S, Carnahan J, Nawa H, Marescaux C, Depaulis A. Protective effects of brain-derived neurotrophic factor on the development of hippocampal kindling in the rat. Neuroreport. 1995;6:1937–1941. doi: 10.1097/00001756-199510020-00027. [DOI] [PubMed] [Google Scholar]
  181. Lee BH, Kim YK. The roles of BDNF in the pathophysiology of major depression and in antidepressant treatment. Psychiatry Investig. 2010;7:231–235. doi: 10.4306/pi.2010.7.4.231. [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Lee HM, Hahn SJ, Choi BH. Open channel block of Kv1.5 currents by citalopram. Acta Pharmacol Sin. 2010;31:429–435. doi: 10.1038/aps.2010.14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Li N, Lee B, Liu RJ, Banasr M, Dwyer JM, Iwata M, et al. mTOR-dependent synapse formation underlies the rapid antidepressant effects of NMDA antagonists. Science. 2010;329:959–964. doi: 10.1126/science.1190287. [DOI] [PMC free article] [PubMed] [Google Scholar]
  184. Li Y, Luikart BW, Birnbaum S, Chen J, Kwon CH, Kernie SG, et al. TrkB regulates hippocampal neurogenesis and governs sensitivity to antidepressive treatment. Neuron. 2008;59:399–412. doi: 10.1016/j.neuron.2008.06.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Lin JJ, Mula M, Hermann BP. Uncovering the neurobehavioural comorbidities of epilepsy over the lifespan. Lancet. 2012;380:1180–1192. doi: 10.1016/S0140-6736(12)61455-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Lothe A, Didelot A, Hammers A, Costes N, Saoud M, Gilliam F, et al. Comorbidity between temporal lobe epilepsy and depression: a [18F]MPPF PET study. Brain. 2008;131(Pt 10):2765–2782. doi: 10.1093/brain/awn194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  187. Lu KT, Gean PW. Endogenous serotonin inhibits epileptiform activity in rat hippocampal CA1 neurons via 5-hydroxytryptamine1A receptor activation. Neuroscience. 1998;86:729–737. doi: 10.1016/s0306-4522(98)00106-7. [DOI] [PubMed] [Google Scholar]
  188. Luchins DJ, Oliver AP, Wyatt RJ. Seizures with antidepressants: an in vitro technique to assess relative risk. Epilepsia. 1984;25:25–32. doi: 10.1111/j.1528-1157.1984.tb04151.x. [DOI] [PubMed] [Google Scholar]
  189. Lupien SJ, McEwen BS, Gunnar MR, Heim C. Effects of stress throughout the lifespan on the brain, behaviour and cognition. Nat Rev Neurosci. 2009;10:434–445. doi: 10.1038/nrn2639. [DOI] [PubMed] [Google Scholar]
  190. McCloskey DP, Hintz TM, Pierce JP, Scharfman HE. Stereological methods reveal the robust size and stability of ectopic hilar granule cells after pilocarpine-induced status epilepticus in the adult rat. Eur J Neurosci. 2006;24:2203–2210. doi: 10.1111/j.1460-9568.2006.05101.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  191. McLaughlin DP, Pachana NA, McFarland K. Depression in a community-dwelling sample of older adults with late-onset or lifetime epilepsy. Epilepsy Behav. 2008;12:281–285. doi: 10.1016/j.yebeh.2007.10.005. [DOI] [PubMed] [Google Scholar]
  192. Macrodimitris S, Wershler J, Hatfield M, Hamilton K, Backs-Dermott B, Mothersill K, et al. Group cognitive-behavioral therapy for patients with epilepsy and comorbid depression and anxiety. Epilepsy Behav. 2011;20:83–88. doi: 10.1016/j.yebeh.2010.10.028. [DOI] [PubMed] [Google Scholar]
  193. Maes M, Meltzer HY, Bosmans E, Bergmans R, Vandoolaeghe E, Ranjan R, et al. Increased plasma concentrations of interleukin-6, soluble interleukin-6, soluble interleukin-2 and transferrin receptor in major depression. J Affect Disord. 1995;34:301–309. doi: 10.1016/0165-0327(95)00028-l. [DOI] [PubMed] [Google Scholar]
  194. Mainie I, McGurk C, McClintock G, Robinson J. Seizures after buproprion overdose. Lancet. 2001;357:1624. doi: 10.1016/S0140-6736(00)04770-X. [DOI] [PubMed] [Google Scholar]
  195. Malberg JE, Eisch AJ, Nestler EJ, Duman RS. Chronic antidepressant treatment increases neurogenesis in adult rat hippocampus. J Neurosci. 2000;20:9104–9110. doi: 10.1523/JNEUROSCI.20-24-09104.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  196. Maroso M, Balosso S, Ravizza T, Liu J, Aronica E, Iyer AM, et al. Toll-like receptor 4 and high-mobility group box-1 are involved in ictogenesis and can be targeted to reduce seizures. Nat Med. 2010;16:413–419. doi: 10.1038/nm.2127. [DOI] [PubMed] [Google Scholar]
  197. Mathern GW, Babb TL, Micevych PE, Blanco CE, Pretorius JK. Granule cell mRNA levels for BDNF, NGF, and NT-3 correlate with neuron losses or supragranular mossy fiber sprouting in the chronically damaged and epileptic human hippocampus. Mol Chem Neuropathol. 1997;30:53–76. doi: 10.1007/BF02815150. [DOI] [PubMed] [Google Scholar]
  198. Mathern GW, Pretorius JK, Mendoza D, Lozada A, Kornblum HI. Hippocampal AMPA and NMDA mRNA levels correlate with aberrant fascia dentata mossy fiber sprouting in the pilocarpine model of spontaneous limbic epilepsy. J Neurosci Res. 1998;54:734–753. doi: 10.1002/(SICI)1097-4547(19981215)54:6<734::AID-JNR2>3.0.CO;2-P. [DOI] [PubMed] [Google Scholar]
  199. Mazarati A, Shin D, Auvin S, Caplan R, Sankar R. Kindling epileptogenesis in immature rats leads to persistent depressive behavior. Epilepsy Behav. 2007;10:377–383. doi: 10.1016/j.yebeh.2007.02.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  200. Mazarati A, Siddarth P, Baldwin RA, Shin D, Caplan R, Sankar R. Depression after status epilepticus: behavioural and biochemical deficits and effects of fluoxetine. Brain. 2008;131(Pt 8):2071–2083. doi: 10.1093/brain/awn117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  201. Mazarati AM, Shin D, Kwon YS, Bragin A, Pineda E, Tio D, et al. Elevated plasma corticosterone level and depressive behavior in experimental temporal lobe epilepsy. Neurobiol Dis. 2009;34:457–461. doi: 10.1016/j.nbd.2009.02.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  202. Mazarati AM, Pineda E, Shin D, Tio D, Taylor AN, Sankar R. Comorbidity between epilepsy and depression: role of hippocampal interleukin-1beta. Neurobiol Dis. 2010;37:461–467. doi: 10.1016/j.nbd.2009.11.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  203. Mensah SA, Beavis JM, Thapar AK, Kerr MP. A community study of the presence of anxiety disorder in people with epilepsy. Epilepsy Behav. 2007;11:118–124. doi: 10.1016/j.yebeh.2007.04.012. [DOI] [PubMed] [Google Scholar]
  204. Mesquita AR, Tavares HB, Silva R, Sousa N. Febrile convulsions in developing rats induce a hyperanxious phenotype later in life. Epilepsy Behav. 2006;9:401–406. doi: 10.1016/j.yebeh.2006.07.012. [DOI] [PubMed] [Google Scholar]
  205. Metternich B, Wagner K, Brandt A, Kraemer R, Buschmann F, Zentner J, et al. Preoperative depressive symptoms predict postoperative seizure outcome in temporal and frontal lobe epilepsy. Epilepsy Behav. 2009;16:622–628. doi: 10.1016/j.yebeh.2009.09.017. [DOI] [PubMed] [Google Scholar]
  206. Mikova O, Yakimova R, Bosmans E, Kenis G, Maes M. Increased serum tumor necrosis factor alpha concentrations in major depression and multiple sclerosis. Eur Neuropsychopharmacol. 2001;11:203–208. doi: 10.1016/s0924-977x(01)00081-5. [DOI] [PubMed] [Google Scholar]
  207. Miller BH, Schultz LE, Gulati A, Cameron MD, Pletcher MT. Genetic regulation of behavioral and neuronal responses to fluoxetine. Neuropsychopharmacology. 2008;33:1312–1322. doi: 10.1038/sj.npp.1301497. [DOI] [PubMed] [Google Scholar]
  208. Minami M, Kuraishi Y, Satoh M. Effects of kainic acid on messenger RNA levels of IL-1 beta, IL-6, TNF alpha and LIF in the rat brain. Biochem Biophys Res Commun. 1991;176:593–598. doi: 10.1016/s0006-291x(05)80225-6. [DOI] [PubMed] [Google Scholar]
  209. Mirescu C, Gould E. Stress and adult neurogenesis. Hippocampus. 2006;16:233–238. doi: 10.1002/hipo.20155. [DOI] [PubMed] [Google Scholar]
  210. Miro X, Perez-Torres S, Artigas F, Puigdomenech P, Palacios JM, Mengod G. Regulation of cAMP phosphodiesterase mRNAs expression in rat brain by acute and chronic fluoxetine treatment. An in situ hybridization study. Neuropharmacology. 2002;43:1148–1157. doi: 10.1016/s0028-3908(02)00220-4. [DOI] [PubMed] [Google Scholar]
  211. Morgan VA, Croft ML, Valuri GM, Zubrick SR, Bower C, McNeil TF, et al. Intellectual disability and other neuropsychiatric outcomes in high-risk children of mothers with schizophrenia, bipolar disorder and unipolar major depression. Br J Psychiatry. 2012;200:282–289. doi: 10.1192/bjp.bp.111.093070. [DOI] [PubMed] [Google Scholar]
  212. Morimoto K, Fahnestock M, Racine RJ. Kindling and status epilepticus models of epilepsy: rewiring the brain. Prog Neurobiol. 2004;73:1–60. doi: 10.1016/j.pneurobio.2004.03.009. [DOI] [PubMed] [Google Scholar]
  213. Mula M, Monaco F. Antiepileptic drugs and psychopathology of epilepsy: an update. Epileptic Disord. 2009a;11:1–9. doi: 10.1684/epd.2009.0238. [DOI] [PubMed] [Google Scholar]
  214. Mula M, Schmitz B. Depression in epilepsy: mechanisms and therapeutic approach. Ther Adv Neurol Disord. 2009b;2:337–344. doi: 10.1177/1756285609337340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  215. Mula M, Schmitz B, Sander JW. The pharmacological treatment of depression in adults with epilepsy. Expert Opin Pharmacother. 2008;9:3159–3168. doi: 10.1517/14656560802587024. [DOI] [PubMed] [Google Scholar]
  216. Mula M, Jauch R, Cavanna A, Gaus V, Kretz R, Collimedaglia L, et al. Interictal dysphoric disorder and periictal dysphoric symptoms in patients with epilepsy. Epilepsia. 2010;51:1139–1145. doi: 10.1111/j.1528-1167.2009.02424.x. [DOI] [PubMed] [Google Scholar]
  217. Muller CJ, Bankstahl M, Groticke I, Loscher W. Pilocarpine vs. lithium-pilocarpine for induction of status epilepticus in mice: development of spontaneous seizures, behavioral alterations and neuronal damage. Eur J Pharmacol. 2009a;619:15–24. doi: 10.1016/j.ejphar.2009.07.020. [DOI] [PubMed] [Google Scholar]
  218. Muller CJ, Groticke I, Bankstahl M, Loscher W. Behavioral and cognitive alterations, spontaneous seizures, and neuropathology developing after a pilocarpine-induced status epilepticus in C57BL/6 mice. Exp Neurol. 2009b;219:284–297. doi: 10.1016/j.expneurol.2009.05.035. [DOI] [PubMed] [Google Scholar]
  219. Murphy BL, Pun RY, Yin H, Faulkner CR, Loepke AW, Danzer SC. Heterogeneous integration of adult-generated granule cells into the epileptic brain. J Neurosci. 2011;31:105–117. doi: 10.1523/JNEUROSCI.2728-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  220. Murray KD, Isackson PJ, Eskin TA, King MA, Montesinos SP, Abraham LA, et al. Altered mRNA expression for brain-derived neurotrophic factor and type II calcium/calmodulin-dependent protein kinase in the hippocampus of patients with intractable temporal lobe epilepsy. J Comp Neurol. 2000;418:411–422. doi: 10.1002/(sici)1096-9861(20000320)418:4<411::aid-cne4>3.0.co;2-f. [DOI] [PubMed] [Google Scholar]
  221. Najjar S, Bernbaum M, Lai G, Devinsky O. Immunology and epilepsy. Rev Neurol Dis. 2008;5:109–116. [PubMed] [Google Scholar]
  222. Nestler EJ, Hyman SE, Malenka RC. Molecular Neuropharmacology. New York: McGraw Hill; 2001. [Google Scholar]
  223. Nestler EJ, Barrot M, DiLeone RJ, Eisch AJ, Gold SJ, Monteggia LM. Neurobiology of depression. Neuron. 2002;34:13–25. doi: 10.1016/s0896-6273(02)00653-0. [DOI] [PubMed] [Google Scholar]
  224. Nibuya M, Morinobu S, Duman RS. Regulation of BDNF and trkB mRNA in rat brain by chronic electroconvulsive seizure and antidepressant drug treatments. J Neurosci. 1995;15:7539–7547. doi: 10.1523/JNEUROSCI.15-11-07539.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  225. Nilsson FM, Kessing LV, Bolwig TG. On the increased risk of developing late-onset epilepsy for patients with major affective disorder. J Affect Disord. 2003;76:39–48. doi: 10.1016/s0165-0327(02)00061-7. [DOI] [PubMed] [Google Scholar]
  226. Niquet J, Ben-Ari Y, Represa A. Glial reaction after seizure induced hippocampal lesion: immunohistochemical characterization of proliferating glial cells. J Neurocytol. 1994;23:641–656. doi: 10.1007/BF01191558. [DOI] [PubMed] [Google Scholar]
  227. Noe KH, Locke DE, Sirven JI. Treatment of depression in patients with epilepsy. Curr Treat Options Neurol. 2011;13:371–379. doi: 10.1007/s11940-011-0127-8. [DOI] [PubMed] [Google Scholar]
  228. Okazaki M, Adachi N, Ito M, Watanabe M, Watanabe Y, Kato M, et al. One-year seizure prognosis in epilepsy patients treated with antidepressants. Epilepsy Behav. 2011;22:331–335. doi: 10.1016/j.yebeh.2011.07.016. [DOI] [PubMed] [Google Scholar]
  229. Olsen RW, DeLorey TM, Gordey M, Kang MH. GABA receptor function and epilepsy. Adv Neurol. 1999;79:499–510. [PubMed] [Google Scholar]
  230. Osehobo P, Adams B, Sazgar M, Xu Y, Racine RJ, Fahnestock M. Brain-derived neurotrophic factor infusion delays amygdala and perforant path kindling without affecting paired-pulse measures of neuronal inhibition in adult rats. Neuroscience. 1999;92:1367–1375. doi: 10.1016/s0306-4522(99)00048-2. [DOI] [PubMed] [Google Scholar]
  231. Ottman R, Lipton RB, Ettinger AB, Cramer JA, Reed ML, Morrison A, et al. Comorbidities of epilepsy: results from the Epilepsy Comorbidities and Health (EPIC) survey. Epilepsia. 2011;52:308–315. doi: 10.1111/j.1528-1167.2010.02927.x. [DOI] [PubMed] [Google Scholar]
  232. Owen BM, Eccleston D, Ferrier IN, Young AH. Raised levels of plasma interleukin-1beta in major and postviral depression. Acta Psychiatr Scand. 2001;103:226–228. doi: 10.1034/j.1600-0447.2001.00162.x. [DOI] [PubMed] [Google Scholar]
  233. Pancrazio JJ, Kamatchi GL, Roscoe AK, Lynch C., 3rd Inhibition of neuronal Na+ channels by antidepressant drugs. J Pharmacol Exp Ther. 1998;284:208–214. [PubMed] [Google Scholar]
  234. Panelli RJ, Kilpatrick C, Moore SM, Matkovic Z, D'Souza WJ, O'Brien TJ. The Liverpool Adverse Events Profile: relation to AED use and mood. Epilepsia. 2007;48:456–463. doi: 10.1111/j.1528-1167.2006.00956.x. [DOI] [PubMed] [Google Scholar]
  235. Paparrigopoulos T, Ferentinos P, Brierley B, Shaw P, David AS. Relationship between post-operative depression/anxiety and hippocampal/amygdala volumes in temporal lobectomy for epilepsy. Epilepsy Res. 2008;81:30–35. doi: 10.1016/j.eplepsyres.2008.04.011. [DOI] [PubMed] [Google Scholar]
  236. Parent JM, Yu TW, Leibowitz RT, Geschwind DH, Sloviter RS, Lowenstein DH. Dentate granule cell neurogenesis is increased by seizures and contributes to aberrant network reorganization in the adult rat hippocampus. J Neurosci. 1997;17:3727–3738. doi: 10.1523/JNEUROSCI.17-10-03727.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  237. Parent JM, Elliott RC, Pleasure SJ, Barbaro NM, Lowenstein DH. Aberrant seizure-induced neurogenesis in experimental temporal lobe epilepsy. Ann Neurol. 2006;59:81–91. doi: 10.1002/ana.20699. [DOI] [PubMed] [Google Scholar]
  238. Pariante CM. The glucocorticoid receptor: part of the solution or part of the problem? J Psychopharmacol. 2006;20(4 Suppl):79–84. doi: 10.1177/1359786806066063. [DOI] [PubMed] [Google Scholar]
  239. Pariante CM, Lightman SL. The HPA axis in major depression: classical theories and new developments. Trends Neurosci. 2008;31:464–468. doi: 10.1016/j.tins.2008.06.006. [DOI] [PubMed] [Google Scholar]
  240. Parsons LH, Kerr TM, Tecott LH. 5-HT(1A) receptor mutant mice exhibit enhanced tonic, stress-induced and fluoxetine-induced serotonergic neurotransmission. J Neurochem. 2001;77:607–617. doi: 10.1046/j.1471-4159.2001.00254.x. [DOI] [PubMed] [Google Scholar]
  241. Pasini A, Tortorella A, Gale K. Anticonvulsant effect of intranigral fluoxetine. Brain Res. 1992;593:287–290. doi: 10.1016/0006-8993(92)91320-e. [DOI] [PubMed] [Google Scholar]
  242. Payandemehr B, Bahremand A, Rahimian R, Ziai P, Amouzegar A, Sharifzadeh M, et al. 5-HT(3) receptor mediates the dose-dependent effects of citalopram on pentylenetetrazole-induced clonic seizure in mice: involvement of nitric oxide. Epilepsy Res. 2012;101:217–227. doi: 10.1016/j.eplepsyres.2012.04.004. [DOI] [PubMed] [Google Scholar]
  243. Pepin MC, Beaulieu S, Barden N. Antidepressants regulate glucocorticoid receptor messenger RNA concentrations in primary neuronal cultures. Brain Res Mol Brain Res. 1989;6:77–83. doi: 10.1016/0169-328x(89)90031-4. [DOI] [PubMed] [Google Scholar]
  244. Pericic D, Lazic J, Svob Strac D. Anticonvulsant effects of acute and repeated fluoxetine treatment in unstressed and stressed mice. Brain Res. 2005;1033:90–95. doi: 10.1016/j.brainres.2004.11.025. [DOI] [PubMed] [Google Scholar]
  245. Perr JV, Ettinger AB. Psychiatric illness and psychotropic medication use in epilepsy. In: Trimble M, Schmitz B, editors. The Neuropsychiatry of Epilepsy. Vol. 16. Cambridge: Cambridge University Press; 2011. pp. 165–195. [Google Scholar]
  246. Petrovski S, Szoeke CE, Jones NC, Salzberg MR, Sheffield LJ, Huggins RM, et al. Neuropsychiatric symptomatology predicts seizure recurrence in newly treated patients. Neurology. 2010;75:1015–1021. doi: 10.1212/WNL.0b013e3181f25b16. [DOI] [PubMed] [Google Scholar]
  247. Piletz JE, Halaris A, Iqbal O, Hoppensteadt D, Fareed J, Zhu H, et al. Pro-inflammatory biomakers in depression: treatment with venlafaxine. World J Biol Psychiatry. 2009;10:313–323. doi: 10.3109/15622970802573246. [DOI] [PubMed] [Google Scholar]
  248. Pineda EA, Hensler JG, Sankar R, Shin D, Burke TF, Mazarati AM. Interleukin-1beta causes fluoxetine resistance in an animal model of epilepsy-associated depression. Neurother. 2012;9:477–485. doi: 10.1007/s13311-012-0110-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  249. Pisani F, Spina E, Oteri G. Antidepressant drugs and seizure susceptibility: from in vitro data to clinical practice. Epilepsia. 1999;40(Suppl. 10):S48–S56. doi: 10.1111/j.1528-1157.1999.tb00885.x. [DOI] [PubMed] [Google Scholar]
  250. Pisani F, Oteri G, Costa C, Di Raimondo G, Di Perri R. Effects of psychotropic drugs on seizure threshold. Drug Saf. 2002;25:91–110. doi: 10.2165/00002018-200225020-00004. [DOI] [PubMed] [Google Scholar]
  251. Popoli M, Gennarelli M, Racagni G. Modulation of synaptic plasticity by stress and antidepressants. Bipolar Disord. 2002;4:166–182. doi: 10.1034/j.1399-5618.2002.01159.x. [DOI] [PubMed] [Google Scholar]
  252. Post RM. Do the epilepsies, pain syndromes, and affective disorders share common kindling-like mechanisms? Epilepsy Res. 2002;50:203–219. doi: 10.1016/s0920-1211(02)00081-5. [DOI] [PubMed] [Google Scholar]
  253. Prendiville S, Gale K. Anticonvulsant effect of fluoxetine on focally evoked limbic motor seizures in rats. Epilepsia. 1993;34:381–384. doi: 10.1111/j.1528-1157.1993.tb02425.x. [DOI] [PubMed] [Google Scholar]
  254. Preskorn SH, Fast GA. Tricyclic antidepressant-induced seizures and plasma drug concentration. J Clin Psychiatry. 1992;53:160–162. [PubMed] [Google Scholar]
  255. Pritchard PB, III, Wannamaker BB, Sagel J, Daniel CM. Serum prolactin and cortisol levels in evaluation of pseudoepileptic seizures. Ann Neurol. 1985;18:87–89. doi: 10.1002/ana.410180115. [DOI] [PubMed] [Google Scholar]
  256. Raju SS, Noor AR, Gurthu S, Giriyappanavar CR, Acharya SB, Low HC, et al. Effect of fluoxetine on maximal electroshock seizures in mice: acute vs chronic administration. Pharmacol Res. 1999;39:451–454. doi: 10.1006/phrs.1999.0466. [DOI] [PubMed] [Google Scholar]
  257. Rao ML, Stefan H, Bauer J. Epileptic but not psychogenic seizures are accompanied by simultaneous elevation of serum pituitary hormones and cortisol levels. Neuroendocrinology. 1989;49:33–39. doi: 10.1159/000125088. [DOI] [PubMed] [Google Scholar]
  258. Reibel S, Larmet Y, Carnahan J, Marescaux C, Depaulis A. Endogenous control of hippocampal epileptogenesis: a molecular cascade involving brain-derived neurotrophic factor and neuropeptide Y. Epilepsia. 2000a;41(Suppl. 6):S127–S133. doi: 10.1111/j.1528-1157.2000.tb01571.x. [DOI] [PubMed] [Google Scholar]
  259. Reibel S, Larmet Y, Le BT, Carnahan J, Marescaux C, Depaulis A. Brain-derived neurotrophic factor delays hippocampal kindling in the rat. Neuroscience. 2000b;100:777–788. doi: 10.1016/s0306-4522(00)00351-1. [DOI] [PubMed] [Google Scholar]
  260. Reul JM, Stec I, Soder M, Holsboer F. Chronic treatment of rats with the antidepressant amitriptyline attenuates the activity of the hypothalamic-pituitary-adrenocortical system. Endocrinology. 1993;133:312–320. doi: 10.1210/endo.133.1.8391426. [DOI] [PubMed] [Google Scholar]
  261. Reul JM, Labeur MS, Grigoriadis DE, De Souza EB, Holsboer F. Hypothalamic-pituitary-adrenocortical axis changes in the rat after long-term treatment with the reversible monoamine oxidase-A inhibitor moclobemide. Neuroendocrinology. 1994;60:509–519. doi: 10.1159/000126788. [DOI] [PubMed] [Google Scholar]
  262. Richardson EJ, Griffith HR, Martin RC, Paige AL, Stewart CC, Jones J, et al. Structural and functional neuroimaging correlates of depression in temporal lobe epilepsy. Epilepsy Behav. 2007;10:242–249. doi: 10.1016/j.yebeh.2006.11.013. [DOI] [PubMed] [Google Scholar]
  263. Richman A, Heinrichs SC. Seizure prophylaxis in an animal model of epilepsy by dietary fluoxetine supplementation. Epilepsy Res. 2007;74:19–27. doi: 10.1016/j.eplepsyres.2006.11.007. [DOI] [PubMed] [Google Scholar]
  264. Ridsdale L, Charlton J, Ashworth M, Richardson MP, Gulliford MC. Epilepsy mortality and risk factors for death in epilepsy: a population-based study. Br J Gen Pract. 2011;61:e271–e278. doi: 10.3399/bjgp11X572463. [DOI] [PMC free article] [PubMed] [Google Scholar]
  265. Robertson MM, Trimble MR, Townsend HR. Phenomenology of depression in epilepsy. Epilepsia. 1987;28:364–372. doi: 10.1111/j.1528-1157.1987.tb03659.x. [DOI] [PubMed] [Google Scholar]
  266. Rudge JS, Mather PE, Pasnikowski EM, Cai N, Corcoran T, Acheson A, et al. Endogenous BDNF protein is increased in adult rat hippocampus after a kainic acid induced excitotoxic insult but exogenous BDNF is not neuroprotective. Exp Neurol. 1998;149:398–410. doi: 10.1006/exnr.1997.6737. [DOI] [PubMed] [Google Scholar]
  267. Russo-Neustadt AA, Beard RC, Huang YM, Cotman CW. Physical activity and antidepressant treatment potentiate the expression of specific brain-derived neurotrophic factor transcripts in the rat hippocampus. Neuroscience. 2000;101:305–312. doi: 10.1016/s0306-4522(00)00349-3. [DOI] [PubMed] [Google Scholar]
  268. Saarelainen T, Hendolin P, Lucas G, Koponen E, Sairanen M, MacDonald E, et al. Activation of the TrkB neurotrophin receptor is induced by antidepressant drugs and is required for antidepressant-induced behavioral effects. J Neurosci. 2003;23:349–357. doi: 10.1523/JNEUROSCI.23-01-00349.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  269. Salgado PC, Yasuda CL, Cendes F. Neuroimaging changes in mesial temporal lobe epilepsy are magnified in the presence of depression. Epilepsy Behav. 2010;19:422–427. doi: 10.1016/j.yebeh.2010.08.012. [DOI] [PubMed] [Google Scholar]
  270. Salzberg M. Neurobiological links between epilepsy and mood disorders. Neurology Asia. 2011;16:37–40. [Google Scholar]
  271. Salzberg M, Kumar G, Supit L, Jones NC, Morris MJ, Rees S, et al. Early postnatal stress confers enduring vulnerability to limbic epileptogenesis. Epilepsia. 2007;48:2079–2085. doi: 10.1111/j.1528-1167.2007.01246.x. [DOI] [PubMed] [Google Scholar]
  272. Salzberg MR, Vajda FJ. Epilepsy, depression and antidepressant drugs. J Clin Neurosci. 2001;8:209–215. doi: 10.1054/jocn.2000.0896. [DOI] [PubMed] [Google Scholar]
  273. Santarelli L, Saxe M, Gross C, Surget A, Battaglia F, Dulawa S, et al. Requirement of hippocampal neurogenesis for the behavioral effects of antidepressants. Science. 2003;301:805–809. doi: 10.1126/science.1083328. [DOI] [PubMed] [Google Scholar]
  274. Santos JG, Jr, Do Monte FH, Russi M, Agustine PE, Lanziotti VM. Proconvulsant effects of high doses of venlafaxine in pentylenetetrazole-convulsive rats. Braz J Med Biol Res. 2002;35:469–472. doi: 10.1590/s0100-879x2002000400010. [DOI] [PubMed] [Google Scholar]
  275. Sarkisova K, van Luijtelaar G. The WAG/Rij strain: a genetic animal model of absence epilepsy with comorbidity of depression. Prog Neuropsychopharmacol Biol Psychiatry. 2011;35:854–876. doi: 10.1016/j.pnpbp.2010.11.010. [corrected] [DOI] [PubMed] [Google Scholar]
  276. Sarnyai Z, Sibille EL, Pavlides C, Fenster RJ, McEwen BS, Toth M. Impaired hippocampal-dependent learning and functional abnormalities in the hippocampus in mice lacking serotonin(1A) receptors. Proc Natl Acad Sci U S A. 2000;97:14731–14736. doi: 10.1073/pnas.97.26.14731. [DOI] [PMC free article] [PubMed] [Google Scholar]
  277. Sayyah M, Javad-Pour M, Ghazi-Khansari M. The bacterial endotoxin lipopolysaccharide enhances seizure susceptibility in mice: involvement of proinflammatory factors: nitric oxide and prostaglandins. Neuroscience. 2003;122:1073–1080. doi: 10.1016/j.neuroscience.2003.08.043. [DOI] [PubMed] [Google Scholar]
  278. Scharfman HE. The neurobiology of epilepsy. Curr Neurol Neurosci Rep. 2007;7:348–354. doi: 10.1007/s11910-007-0053-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  279. Scharfman H, Goodman J, McCloskey D. Ectopic granule cells of the rat dentate gyrus. Dev Neurosci. 2007;29:14–27. doi: 10.1159/000096208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  280. Scharfman HE, Goodman JH, Sollas AL, Croll SD. Spontaneous limbic seizures after intrahippocampal infusion of brain-derived neurotrophic factor. Exp Neurol. 2002;174:201–214. doi: 10.1006/exnr.2002.7869. [DOI] [PubMed] [Google Scholar]
  281. Schmitz B. The effects of antiepileptic drugs on behavior. In: Trimble M, Schmitz B, editors. The Neuropsychiatry of Epilepsy. 2nd edn. Cambridge: Cambridge University Press; 2011. pp. 133–142. [Google Scholar]
  282. Settle EC, Stahl SM, Batey SR, Johnston JA, Ascher JA. Safety profile of sustained-release bupropion in depression: results of three clinical trials. Clin Ther. 1999;21:454–463. doi: 10.1016/s0149-2918(00)88301-0. [DOI] [PubMed] [Google Scholar]
  283. Shin RS, McIntyre DC. Differential noradrenergic influence on seizure expression in genetically Fast and Slow kindling rat strains during massed trial stimulation of the amygdala. Neuropharmacology. 2007;52:321–332. doi: 10.1016/j.neuropharm.2006.08.004. [DOI] [PubMed] [Google Scholar]
  284. Shorvon SD. Handbook of Epilepsy Treatment. Oxford: Wiley-Blackwell; 2010. [Google Scholar]
  285. Siegal J, Murphy GJ. Serotonergic inhibition of amygdala-kindled seizures in cats. Brain Res. 1979;147:337–340. doi: 10.1016/0006-8993(79)90858-8. [DOI] [PubMed] [Google Scholar]
  286. Sloviter RS. On the relationship between neuropathology and pathophysiology in the epileptic hippocampus of humans and experimental animals. Hippocampus. 1994a;4:250–253. doi: 10.1002/hipo.450040304. [DOI] [PubMed] [Google Scholar]
  287. Sloviter RS. The functional organization of the hippocampal dentate gyrus and its relevance to the pathogenesis of temporal lobe epilepsy. Ann Neurol. 1994b;35:640–654. doi: 10.1002/ana.410350604. [DOI] [PubMed] [Google Scholar]
  288. Small SA, Schobel SA, Buxton RB, Witter MP, Barnes CA. A pathophysiological framework of hippocampal dysfunction in ageing and disease. Nat Rev Neurosci. 2011;12:585–601. doi: 10.1038/nrn3085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  289. Song C, Zhang XY, Manku M. Increased phospholipase A2 activity and inflammatory response but decreased nerve growth factor expression in the olfactory bulbectomized rat model of depression: effects of chronic ethyl-eicosapentaenoate treatment. J Neurosci. 2009;29:14–22. doi: 10.1523/JNEUROSCI.3569-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  290. Sorrells SF, Sapolsky RM. An inflammatory review of glucocorticoid actions in the CNS. Brain Behav Immun. 2007;21:259–272. doi: 10.1016/j.bbi.2006.11.00. [DOI] [PMC free article] [PubMed] [Google Scholar]
  291. Specchio LM, Iudice A, Specchio N, La Neve A, Spinelli A, Galli R, et al. Citalopram as treatment of depression in patients with epilepsy. Clin Neuropharmacol. 2004;27:133–136. doi: 10.1097/00002826-200405000-00009. [DOI] [PubMed] [Google Scholar]
  292. Stahl S. Stahl's Essential Psychopharmacology: Neuroscientific Basis and Practical Applications. Cambridge: Cambridge University Press; 2008. [Google Scholar]
  293. Statnick M, Dailey J, Jobe P, Browning R. Neither intranigral fluoxetine nor 5,7-dihydroxytryptamine alter audiogenic seizures in genetically epilepsy-prone rats. Eur J Pharmacol. 1996;299:93–102. doi: 10.1016/0014-2999(95)00839-x. [DOI] [PubMed] [Google Scholar]
  294. Sutcigil L, Oktenli C, Musabak U, Bozkurt A, Cansever A, Uzun O, et al. Pro- and anti-inflammatory cytokine balance in major depression: effect of sertraline therapy. Clin Dev Immunol. 2007;2007:76396. doi: 10.1155/2007/76396. [DOI] [PMC free article] [PubMed] [Google Scholar]
  295. Taher TR, Salzberg M, Morris MJ, Rees S, O'Brien TJ. Chronic low-dose corticosterone supplementation enhances acquired epileptogenesis in the rat amygdala kindling model of TLE. Neuropsychopharmacology. 2005;30:1610–1616. doi: 10.1038/sj.npp.1300709. [DOI] [PubMed] [Google Scholar]
  296. Takahashi M, Hayashi S, Kakita A, Wakabayashi K, Fukuda M, Kameyama S, et al. Patients with temporal lobe epilepsy show an increase in brain-derived neurotrophic factor protein and its correlation with neuropeptide Y. Brain Res. 1999;818:579–582. doi: 10.1016/s0006-8993(98)01355-9. [DOI] [PubMed] [Google Scholar]
  297. Takeshita H, Kawahara R, Nagabuchi T, Mizukawa R, Hazama H. Serum prolactin, cortisol and growth hormone concentrations after various epileptic seizures. Jpn J Psychiatry Neurol. 1986;40:617–623. doi: 10.1111/j.1440-1819.1986.tb03176.x. [DOI] [PubMed] [Google Scholar]
  298. Tauck DL, Nadler JV. Evidence of functional mossy fiber sprouting in hippocampal formation of kainic acid-treated rats. J Neurosci. 1985;5:1016–1022. doi: 10.1523/JNEUROSCI.05-04-01016.1985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  299. Tellez-Zenteno JF, Patten SB, Jette N, Williams J, Wiebe S. Psychiatric comorbidity in epilepsy: a population-based analysis. Epilepsia. 2007;48:2336–2344. doi: 10.1111/j.1528-1167.2007.01222.x. [DOI] [PubMed] [Google Scholar]
  300. Theodore WH, Hasler G, Giovacchini G, Kelley K, Reeves-Tyer P, Herscovitch P, et al. Reduced hippocampal 5HT1A PET receptor binding and depression in temporal lobe epilepsy. Epilepsia. 2007;48:1526–1530. doi: 10.1111/j.1528-1167.2007.01089.x. [DOI] [PubMed] [Google Scholar]
  301. Thome-Souza MS, Kuczynski E, Valente KD. Sertraline and fluoxetine: safe treatments for children and adolescents with epilepsy and depression. Epilepsy Behav. 2007;10:417–425. doi: 10.1016/j.yebeh.2007.01.004. [DOI] [PubMed] [Google Scholar]
  302. Thompson NJ, Walker ER, Obolensky N, Winning A, Barmon C, Diiorio C, et al. Distance delivery of mindfulness-based cognitive therapy for depression: project UPLIFT. Epilepsy Behav. 2010;19:247–254. doi: 10.1016/j.yebeh.2010.07.031. [DOI] [PubMed] [Google Scholar]
  303. Tiemeier H, Hofman A, van Tuijl HR, Kiliaan AJ, Meijer J, Breteler MM. Inflammatory proteins and depression in the elderly. Epidemiology. 2003;14:103–107. doi: 10.1097/00001648-200301000-00025. [DOI] [PubMed] [Google Scholar]
  304. Traboulsie A, Chemin J, Kupfer E, Nargeot J, Lory P. T-type calcium channels are inhibited by fluoxetine and its metabolite norfluoxetine. Mol Pharmacol. 2006;69:1963–1968. doi: 10.1124/mol.105.020842. [DOI] [PubMed] [Google Scholar]
  305. Trimble M. Non-monoamine oxidase inhibitor antidepressants and epilepsy: a review. Epilepsia. 1978;19:241–250. doi: 10.1111/j.1528-1157.1978.tb04486.x. [DOI] [PubMed] [Google Scholar]
  306. Turrin NP, Rivest S. Innate immune reaction in response to seizures: implications for the neuropathology associated with epilepsy. Neurobiol Dis. 2004;16:321–334. doi: 10.1016/j.nbd.2004.03.010. [DOI] [PubMed] [Google Scholar]
  307. Tutka P, Barczynski B, Wielosz M. Convulsant and anticonvulsant effects of bupropion in mice. Eur J Pharmacol. 2004;499:117–120. doi: 10.1016/j.ejphar.2004.07.105. [DOI] [PubMed] [Google Scholar]
  308. Ugale RR, Mittal N, Hirani K, Chopde CT. Essentiality of central GABAergic neuroactive steroid allopregnanolone for anticonvulsant action of fluoxetine against pentylenetetrazole-induced seizures in mice. Brain Res. 2004;1023:102–111. doi: 10.1016/j.brainres.2004.07.018. [DOI] [PubMed] [Google Scholar]
  309. Vanpee D, Laloyaux P, Gillet JB. Seizure and hyponatraemia after overdose of trazadone. Am J Emerg Med. 1999;17:430–431. doi: 10.1016/s0735-6757(99)90104-3. [DOI] [PubMed] [Google Scholar]
  310. Velissaris SL, Wilson SJ, Newton MR, Berkovic SF, Saling MM. Cognitive complaints after a first seizure in adulthood: influence of psychological adjustment. Epilepsia. 2009;50:1012–1021. doi: 10.1111/j.1528-1167.2008.01893.x. [DOI] [PubMed] [Google Scholar]
  311. Vermoesen K, Serruys AS, Loyens E, Afrikanova T, Massie A, Schallier A, et al. Assessment of the convulsant liability of antidepressants using zebrafish and mouse seizure models. Epilepsy Behav. 2011;22:450–460. doi: 10.1016/j.yebeh.2011.08.016. [DOI] [PubMed] [Google Scholar]
  312. Vermoesen K, Massie A, Smolders I, Clinckers R. The antidepressants citalopram and reboxetine reduce seizure frequency in rats with chronic epilepsy. Epilepsia. 2012;53:870–878. doi: 10.1111/j.1528-1167.2012.03436.x. [DOI] [PubMed] [Google Scholar]
  313. Vezzani A, Moneta D, Conti M, Richichi C, Ravizza T, De Luigi A, et al. Powerful anticonvulsant action of IL-1 receptor antagonist on intracerebral injection and astrocytic overexpression in mice. Proc Natl Acad Sci U S A. 2000;97:11534–11539. doi: 10.1073/pnas.190206797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  314. Vezzani A, French J, Bartfai T, Baram TZ. The role of inflammation in epilepsy. Nat Rev Neurol. 2011;7:31–40. doi: 10.1038/nrneurol.2010.178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  315. Vezzani A, Friedman A, Dingledine RJ. The role of inflammation in epileptogenesis. Neuropharmacology. 2012 doi: 10.1016/j.neuropharm.2012.04.004. doi: 10.1016/j.neuropharm.2012.04.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  316. Wada Y, Nakamura M, Hasegawa H, Shiraishi J, Yamaguchi N. Microinjection of the serotonin uptake inhibitor fluoxetine elevates hippocampal seizure threshold in rats. Neurosci Res Commun. 1993;13:143–148. [Google Scholar]
  317. Wada Y, Shiraishi J, Nakamura M, Hasegawa H. Prolonged but not acute fluoxetine administration produces its inhibitory effect on hippocampal seizures in rats. Psychopharmacology (Berl) 1995;118:305–309. doi: 10.1007/BF02245959. [DOI] [PubMed] [Google Scholar]
  318. Wada Y, Shiraishi J, Nakamura M, Koshino Y. Role of serotonin receptor subtypes in the development of amygdaloid kindling in rats. Brain Res. 1997;747:338–342. doi: 10.1016/s0006-8993(96)01322-4. [DOI] [PubMed] [Google Scholar]
  319. Wada Y, Hirao N, Shiraishi J, Nakamura M, Koshino Y. Pindolol potentiates the effect of fluoxetine on hippocampal seizures in rats. Neurosci Lett. 1999;267:61–64. doi: 10.1016/s0304-3940(99)00321-3. [DOI] [PubMed] [Google Scholar]
  320. Walker ER, Obolensky N, Dini S, Thompson NJ. Formative and process evaluations of a cognitive-behavioral therapy and mindfulness intervention for people with epilepsy and depression. Epilepsy Behav. 2010;19:239–246. doi: 10.1016/j.yebeh.2010.07.032. [DOI] [PubMed] [Google Scholar]
  321. Wang GK, Mitchell J, Wang SY. Block of persistent late Na+ currents by antidepressant sertraline and paroxetine. J Membr Biol. 2008;222:79–90. doi: 10.1007/s00232-008-9103-y. [DOI] [PubMed] [Google Scholar]
  322. Watanabe K, Ashby CR, Jr, Katsumori H, Minabe Y. The effect of the acute administration of various selective 5-HT receptor antagonists on focal hippocampal seizures in freely-moving rats. Eur J Pharmacol. 2000;398:239–246. doi: 10.1016/s0014-2999(00)00258-2. [DOI] [PubMed] [Google Scholar]
  323. Weiss G, Lucero K, Fernandez M, Karnaze D, Castillo N. The effect of adrenalectomy on the circadian variation in the rate of kindled seizure development. Brain Res. 1993;612:354–356. doi: 10.1016/0006-8993(93)91686-m. [DOI] [PubMed] [Google Scholar]
  324. Wong M. Mammalian target of rapamycin (mTOR) inhibition as a potential antiepileptogenic therapy: from tuberous sclerosis to common acquired epilepsies. Epilepsia. 2010;51:27–36. doi: 10.1111/j.1528-1167.2009.02341.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  325. Wroblewski BA, McColgan K, Smith K, Whyte J, Singer WD. The incidence of seizures during tricyclic antidepressant drug treatment in a brain-injured population. J Clin Psychopharmacol. 1990;10:124–128. doi: 10.1097/00004714-199004000-00009. [DOI] [PubMed] [Google Scholar]
  326. Xu H, Steven Richardson J, Li XM. Dose-related effects of chronic antidepressants on neuroprotective proteins BDNF, Bcl-2 and Cu/Zn-SOD in rat hippocampus. Neuropsychopharmacology. 2003;28:53–62. doi: 10.1038/sj.npp.1300009. [DOI] [PubMed] [Google Scholar]
  327. Xu JH, Long L, Tang YC, Zhang JT, Hut HT, Tang FR. CCR3, CCR2A and macrophage inflammatory protein (MIP)-1a, monocyte chemotactic protein-1 (MCP-1) in the mouse hippocampus during and after pilocarpine-induced status epilepticus (PISE) Neuropathol Appl Neurobiol. 2009;35:496–514. doi: 10.1111/j.1365-2990.2009.01022.x. [DOI] [PubMed] [Google Scholar]
  328. Yan QS, Jobe PC, Dailey JW. Noradrenergic mechanisms for the anticonvulsant effects of desipramine and yohimbine in genetically epilepsy-prone rats: studies with microdialysis. Brain Res. 1993;610:24–31. doi: 10.1016/0006-8993(93)91212-b. [DOI] [PubMed] [Google Scholar]
  329. Yan QS, Jobe PC, Dailey JW. Evidence that a serotonergic mechanism is involved in the anticonvulsant effect of fluoxetine in genetically epilepsy-prone rats. Eur J Pharmacol. 1994;252:105–112. doi: 10.1016/0014-2999(94)90581-9. [DOI] [PubMed] [Google Scholar]
  330. Yeung SY, Millar JA, Mathie A. Inhibition of neuronal KV potassium currents by the antidepressant drug, fluoxetine. Br J Pharmacol. 1999;128:1609–1615. doi: 10.1038/sj.bjp.0702955. [DOI] [PMC free article] [PubMed] [Google Scholar]
  331. Yirmiya R, Pollak Y, Barak O, Avitsur R, Ovadia H, Bette M, et al. Effects of antidepressant drugs on the behavioral and physiological responses to lipopolysaccharide (LPS) in rodents. Neuropsychopharmacology. 2001;24:531–544. doi: 10.1016/S0893-133X(00)00226-8. [DOI] [PubMed] [Google Scholar]
  332. Zarate CA, Jr, Singh JB, Carlson PJ, Brutsche NE, Ameli R, Luckenbaugh DA, et al. A randomized trial of an N-methyl-D-aspartate antagonist in treatment-resistant major depression. Arch Gen Psychiatry. 2006;63:856–864. doi: 10.1001/archpsyc.63.8.856. [DOI] [PubMed] [Google Scholar]
  333. Zattoni M, Mura ML, Deprez F, Schwendener RA, Engelhardt B, Frei K, et al. Brain infiltration of leukocytes contributes to the pathophysiology of temporal lobe epilepsy. J Neurosci. 2011;31:4037–4050. doi: 10.1523/JNEUROSCI.6210-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  334. Zhao C, Deng W, Gage FH. Mechanisms and functional implications of adult neurogenesis. Cell. 2008;132:645–660. doi: 10.1016/j.cell.2008.01.033. [DOI] [PubMed] [Google Scholar]
  335. Zhu WJ, Roper SN. Brain-derived neurotrophic factor enhances fast excitatory synaptic transmission in human epileptic dentate gyrus. Ann Neurol. 2001;50:188–194. doi: 10.1002/ana.1074. [DOI] [PubMed] [Google Scholar]
  336. Zienowicz M, Wislowska A, Lehner M, Taracha E, Skorzewska A, Maciejak P, et al. The effect of fluoxetine in a model of chemically induced seizures – behavioral and immunocytochemical study. Neurosci Lett. 2005;373:226–231. doi: 10.1016/j.neulet.2004.10.009. [DOI] [PubMed] [Google Scholar]
  337. Zobel A, Wellmer J, Schulze-Rauschenbach S, Pfeiffer U, Schnell S, Elger C, et al. Impairment of inhibitory control of the hypothalamic pituitary adrenocortical system in epilepsy. Eur Arch Psychiatry Clin Neurosci. 2004;254:303–311. doi: 10.1007/s00406-004-0499-9. [DOI] [PubMed] [Google Scholar]

Articles from British Journal of Pharmacology are provided here courtesy of The British Pharmacological Society

RESOURCES