Skip to main content

The interaction between circadian rhythm and epilepsy

Abstract

Evidence about the interaction between circadian rhythms (CR) and epilepsy has been expanded with the application of advanced detection technology. An adequate understanding of how circadian system and epilepsy interact with each other could contribute to more accurate seizure prediction as well as rapid development of potential treatment timed to specific phases of CR. In this review, we present the reciprocal relationship between CR and epileptic activities from aspects of sleep effect, genetic modulation and brain biochemistry. It has been found that sleep-wake patterns, circadian timing systems and multidien rhythms have essential roles in seizure activities and interictal epileptiform discharge (IED). For instance, specific distribution patterns of seizures and IED have been reported, i.e., lighter non-rapid eye movement (NREM) sleep stage (stage 2) induces seizures while deeper NREM sleep stage (stage 3) activates IEDs. Furthermore, the epilepsy type, seizure type and seizure onset zone can significantly affect the rhythms of seizure occurrence. Apart from the common seizure types, several specific epilepsy syndromes also have a close correlation with sleep-wakefulness patterns. Sleep influences the epilepsy rhythm, and conversely, epilepsy alters the sleep rhythm through multiple pathways. Clock genes accompanied by two feedback loops of regulation have an important role in cortical excitability and seizure occurrence, which may be involved in the mTORopathy. The suprachiasmatic nuclei (SCN) has a rhythm of melatonin and cortisol secretion under the circadian pattern, and then these hormones can feed back into a central oscillator to affect the SCN-dependent rhythms, leading to variable but prominent influence on epilepsy. Furthermore, we discuss the precise predictive algorithms and chronotherapy strategies based on different temporal patterns of seizure occurrence for patients with epilepsy, which may offer a valuable indication for non-invasive closed-loop treatment system. Optimization of the time and dose of antiseizure medications, and resynchronization of disturbed CR (by hormone therapy, light exposure, ketogenic diet, novel small molecules) would be beneficial for epileptic patients in the future. Before formal clinical practice, future large-scale studies are urgently needed to assist prediction and treatment of circadian seizure activities and address unsolved restrictions.

Background

Circadian rhythm (CR), one of the most basic rhythms in mammalian behavioural and physiological processes, has been studied for hundred years. Under the 24-h day-night cycle, endogenous biological rhythms are generated by a central pacemaker of the suprachiasmatic nuclei (SCN) in the hypothalamus and control various organ systems through numerous neuronal connections and hormonal factors [1,2,3]. Emerging evidence suggests that the disrupted CR leads to multiple diseases such as cancer, mood disorder, and neurogenerative diseases [4,5,6]. Epilepsy is one of the most common neurological disorders characterized by spontaneous and recurrent seizures [7], with 30–40% of patients remaining refractory to current antiseizure medications (ASMs) [8]. Thus, alternative or adjunctive treatments are urgently required. CR and epilepsy have a crosstalk in several physiological processes such as sleep-wakefulness cycle, expression of clock-related genes, hormone secretion and epileptic activities [9,10,11,12]. A better understanding of the interaction between CR and epilepsy could facilitate more accurate prediction and potential therapeutic strategies. In this review, we will describe the reciprocal relationship between CR and epilepsy from several aspects including the sleep effect, genetic modulation and brain biochemistry, and then discuss accurate prediction algorithms and therapeutic strategies based on the chronoepileptology of patients with epilepsy.

Sleep pattern and epilepsy

Effects of sleep on epilepsy

CR modulates the brain function in several aspects, of which the regulation of sleep-wakefulness rhythm is the most important due to the different brain excitation-inhibition balances during sleep and wakefulness phases [13, 14]. A study has shown that 42% of seizures are diurnal, 21% are nocturnal and the remaining 37% are of the mixed type, indicating a significant difference in baseline seizure amount between the sleep and wakefulness phases [15]. Nearly half of the seizures occur preferentially at sleep [16] and the majority of seizures in sleep tend to begin during non-rapid eye movement sleep (NREM) [17]. In addition, a study investigating the seizure frequency in various sleep stages in patients with epilepsy showed that seizures occurred frequently at NREM sleep stage 2, followed by stage 1 and then stage 4 and stage 3 (61% in stage 2, 20% in stage 1, 14% in stage 4 and stage 3) [18]. With the advent of long-term video electroencephalography (EEG) recording of epileptiform activity, distribution of interictal epileptiform discharges (IEDs) in longer periods of time has been revealed [19, 20]. The IEDs tend to be activated during NREM sleep stage 3 and relatively inhibited during REM sleep in patients with epilepsy [21]. A comprehensive meta-analysis showed that the IED incidence is 2.46 times higher in NREM sleep stage 3, 1.75 times higher in stage 1 and 1.69 times higher in stage 2 than in REM sleep [22]. However, IEDs during REM sleep stage are more accurate for seizure localization while IEDs that occur during NREM sleep stage are likely to be widespread [23]. There is a specific distribution pattern of seizure and IED, with lighter NREM sleep (stage 2) inducing seizures while deeper NREM sleep (stage 3) activating IEDs. Interestingly, these two stages both have spindle waves in EEG recording. The frequent activation of IED or seizure during NREM sleep may be attributed to neurotransmitter change and progressive neural synchronization [18]. The gamma-aminobutyric acid (GABA) regulation decreases with the gradual reduction of the acetylcholine (Ach) level, which deactivates the suppressive GABAergic neurons of the reticular nuclei, mediating hyperpolarization of the excitatory thalamocortical network [24], and hypersynchronization during NREM sleep facilitates thalamocortical spindle wave to transform into spike wave [25]. On the contrary, the REM sleep stage characterized by suppression of the thalamocortical synchronization inhibits the spread of epileptiform discharge [26]. Similar effects on the thalamocortical circuitry have also been demonstrated for other monoamine neurotransmitters [27]. In addition, under normal conditions, concurrent drop of adenosine level and increase of adenosine kinase at the transition from wakefulness to sleep have been described, which may lead to lowering of seizure threshold in lighter NREM sleep stage and also be involved in transition of sleep stages and sleep arousal, therefore affecting seizure distribution [28]. In addition, adenosine exhibits an anticonvulsant effect while adenosine kinase shows a reverse effect [29].

In a study, epileptiform discharges were shown in 52% of patients during EEG after sleep deprivation, independent of sleep depth [30]. Therefore, in terms of EEG monitoring, sleep deprivation can be considered as a common firing method to induce epileptiform activity. Clinically, sleep deprivation after the morning also increases the seizure probability [14]. Arousal from sleep would lead to transition into lighter sleep stage and then back to deeper sleep stage, and seizures tend to occur on this descent back to deeper sleep [9]. Recently, in a longitudinal EEG investigation, sleeping longer by 1.66 ± 0.52 h, may offer protective effects in patients with refractory focal epilepsy, reducing the seizure odds by 27% in the following 48 h [31]. Cyclical alternating patterns, characterized by more sleep fragmentation, are more likely to promote seizures in sleep [32].

Frontal lobe seizures (FLS) tend to occur during sleep compared to temporal lobe seizures (TLS) [33]. For instance, 61% of FLS occur during sleep compared to only 11% of TLS based on a study of 30 patients receiving 5 days of continuous video-EEG monitoring. In addition, 22% of FLS have secondary generalization seizures during sleep compared to 20% during wakefulness, whereas TLS secondary generalization occurs at a rate of 35% during sleep and 18% during wakefulness [34]. It has been indicated that FLS occur more often during sleep than TLS; however, compared with FLS, TLS tend to undergo secondary generalization seizures when occurring during sleep [9]. In addition, seizures are more likely to occur frequently during sleep independent of day-night cycles for FLS. Sleep, especially NREM sleep, is the most robust predictor for seizure occurrence in lesional or genetic form of epilepsy, regardless of the location of ictal-seizure onset zone [35]. In contrast, a majority of TLS tend to occur following the clock phase of CR, independent of the vigilance state [14]. As shown in several studies, TLS have a higher seizure frequency in the light than the dark phase in both animals and humans [36, 37]. The seizure-like discharges displayed two peaks during the light and dark phases, respectively, in an atypical absence epilepsy model [38]. Generalized tonic-clonic seizures reached one peak during the dark phase in Aldh5a1−/− mice model [39]. Contrary to these results, evidence suggests that slow-wave sleep and passive wakefulness seemed to promote the occurrence of spike-wave discharge (SWD) in a WAG/Rij rat model, regardless of the environment condition (light) [40]. This discrepancy may be explained by the uncoupling of SWD and motor rhythm during light conditions, because the sleep-wakefulness rhythm, considered to be governed by a weaker oscillator, loses its organization earlier than the more robust oscillator of motor rhythm [40]. Another potential explanation is the differences in animal models, seizure onset zones and epilepsy types between these studies [41]. With the clinical application of closed-loop implanted neurostimulator system, a longer-period monitoring of neuronal activity has become possible. In a study of 37 drug-intractable epileptic patients implanted with a closed-loop system, IED activities were found with a multidien period of length (commonly 20–30 days), and the multidien periodicities were stable for up to 10 years [42]. Similarly, another study also reported multidien periods of seizures and IEDs (6 days and 5–7 days, respectively) [43]. Therefore, sleep-wake cycles, circadian clock systems, and multidien rhythms have essential roles in seizure occurrence and IED activities [14].

Moreover, epilepsy type, seizure type, and seizure onset zone can affect the rhythm of seizure occurrences [44, 45]. In patients with idiopathic generalized epilepsy, tonic-clonic and tonic seizures occur predominantly during sleep, whereas other generalized seizures, such as myoclonic, clonic, atonic and absence types, occur mostly during wakefulness [46]. There are close connections between seizure onset zone and seizure rhythms in focal epilepsy. In focal epileptic patients with seizures originating from the frontal lobe, most of their seizures occur at night (4:00–7:00 a.m.) while the occipital seizures and temporal seizures reach a peak in the afternoon (16:00–19:00 and 13:00–16:00, respectively) [47]. A mainly bimodal distribution of seizures has been shown in mesial temporal lobe epilepsy, with peak frequencies in the early morning and the late afternoon [48, 49]. Currently, the parietal seizures have not shown a definite pattern possibly due to the anatomical ambiguity [41]. The parietal lobe is subdivided into the fronto-parietal and occipito-parietal regions on the basis of combination with adjacent areas [50]. Of note, the occurrence of frontal lobe seizures is 12-h out of phase with occipital lobe seizures, with the frontal lobe seizure reaching a trough, while the occipital lobe seizure occurrence being at a peak between 16:00 and 19:00 [47]. In conclusion, further studies are needed to determine whether all these results could be applied to predict the seizure occurrence for epileptic patients in clinical practice.

Sleep-related epilepsy syndromes

In addition to sleep-wake patterns affecting common epilepsy types, several specific epilepsy syndromes also show a tendency of correlation with sleep. One of the most common idiopathic generalized epilepsy syndromes, juvenile myoclonic epilepsy (JME), tends to occur on the transition from sleep to wakefulness in the early morning. In patients with idiopathic focal epilepsy syndromes, benign epilepsy with centrotemporal spikes (BECTs), autosomal dominant nocturnal frontal lobe epilepsy (ADNFLE), Landau-Kleffner syndrome (LKS) and epilepsy of childhood with occipital paroxysms (BECOP) and electrical status epileptitus during slow wave sleep (ESES) also have a well-documented relationship with sleep [51, 52].

JME, characterized by recurrent myoclonic jerk shortly after awaking in the early morning, is extremely susceptive to sleep deprivation. A potential explanation of higher seizure susceptibility at this time of the day is that, using transcranial magnetic stimulation, increased cortical excitability occurs  in JME patients during the early morning compared with focal epilepsy patients or healthy control subjects [53]. Sometimes the myoclonic jerk is slight and even ignored because of its presentation as a simple clumsiness; furthermore, generalized tonic-clonic seizures could begin independently or precede the myoclonic jerk [52, 54]. In addition, patients with JME favor the evening chronotype and often suffer cognitive impairment possibly as an early awakening during the morning [12].

Accompanied by the secondary generalization with characteristic central and temporal spikes unilaterally or bilaterally in the EEG, BECT has a typical presentation with paresthesia or clonic or tonic activity of the face, tongue and throat [55]. The centrotemporal discharges become more frequent with higher amplitude during NREM sleep compared to the wakefulness. The progress of BECT is generally benign with an excellent response to medicine and remission in adolescence [9]. However, BECT patients may degenerate into epileptic encephalopathy in the form of ESEC and often suffer memory and cognitive dysfunction [56]. Impaired cognitive function may lead to increased nocturnal discharge that is involved in the plastic function that refreshes synapses during sleep [57]. However, it has been found that lowering the nocturnal discharge rate is still beneficial for improving cognitive function, despite the possible side effects of anti-IEDs medication (e.g., diazepam) [58].

ADNFLE is a type of idiopathic focal epilepsy that is identified with a genetic basis, with clusters of complex and stereotyped hypermotor seizures in the frontal lobe during NREM [59]. It is similar in clinical and neurophysiological manifestations with sporadic nocturnal frontal lobe epilepsy, frequently accompanied by paroxysmal dystonia, sudden arousal and nocturnal wandering [9]. In addition, over half of the patients have normal in interictal EEG during sleep and only a few patients display a clear-cut epileptic activity in ictal EEG recordings. ADNFLE diagnosis is difficult in clinical practice because the nocturnal manifestation of symptoms resembles other sleep disorders [60]. About 10–15% (approximately 12%) of the ADNFLE families carry mutational genes that code for subunits of the neuronal ACh receptor (nAChR) [59]. The nAChR is a pentameric ion channel permeable to cations such as Ca2+, and can result in membrane depolarization and then induce post-synaptic excitation or stimulate neurotransmitter release upon activation by ACh [61]. The dysfunction of inhibitory and excitatory transmitters mediated by mutation of the nAChR gene may trigger frequent seizures. The sleep-wake cycle is regulated by inhibitory and excitatory transmitters in several brain regions such as the locus coeruleus, the intralaminar nucleus of thalamus, the basal forebrain and the neocortex. In particular, the cholinergic projections into the thalamus nuclei are related with the regulation of thalamocortical oscillations occurring during sleep. Impairment of the function of AChRs interferes with the sleep-wake cycle and provokes seizures, thus explaining why seizures occur prevalently during sleep [60]. Moreover, emerging evidence suggests that nAChRs are a promising therapeutic target for ADNFLE patients and some typical ASMs can regulate the neurotransmitter-gated ion channel mediated by AChRs [59]. Further studies are still needed to validate these findings.

LKS is a type of acquired epileptic aphasia with language regression and epileptiform activity in children aged between 3 and 9 years [62]. LKS often occurs with continuous local spikes during sleep and also atypical spike waves in the temporal regions bilaterally [9, 63]. Moreover, the epileptiform activity is associated with deterioration of language acquisition and persistent neuropsychological deficits [57, 64]. In patients with LKS, early administration of steroids or adrenocorticotropic hormone (ACTH) can alleviate symptoms and normalize EEG [65]. BECOP is a type of idiopathic localization-related epilepsy in children and is characterized by various visual symptoms, convulsive phenomena and autonomic instability during the sleep. EEG recording of BECOP shows unilateral or bilateral spike-wave complexes over the occipital regions in sleep, which are, however, suppressed by the wakefulness [66].

Effect of epilepsy on sleep

Sleep influences the epilepsy rhythm, and epilepsy in turn alters the sleep rhythm. Epilepsy per se, seizure and interictal epileptiform activity would disrupt sleep including sleep quantity, sleep quality and sleep architecture [67]. Emerging evidence suggests that the sleep architecture is often altered in patients with epilepsy, including reduction of total sleep time and REM sleep, and increase of arousal and stage shifts, especially for those whose seizures are poorly controlled [68]. Furthermore, the occurrence of a seizure may disrupt sleep patterns by elongating sleep and, if the seizure occurs during sleep, reducing the quality of sleep [31]. In a study in 2017, the REM percentage and sleep efficiency were decreased, the total arousal index was increased, and the REM latency was prolonged in patients with epilepsy, compared with the healthy controls [69]. In a study comparing the effect of seizures occurring at night or day on sleep architecture, the REM sleep time was reduced from 18 to 12% by the seizures in the daytime and from 16 to 7% by seizures at night [70]. In addition, in seizure-free night, sleep is disintegrated more often in patients with epilepsy than in non-epilepsy subjects. These results may be attributed to the disruption of sleep and restraint of normal sleep progression by IED [71]. Regarding the effect of seizure type, patients with generalized seizures tend to suffer sleep abnormalities compared to those with simple or complex partial seizures [9]. Patients who become seizure-free following epilepsy surgery or ASM administration have a normalized sleep architecture and an improved sleep quality [72, 73]. There are two potential mechanisms mediating the disrupted sleep architecture in epilepsy. First, reticular nucleus of the thalamus (RNT) is involved in triggering of wakefulness [74]. A study has found that sleep disruption is associated with decreased excitability of RNT interneurons possibly due to the declined rebound firing after hyperpolarization, leading to sleep impairment [75]. The second mechanism of disturbed sleep is the dysfunction of hypothalamus in epilepsy. Spontaneous epileptic rats show neuronal loss in the dorsomedial hypothalamus, an important region for normal sleep regulation. Thus, hypothalamic pathology may be a link between epileptiform activity and altered sleep architecture [76].

Epileptic patients have a higher prevalence of comorbid sleep disorders than healthy controls, and patients with current seizures suffer more sleep disorders than these seizure-free epileptic patients. These sleep disorders are more likely to lead to sleep deprivation, sleep fragmentation and hypoxia, therefore, reducing the seizure threshold and promoting epileptogenesis [77]. On the contrary, multiple factors can result in breathing abnormalities during seizures. For instance, seizures spread from local brain lobes to the brainstem respiratory center and eventually cause central sleep apnea [9]. These possibilities may also lead to sudden unexpected death in the epilepsy population. ASMs are still the most potent first-line treatment for epileptic patients and ASMs exhibit both beneficial and harmful effects on sleep independent of their antiseizure effect. Phenytoin may disrupt sleep, whereas carbamazepine and newer-generation ASMs (lamotrigine, pregabalin and gabapentin) appear to consolidate sleep in a variety of common conditions [78]. In brief, sleep disruption may be related to epilepsy per se, seizure activity, comorbid sleep disorders and the effect of antiseizure medication in epileptic patients, which in turn result in poor seizure control. Apart from effects on sleep-wake patterns, epilepsy also has adverse effects on other circadian activities, including temperature rhythm and rest-activity period [49, 79]. A better understanding of the mechanisms underlying sleep and CR disruptions can help epileptic patients with sleep disorder and CR disturbance.

Clock genes and epilepsy

To our knowledge, the imbalance of inhibitory and excitatory neurotransmitters is a currently recognized mechanism of epileptogenesis and the expression of numerous ion channels and neurotransmitter receptors is modulated by the circadian rhythms [14]. With exposure to a 24-h light/dark cycle, every single cell exhibits a genetically programmed rhythm, coordinated by the SCN involved in positive and negative feedback loops of clock-related gene transcription and translation [80, 81]. By serving as transcriptional activators, clock-related genes have rapid responses to the environment condition, which influences the phase of clock to modulate the daily rhythms of behavior and physiology [82]. There are two vital genes for the generation of the circadian clock in the SCN cells, including circadian locomotor output cycles (Clock) gene (a clock circadian regulator) and arylhydrocarbon receptor nuclear translocator-like (Bmal1) gene. Transcripts of these genes are translated and the proteins form the CLOCK/BMAL1 heterodimeric complex in the cytoplasm [80]. The CLOCK/BMAL1 complex translocates to the nucleus, where it binds to specific E-box motifs and then drives the transcription of target genes including Cryptochrome (Cry1, Cry2) and Period (Per1, Per2, Per3) and other downstream genes (TEF, DBP, HLE) [14]. mRNAs of Pers and Crys are then translated into PERs and CRYs proteins in the cytoplasm. Upon increase of levels, PERs and CRYs dimerize and translocate into the nucleus where they reduce the transcriptional activity of the CLOCK/BMAL1 complex, further restraining their own transcription (Fig. 1). This negative feedback loop is also indirectly supported by the phenomenon that the Crys and Pers reach their peaks of expression during the day, whereas the transcription of Bmal1 reaches its peak at midnight (anti-phase to Crys and Pers expression) [83]. This feedback loop along with others is crucial to the generation of endogenous rhythms for adaption to the diurnal-nocturnal cycle.

Fig. 1
figure 1

A general overview of molecular mechanisms. There are two vital genes for the generation of the circadian clock in the SCN cells, Clock gene and Bmal1 gene. The protein products of these genes form a heterodimeric CLOCK/BMAL1 complex in the cytoplasm. The CLOCK/BMAL1 complex translocates to the nucleus where it binds to specific E-box motifs and then drives the transcription of target genes including Cry1-2, Per1-3, and other downstream genes (PAR bZIP transcription factor genes TEF, DBP, HLE). The mRNAs of Pers and Crys are transcribed and then translated into PERs and CRYs in the cytoplasm. The increased PERs and CRYs dimerize and translocate into the nucleus where they reduce the transcriptional activity of the CLOCK/BMAL1 complex, resulting in restraint of their own transcription. Additional feedback loops begin when the transcription of two retinoic acid-related organ receptor response elements REV-ERBA and RORA is activated by the CLOCK/BMAL1 complex. The protein products of REV-ERBA (repressor of Baml1 transcription) and RORA (activator of Baml1 transcription) enter the nucleus where they competitively bind to the promoter of Baml1 gene to stimulate the expression of Baml1. The GATOR1 complex can suppress the activity of the mTOR pathway. The S6K1 kinase of the mTOR pathway can activate Baml1 via phosphorylation to regulate the function of BMAL1 and control circadian protein synthesis. mTORC1 regulates the phosphorylation-dependent S6K and eukaryotic translation initiation factor 4E-BPs. The activated S6K stimulates S6 phosphorylation while 4E-BP1 phosphorylation causes its dissociation from eIF4E, which increases protein synthesis and then leads to a rapid clock system regulation. Clock = circadian locomotor output cycles, Bmal1= brain and muscle ARNT-like 1, Cry = Cryptochrome, Per1 = Period, TEF = translation elongation factor, DBP = D-box binding protein, RORA = RAR-related orphan receptor α, REV-ERBA = reverse erythroblastosis virus α, mTOR = mammalian target of rapamycin, GATOR1 = GAP activity toward Rags 1, S6K1 = ribosomal S6 protein kinase 1, 4E-BP = 4E-binding protein, eIF4E = eukaryotic translation initiation factor 4e

The clock-related genes not only affect the CR, but also modulate the epileptic susceptibility [84]. In a mouse model with deletion of three PAR bZIP (proline and acidic amino acid-rich basic leucine zipper) transcription factors, the audiogenic seizure threshold is decreased and spontaneous seizures occur with a circadian trend similar to the circadian distribution of sleep [85]. These data provide evidence that the downstream clock-related genes are associated with seizure susceptibility. Further study by Gerstner et al. supported for the first time the direct relationship between Baml1 gene and circadian epileptic activities in an animal model of electrical stimulation-induced seizure [11]. In their study, deletion of Baml1 gene not only abolished the circadian profile of seizure susceptibility in the mice, but also significantly reduced the seizure threshold compared to the wild-type mice. Regarding the Clock gene, its protein product, which is a core component of the CLOCK/BMAL1 complex, plays an important role in epileptonenesis. A study based on targeted deletions of the Clock gene in excitatory neurons in mice showed a lowered seizure threshold and sleep-related seizure occurrence, independent of SCN regulation of the sleep/wake cycle as Clock was preserved in SCN [86]. Interestingly, mice with deletion of Clock gene have decreased dendritic spine formation and altered electrophysiological hallmarks of neuronal microcircuits. These results supported that the loss of function of clock-related genes may disrupt the balance between neuronal excitability and inhibition, leading to the occurrence of seizures [87, 88]. On the other hand, some studies have also found disturbed expression of clock-related genes in epilepsy [79, 89]. The expression of Clock, Baml1 and Cry2 is decreased in both the early post-status epileptic (SE) phase and the epileptic phase, while the expression of Cry1 and Per1 is increased in the early post-SE phase and then decreased to the baseline level in the epileptic phase compared to the control condition [89]. In the epileptic Kcna1-null mice, in addition to disrupted diurnal and circadian rest-activity patterns, oscillations of several clock-related genes including Clock, Bmal1, Per1, and Per2 are also attenuated in the anterior hypothalamus [79]. Therefore, the disrupted oscillatory expression of clock genes is associated with altered CR in epilepsy [83].

In addition, second feedback loops are also involved in the regulation of CR. The transcription of two retinoic acid-related organ receptor response elements including reverse erythroblastosis virus ɑ (REV-ERBA) and ROR-related orphan receptor ɑ (RORA) is activated by the CLOCK/BMAL1 complex, and the REV-ERBA and RORA proteins enter the nucleus where they competitively bind to the rev response element, the promoter of Baml1 gene, to modulate the expression of Baml1 [87]. Therefore, REV-ERBA (repressor of Baml1 transcription) and RORA (activator of Baml1 transcription) are connected with negative and positive circadian oscillation of Baml1 expression, respectively. Other auxiliary loops are also mediated by the REVA-ERB/RORA-mediated transcription of E4 promoter-binding protein 4 (E4bp4) gene and the CLOCK/BMAL1-mediated transcription of the Dbp gene. Their protein products (DBP and E4BP4) dimerize and bind to the D-box elements on the promoters of Rev-erbs/Rors, serving as an additional layer of regulation [90] (Fig. 1). A study investigating the expression of RORA in epileptic rats induced by pilocarpine showed that the RORA expression was declined in the hippocampus at both acute and chronic phases [91], indicating that the RORA expression may contribute to the epileptogenic process possibly through its regulation of CR [92].

Under CR control, the mammalian target of rapamycin (mTOR) signaling pathway is a critical system that modulates cell functions (lipid and protein synthesis, proliferation and growth of cells) [93]. The GAP activity toward Rags 1 (GATOR1) complex can suppress the activity of the mTOR pathway. The ribosomal S6 protein kinase 1 in the mTOR pathway can phosphorylate thereby activating BMAL1 and control the synthesis of circadian proteins [94]. Moreover, mTORC1 regulates the phosphorylation of S6K and the eukaryotic translation initiation factor 4E-binding proteins (4E-BPs). The activated S6K stimulates S6 phosphorylation while 4E-BP1 phosphorylation causes its dissociation from eIF4E (eukaryotic translation initiation factor 4e), which increases the CAP-dependent translation and then leads to a rapid clock system regulation [95] (Fig. 1). Thus, the mTOR signaling pathway is involved in the modulation of functions of central and peripheral clock (amplitude and period) and the abnormal mTOR signaling may disturb the CLOCK/BMAL1 complex and its downstream products, resulting in epileptogenesis [14, 96]. Mutation of the GATOR1 complex and activation of the mTOR signaling pathway may both underlie the focal epilepsy [97, 98]. In addition, mTOR signaling inhibitors such as rapamycin and everolimus can reduce seizures in patients with tuberous sclerosis epilepsy [99, 100], which further support the role of the mTOR pathway in the epileptic seizures. Cdc42 that encodes cell division cycle 42 (GTP-binding protein) is associated with the mTOR protein [101]. It has been shown that Cdc42 influences the seizure threshold through the mTOR signaling pathway, which acts as a downstream target of Cdc42 [102]. Considering all these findings, the clock-related genes may contribute to structural epilepsy mediated by the mTORopathies, which may shed light into therapeutic approaches for epileptic patients, especially for structural epilepsy.

Circadian hormone pattern and epilepsy

After receiving direct environmental signal (light) from the retinal photoreceptor via the retinohypothalamic tract (or via the geniculohypothalamic indirectly), SCN transmits the signal to the pineal gland, which then secretes melatonin in a circadian pattern. The release of melatonin is at a high level during the dark period (peak at 2:00 am) and at a low level during the light period [103]. Melatonin can also act as an internal nighttime signal for the circadian system as the length of night is positively correlated with the duration of melatonin secretion [104]. Melatonin can modulate CR through activating the melatonin receptors (MT1 and MT2) in the central biological clock. Thus, melatonin can feed back into a central oscillator to affect rhythms in a SCN-dependent manner while its secretion is also modulated by the SCN clock. Disrupted circadian fluctuation of corticosterone levels has been displayed in mice with MT1 receptor knockout [105].

Notably, melatonin has a variable but prominent action on epilepsy. In patients with intractable epilepsy, the level of baseline melatonin is low during the interictal period while a high level of salivary melatonin would occur during the postictal period [106]. Furthermore, melatonin receptor whose expression depends on light and melatonin which regulates the circadian rhythms, has demonstrated an obvious change of secretion amplitude in epileptic rats with a decrease at zeitgeber time (ZT) point 0 for MT1 receptor and an increase at ZT18 for MT2 receptor, which may compensate for the severe epileptic condition [106]. Interestingly, pinealectomy significantly increases the seizure activity and facilitates epileptogenesis in a penicillin-induced model [107]. In addition, exogenous melatonin has been shown to exert beneficial effects against seizure in numerous experimental and clinical studies [108, 109]. However, the anticonvulsion effect of melatonin remains controversial. Melatonin exerts a proconvulsive effect at high doses while possessing an anticonvulsion action at lower doses [110]. This dose-dependent switch of effect may be because that melatonin at low doses can increase cortical and hypothalamic GABA levels, while high doses of melatonin reduce the GABA level in these regions [111]. This result also indicates a possible time-dependent response of melatonin. Interestingly, it has been found that melatonin has diurnal variation of susceptibility to the pentylenetetrazole-induced seizures accompanied by the lowest kindling threshold at night [112]. Agomelatine and ramelteon, novel analogues of melatonin, possess anticonvulsant effects in many experimental epilepsy models [113, 114]. However, on the other hand, agomelatine is ineffective in preventing epileptogenesis in a kainate acid-induced epilepsy model although it can suppress inflammatory factors and has neuroprotective effects against neuronal loss and gliosis [115]. Thus, although melatonin and its analogues show a potential anticonvulsion property, extensive research is still needed to clarify their effects in patients with epilepsy.

SCN also has a rhythm of cortisol secretion mediated by the hypothalamic-pituitary-adrenal (HPA) axis in a circadian pattern. The cortisol pulses reach a peak in the morning, and then gradually decline during the day, reaching a low trough at the end of active periods [116]. HPA dysfunction and rapid changes in cortisol level are involved in signal transduction and epileptiform seizure of circadian rhythm [117]. In addition, it has been found that the levels of ACTH and cortisol are lower in the CSF of infants with infantile spasms than in normal infants [118]. ACTH and corticosterone have been considered as the first-line drugs to treat infantile spasm while traditional ASMs have difficulty to control seizures in patients with infantile spasm [119]. This cortisol concentration has an important role in the homeostasis of neuronal excitability and inhibition. With the gradual reduction of ACTH and cortisol after sleep, cerebral cortex inhibition is gradually diminished, which increases the probability of seizure occurrence in epileptic patients [120]. The levels of ACTH and cortisol are lowered in chronic epilepsy, increasing the secretion of corticotropin- releasing hormone (CRH) as an enhanced negative feedback. It is worth noting that CRH, an excitatory neuropeptide, can reduce the seizure threshold, resulting in seizure occurrences [121]. Furthermore, cortisol can improve the GABA level to reduce the neuron excitation through inhibiting GABA uptake by synapses and promoting the rapid synthesis of GABA [122]. In future studies, application of exogenous cortisol or measurement of endogenous cortisol levels can help further reveal the effects of circadian cortisol variability on seizure activities, using high temporal-resolution EEG recording.

Seizure prediction

The use of advanced techniques could advance the understanding of the relationship between epileptic activities and circadian rhythm, facilitating accurate seizure prediction. With accurate seizure prediction, seizure risk can be reduced and drug-related side effects attenuated by applying ASMs at differential dosing times. Moreover, patients and caregivers can also take some measures to avoid specific situation where seizures tend to occur at a certain time, if mathematical predictability algorithms are available. Thus, mathematical predictability models should be developed to predict the circadian fluctuation of seizure susceptibility based on age, sleep-wake pattern, rest-activity state, hormones, and seizure type. A previous study showed that combination of the temporal distribution of seizures with chronic intracranial EEG signals improved the seizure forecasting by algorithms [123]. In addition, the accuracy of seizure-forecasting algorithms will be enhanced by addition of multidien IED rhythms because the seizure occurrence also displays a multidien rhythm [42]. Nevertheless, it is difficult to generally apply this invasive device in the epileptic population; this prediction is feasible only when the benefits of detecting seizures outweigh the risk of neurosurgery for implantation of the intracranial device. Furthermore, the negative and positive predictive values of the forecasting algorithms are still not fully suitable for clinically meaningful prediction. Thus, further studies are needed to investigate the novel non-invasive and accurate devices (such as the wearable detection device); meanwhile, the factors that could affect seizure activities such as age, seizure type, and seizure origin zone should be considered. The rapid growth of clinical databases, and the integration of learning algorithms into wearable detection devices may help develop closed-loop systems and improve seizure- prediction models [124, 125].

Chronotherapy

Epileptic chronotherapy, which is a treatment strategy of drug administration at times of higher seizure susceptibility, can enhance the efficacy and reduce the side-effect of ASM. Apart from optimizing the administration timing of ASMs according to the circadian rhythms, the pharmacokinetic characteristics should also be taken into account, including the absorption, metabolism, distribution and elimination of the drugs [14]. This chronotherapy has been applied in several chronic diseases with a circadian pattern, including epilepsy [126]. One study investigated whether clobazam treatment tailored to the timing of seizures could improve seizure control in 27 epileptic patients. In this study, higher-evening differential dose of clobazam achieved better seizure control in patients with predominantly nighttime and early-morning seizures, possibly because the differential dosing provided higher overall treatment doses at times of greatest seizure occurrence without increased side-effects at other times [127]. Another study found that a chronotherapeutic dosing schedule of carbamazepine and phenytoin involving administration of most or all daily doses of ASM at 8:00 p.m. can improve seizure control than standard treatment regimens [128]. In 17 patients with nocturnal or early-morning seizures, taking a higher ASM dose at bedtime than in the morning led to seizure freedom in 64.7% (11/17) of patients and 88.2% (15/17) experienced ≥50% reductions of seizures, along with only 11.8% of patients experiencing fatigue [129]. Just like the closed-loop diabetes pump treatment, closed-loop detection-treatment systems may be a better choice for patients with epilepsy, especially for refractory epilepsy. Development of portable devices can decrease EEG recordings and hospital appointment, allowing for rapid detection of seizures and immediate treatment of patients. For generalized tonic-clonic seizures, some algorithms have proven useful in clinical practice [130]. However, a large-scale clinical trial of closed-loop chronotherapy has not been successfully conducted. Further studies are still needed to investigate the value of closed-loop chronotherapy in clinical practice.

Resetting the disturbed circadian rhythms may improve seizure control as the potential mechanism underlying the circadian pattern of seizure occurrence is associated with dysregulation of clock genes. Melatonin and its analogues display a potential anticonvulsion property, probably due to their capacity to synchronize the disrupted circadian rhythms. However, there are inconsistent reports on the effect of melatonin on seizure control across different studies, even in systematic reviews. Light, as a crucial environmental cue, can reset the endogenous rhythms possibly because it can synchronize the clock system, even after a quick exposure [131]. In a randomized control trial, the light therapy significantly reduced symptoms of depression and anxiety in focal epileptic patients [132]. Another double-blinded randomized trial reported that the bright light exposure significantly reduced the seizure frequency in patients with drug-intractable epilepsy, but also induced seizures in patients with extra-TLE [133]. Clinical data on light therapy are still limited, which reinforces the necessity for further clinical studies to evaluate the specific effects of light therapy in epilepsy management. Ketogenic diet (KD), a high fat and low carbohydrate diet, also shows anticonvulsant and antiepileptogenic effects for intractable epilepsy [134]. In Kcna1-knockout mice that experience seizures with impaired circadian behavior, KD reduces the seizure frequency and restores the disturbed diurnal rhythmicity, which may be because that the diet, similar to light cue, can gradually reset the phase of rhythmic gene expression [135]. Moreover, KD exerts the anticonvulsant or antiepileptogenic effect possibly via inhibition of the mTOR pathway [136]. When circadian clock systems that regulate essential cellular and physiological processes are disturbed, the risk of epileptogenesis and seizure frequency would rise. Developing a ‘clock drug’ targeting the altered clock components or clock-associated cellular pathways is a novel therapeutic strategy for clock-related diseases. Fortunately, several small molecules have been identified as clock modulators that regulate the circadian pattern and the physiological outputs in disease models [137], which may be used as novel therapeutic agents or coadjuvant agents for epileptic patients in the future.

Limitation and direction

Despite great research progress on the relationship between circadian rhythm and epilepsy in animal models, there are also limitations that hinder translation into clinical practice. It is difficult to verify that the seizure occurrence is influenced by the circadian rhythm, sleep-wakefulness state, or light-dark cycles in humans, because it is challenging to apply specific protocols such as consistent light-dark cycles or consistent routine protocols in epileptic patients. Many confounding factors (e.g., ASM treatment, psychiatric comorbidities, and genetic variations) may also limit the observational results in clinical studies. Although the chronotherapy has been applied in a few animal models and in epileptic patients, the individual nature of epilepsy may affect the popularity of chronotherapy. Given the inter-individual variability of seizure occurrence, a large-scale patient-based tracking of seizure cluster would be advantageous in integrating the individual seizure patterns into accurate closed-loop systems. Despite the emerging evidence at the cellular and molecular levels, our knowledge regarding the relationship between epilepsy and circadian rhythm from the microcosmic perspective is far from enough, limited mainly at the brain network level. Thus, further large-scale studies are needed to assist prediction and treatment of circadian seizure activities and address these restrictions.

Conclusions

With the application of long-term video EEG and implantation of closed-loop neurostimulator system, our knowledge on the interaction between circadian rhythms and epilepsy has been greatly enhanced. It has been found that sleep-wake patterns, circadian timing systems and multidien rhythms have an essential role in seizure activities and IEDs. For instance, seizures and IEDs occur preferentially at the sleep state especially in the NREM stage, and sleep deprivation after the morning also increases the seizure risk. In addition, epilepsy type, seizure type and seizure onset zone significantly affected the rhythm of seizure occurrences, which should be considered into the seizure forecasting algorithms for accurate prediction. Apart from the effect of sleep on the occurrence of seizures of common epilepsy types, several specific epilepsy syndromes have a close correlation with the sleep-wakefulness patterns such as JME, BECTs, ADNFLE, LKS, BECOP, and ESES. Sleep influences the epilepsy rhythm and epilepsy in turn alters the sleep rhythm via the pathway of epilepsy pe se, seizures activity, comorbid sleep disorders and the effect of ASMs. Clock-related genes accompanied by two feedback loops of regulation play an important role in cortical excitability and seizure occurrence, which may be involved in the mTORopathies. These are considered as a useful model for advancing understanding of how the clock genes affect seizure occurrence and inspiring epileptic gene therapies in the future. SCN forms a rhythm of secretion of melatonin and cortisol under the circadian pattern, which then form a feedback as a central oscillator to affect rhythms, exerting variable but prominent actions on epilepsy. A time-dependent hormone therapy may be a potential option for epileptic patients, although their anticonvulsive action remains still controversial. Deeper understanding of the relationship between epileptic activities and circadian rhythm may facilitate accurate seizure prediction. Integration of prediction algorithms into closed-loop detection-treatment systems with non-invasive devices may be a better choice for individualized intervention. Furthermore, optional timing and dosing of ASMs, and resynchronization of disturbed CR (by hormone therapy, light exposure, KD, or novel small molecules) would be beneficial therapeutic strategies for epileptic patients in the future.

Availability of data and materials

Not applicable.

Abbreviations

Ach:

Acetylcholine

ACTH:

Adrenocorticotropic hormone

ADNFLE:

Autosomal dominant nocturnal frontal lobe epilepsy

ASMs:

Antiseizure medications

BECOP:

Benign epilepsy of childhood with occipital paroxysms

BECT:

Benign epilepsy with centrotemporal spike

Bmal1:

Brain and muscle ARNT-like 1

CR:

Circadian rhythm

CRH:

Corticotropin releasing hormone

Cry:

Cryptochrome

DBP:

D-box binding protein

EEG:

electroencephalograph

ESES:

Electrical status epileptitus during slow wave sleep

FLS:

Frontal lobe seizures

GATOR1:

GAP activity toward Rags 1

HPA:

Hypothalamic pituitary adrenal

IED:

Interictal epileptiform discharge

JME:

Myoclonic epilepsy

KD:

Ketogenic diet

LKS:

Landau-Kleffner syndrome

mTOR:

Mammalian target of rapamycin

nAChR:

Neuronal acetylcholine receptor

NREM:

Non-rapid eye movement

REV-ERBA:

Reverse erythroblastosis virus α

RNT:

Reticular nucleus of the thalamus

RORA:

RAR-related orphan receptor α

SCN:

Suprachiasmatic nuclei

SWD:

Spike-wave discharge

TEF:

Translation elongation factor

TLS:

Temporal lobe seizures

4E-BP:

4E-binding protein

References

  1. Patke A, Young MW, Axelrod S. Molecular mechanisms and physiological importance of circadian rhythms. Nat Rev Mol Cell Biol. 2020;21:67–84.

    Article  CAS  PubMed  Google Scholar 

  2. Partch CL, Green CB, Takahashi JS. Molecular architecture of the mammalian circadian clock. Trends Cell Biol. 2014;24:90–9.

    Article  CAS  PubMed  Google Scholar 

  3. Morin LP, Allen CN. The circadian visual system. Brain Res Rev. 2006;51:1–60.

    Article  CAS  PubMed  Google Scholar 

  4. Wang Z, Wang H, Wang Z, He S, Jiang Z, Yan C, et al. Associated analysis of PER1/TUBB2B with endometrial cancer development caused by circadian rhythm disorders. Med Oncol. 2020;37:90.

    Article  CAS  PubMed  Google Scholar 

  5. Duncan MJ, Smith JT, Franklin KM, Beckett TL, Murphy MP, St Clair DK, et al. Effects of aging and genotype on circadian rhythms, sleep, and clock gene expression in APPxPS1 knock-in mice, a model for Alzheimer’s disease. Exp Neurol. 2012;236:249–58.

    Article  CAS  PubMed  Google Scholar 

  6. Satyanarayanan SK, Su H, Lin YW, Su KP. Circadian rhythm and melatonin in the treatment of depression. Curr Pharm Des. 2018;24:2549–55.

    Article  CAS  PubMed  Google Scholar 

  7. Strine TW, Kobau R, Chapman DP, Thurman DJ, Price P, Balluz LS. Psychological distress, comorbidities and health behaviors among US adults with seizures: results from the 2002 National Health Interview Survey. Epilepsia. 2005;46:1133–9.

    Article  PubMed  Google Scholar 

  8. Brodie MJ, Barry SJ, Bamagous GA, Norrie JD, Kwan P. Patterns of treatment response in newly diagnosed epilepsy. Neurology. 2012;78:1548–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Kothare SV, Kaleyias J. Sleep and epilepsy in children and adolescents. Sleep Med. 2010;11:674–85.

    Article  PubMed  Google Scholar 

  10. Ardura J, Andres J, Garmendia JR, Ardura F. Melatonin in epilepsy and febrile seizures. J Child Neurol. 2010;25:888–91.

    Article  PubMed  Google Scholar 

  11. Gerstner JR, Smith GG, Lenz O, Perron IJ, Buono RJ, Ferraro TN. BMAL1 controls the diurnal rhythm and set point for electrical seizure threshold in mice. Front Syst Neurosci. 2014;8:121.

    Article  PubMed  PubMed Central  Google Scholar 

  12. Pung T, Schmitz B. Circadian rhythm and personality profile in juvenile myoclonic epilepsy. Epilepsia. 2006;47:111–4.

    Article  PubMed  Google Scholar 

  13. Quigg M. Circadian rhythms: interactions with seizures and epilepsy. Epilepsy Res. 2000;42:43–55.

    Article  CAS  PubMed  Google Scholar 

  14. Khan S, Nobili L, Khatami R, Loddenkemper T, Cajochen C, Dijk DJ, et al. Circadian rhythm and epilepsy. Lancet Neurol. 2018;17:1098–108.

    Article  PubMed  Google Scholar 

  15. Jin B, Aung T, Geng Y, Wang S. Epilepsy and its interaction with sleep and circadian rhythm. Front Neurol. 2020;8(11):327.

    Article  Google Scholar 

  16. Herman ST, Walczak TS, Bazil CW. Distribution of partial seizures during the sleep-wake cycle: differences by seizure onset site. Neurology. 2001;56:1453–9.

    Article  CAS  PubMed  Google Scholar 

  17. Mendez M, Radtke RA. Interactions between sleep and epilepsy. J Clin Neurophysiol. 2001;18:106–27.

    Article  CAS  PubMed  Google Scholar 

  18. Minecan D, Natarajan A, Marzec M, Malow B. Relationship of epileptic seizures to sleep stage and sleep depth. Sleep. 2002;25:56–61.

    Article  Google Scholar 

  19. Spencer DC, Sun FT, Brown SN, Jobst BC, Fountain NB, Wong VS, et al. Circadian and ultradian patterns of epileptiform discharges differ by seizure-onset location during long-term ambulatory intracranial monitoring. Epilepsia. 2016;57:1495–502.

    Article  CAS  PubMed  Google Scholar 

  20. Anderson C, Tcheng T, Sun F, Morrell M. Day–night patterns of epileptiform activity in 65 patients with long-term ambulatory electrocorticography. J Clin Neurophysiol. 2015;32:406–12.

    Article  PubMed  Google Scholar 

  21. Sammaritano M, Gigli GL, Gotman J. Interictal spikes during wakefulness and sleep and localization of foci in temporal lobe epilepsy. Neurology. 1991;41:290–7.

    Article  CAS  PubMed  Google Scholar 

  22. Ng M, Pavlova M. Why are seizures rare in rapid eye movement sleep? Review of the frequency of seizures in different sleep stages. Epilepsy Res Treat. 2013;2013:932790.

    PubMed  PubMed Central  Google Scholar 

  23. Okanari K, Baba S, Otsubo H, Widjaja E, Sakuma S, Go CY, et al. Rapid eye movement sleep reveals epileptogenic spikes for resective surgery in children with generalized interictal discharges. Epilepsia. 2015;56:1445–53.

    Article  PubMed  Google Scholar 

  24. Steriade M, McCormick DA, Sejnowski TJ. Thalamocortical oscillations in the sleeping and aroused brain. Science. 1993;262:679–85.

    Article  CAS  PubMed  Google Scholar 

  25. Ferrillo F, Beelke M, Nobili L. Sleep EEG synchronization mechanisms and activation of interictal epileptic spikes. Clin Neurophysiol. 2000;111:S65–73.

    Article  PubMed  Google Scholar 

  26. Shouse MN, Siegel JM, Szymusiak WMF, R, Morrison AR. Mechanisms of seizure suppression during rapid-eye movement sleep in cats. Brain Res. 1989;505:271–82.

    Article  CAS  PubMed  Google Scholar 

  27. McCormick DA. Neurotransmitter actions in the thalamus and cerebral cortexand their role in neuromodulation of thalamocortical activity. Prog Neurobiol. 1992;39:317–88.

    Article  Google Scholar 

  28. De Weerd A, de Haas S, Otte A, Trenite DK, van Erp G, Cohen A, et al. Subjective sleep disturbance in patients with partial epilepsy: a questionnaire–based study on prevalence and impact on quality of life. Epilepsia. 2004;45:1397–404.

    Article  PubMed  Google Scholar 

  29. Meletti S, Cantalupo G, Volpi L, Rubboli G, Magaudda A, Tassinari CA. Rhythmic teeth grinding induced by temporal lobe seizures. Neurology. 2004;62:2306–9.

    Article  CAS  PubMed  Google Scholar 

  30. Fountain NB, Kim JS, Lee SI. Sleep deprivation activates epileptiform discharges independent of the activating effects of sleep. J Clin Neurophysiol. 1998;15:69–75.

    Article  CAS  PubMed  Google Scholar 

  31. Dell KL, Payne DE, Kremen V, Maturana MI, Gerla V, Nejedly P, et al. Seizure likelihood varies with day-to-day variations in sleep duration in patients with refractory focal epilepsy: A longitudinal electroencephalography investigation. EClinicalMedicine. 2021;37:100934.

    Article  PubMed  PubMed Central  Google Scholar 

  32. Miano S, PiaVilla M, Blanco D, Zamora E, Rodriguez R, Ferri R, et al. Development of NREM sleep instability-continuity (cyclic alternating pattern) in healthy term infants aged 1 to 4 months. Sleep. 2009;32:83–90.

    PubMed  PubMed Central  Google Scholar 

  33. Bazil CW, Walczak TS. Effects of sleep and sleep stage on epileptic and nonepileptic seizures. Epilepsia. 1997;38:56–62.

    Article  CAS  PubMed  Google Scholar 

  34. Herman ST, Walczak TS, Bazil CW. Distribution of partial seizures during the sleep-wake cycle: differences by seizures onset site. Neurology. 2001;56:1453–8.

    Article  CAS  PubMed  Google Scholar 

  35. Pavlova MK, Shea SA, Scheer FA, Bromfield EB. Is there a circadian variation of epileptiform abnormalities in idiopathic generalized epilepsy? Epilepsy Behav. 2009;16:461–7.

    Article  PubMed  Google Scholar 

  36. Raedt R, Van Dycke A, Van Melkebeke D, De Smedt T, Claeys P, Wyckhuys T, et al. Seizures in the intrahippocampal kainic acid epilepsy model: characterization using long-term video-EEG monitoring in the rat. Acta Neurol Scand. 2009;119:293–303.

    Article  CAS  PubMed  Google Scholar 

  37. Quigg M, Straume M, Menaker M, Bertram EH III. Temporal distribution of partial seizures: comparison of an animal model with human partial epilepsy. Ann Neurol. 1998;43:748–55.

    Article  CAS  PubMed  Google Scholar 

  38. Stewart LS, Bercovici E, Shukla R, Serbanescu I, Persad V, Mistry N, et al. Daily rhythms of seizure activity and behavior in a model of atypical absence epilepsy. Epilepsy Behav. 2006;9:564–72.

    Article  PubMed  Google Scholar 

  39. Stewart LS, Nylen KJ, Persinger MA, Cortez MA, Gibson KM, Snead OC III. Circadian distribution of generalized tonic-clonic seizures associated with murine succinic semialdehyde dehydrogenase deficiency, a disorder of GABA metabolism. Epilepsy Behav. 2008;13:290–4.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Smyk MK, Coenen AM, Lewandowski MH, van Luijtelaar G. Endogenous rhythm of absence epilepsy: relationship with general motor activity and sleep-wake states. Epilepsy Res. 2011;93:120–7.

    Article  PubMed  Google Scholar 

  41. Gurkas E, Serdaroglu A, Hirfanoglu T, Kartal A, Yilmaz U, Bilir E. Sleepwake distribution and circadian patterns of epileptic seizures in children. Eur J Paediatr Neurol. 2016;20:549–54.

    Article  PubMed  Google Scholar 

  42. Baud MO, Kleen JK, Mirro EA, Andrechak JC, King-Stephens D, Chang EF, et al. Multi-day rhythms modulate seizure risk in epilepsy. Nat Commun. 2018;9:88.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  43. Baud MO, Ghestem A, Benoliel JJ, Becker C, Bernard C. Endogenous multidien rhythm of epilepsy in rats. Exp Neurol. 2019;315:82–7.

    Article  PubMed  Google Scholar 

  44. van Campen JS, Valentijn FA, Jansen FE, Joëls M, Braun KP. Seizure occurrence and the circadian rhythm of cortisol: a systematic review. Epilepsy Behav. 2015;47:132–7.

    Article  PubMed  Google Scholar 

  45. Ramgopal S, Powell C, Zarowski M, Alexopoulos AV, Kothare SV, Loddenkemper T. Predicting diurnal and sleep/wake seizure patterns in paediatric patients of different ages. Epileptic Disord. 2014;16:56–66.

    Article  PubMed  Google Scholar 

  46. Zarowski M, Loddenkemper T, Vendrame M, Alexopoulos AV, Wyllie E, Kothare SV. Circadian distribution and sleep/wake patterns of generalized seizures in children. Epilepsia. 2011;52:1076–83.

    Article  PubMed  Google Scholar 

  47. Durazzo TS, Spencer SS, Duckrow RB, Novotny EJ, Spencer DD, Zaveri HP. Temporal distributions of seizure occurrence from various epileptogenic regions. Neurology. 2008;70:1265–71.

    Article  CAS  PubMed  Google Scholar 

  48. Nzwalo H, Menezes Cordeiro I, Santos AC, Peralta R, Paiva T, Bentes C. 24-hour rhythmicity of seizures in refractory focal epilepsy. Epilepsy Behav. 2015;55:75–8.

    Article  Google Scholar 

  49. Pitsch J, Becker AJ, Schoch S, Müller JA, Curtis M, Gnatkovsky V. Circadian clustering of spontaneous epileptic seizures emerges after pilocarpine-induced status epilepticus. Epilepsia. 2017;58:1159–71.

    Article  CAS  PubMed  Google Scholar 

  50. Hofstra WA, Spetgens WP, Leijten FS, van Rijen PC, Gosselaar P, van der Palen J, et al. Diurnal rhythms in seizures detected by intracranial electrocorticographic monitoring: an observational study. Epilepsy Behavior. 2009;14:617–21.

    Article  PubMed  Google Scholar 

  51. Halasz P. Sleep and epilepsy. Handb Clin Neurol. 2012;107:305–22.

    Article  PubMed  Google Scholar 

  52. Bazil CW. Nocturnal seizures. Semin Neurol. 2004;24:293–300.

    Article  PubMed  Google Scholar 

  53. Badawy RA, MacDonell RA, Jackson GD, Berkovic SF. Why do seizures in generalized epilepsy often occur in the morning? Neurology. 2009;73:218–22.

    Article  CAS  PubMed  Google Scholar 

  54. Xu L, Guo D, Liu YY, Qiao DD, Ye JY, Xue R. Juvenile myoclonic epilepsy and sleep. Epilepsy Behav. 2018;80:326–30.

    Article  PubMed  Google Scholar 

  55. Guerrini R, Pellacani S. Benign childhood focal epilepsies. Epilepsia. 2012;53:9–18.

    Article  PubMed  Google Scholar 

  56. Fejerman N. Atypical rolandic epilepsy. Epilepsia. 2009;50:9–12.

    Article  PubMed  Google Scholar 

  57. Halasz P, Bodizs R, Ujma PP, Fabo D, Szucs A. Strong relationship between NREM sleep, epilepsy and plastic functions-a conceptual review on the neurophysiology background. Epilepsy Res. 2019;150:95–105.

    Article  PubMed  Google Scholar 

  58. Baglietto MG, Battaglia FM, Nobili L, Tortorelli S, De Negri E, Calevo MG, et al. Neuropsychological disorders related to Interictal epileptic discharges during sleep in benign epilepsy of childhood with centrotemporal or rolandic spikes. Dev Med Child Neurol. 2001;43:407–12.

    Article  CAS  PubMed  Google Scholar 

  59. Becchetti A, Aracri P, Meneghini S, Brusco S, Amadeo A. The role of nicotinic acetylcholine receptors in autosomal dominant nocturnal frontal lobe epilepsy. Front Physiol. 2015;6:22.

    Article  PubMed  PubMed Central  Google Scholar 

  60. Combi R, Dalprà L, Tenchini ML, Ferini-Strambi L. Autosomal dominant nocturnal frontal lobe epilepsy-a critical overview. J Neurol. 2004;251:923–34.

    CAS  PubMed  Google Scholar 

  61. Fucile S. Ca2+ permeability of nicotinic acetylcholine receptors. Cell Calcium. 2004;35:1–8.

    Article  CAS  PubMed  Google Scholar 

  62. Conroy J, McGettigan PA, McCreary D, Shah N, Collins K, Parry-Fielder B, et al. Towards the identification of a genetic basis for Landau-Kleffner syndrome. Epilepsia. 2014;55:858–65.

    Article  CAS  PubMed  Google Scholar 

  63. Datta AN, Oser N, Ramelli GP, Zanda Gobbin N, Lantz G, Penner IK, et al. BECTS evolving to Landau-Kleffner syndrome and back by subsequent recovery: A longitudinal language reorganization case study using fMRI, source EEG, and neuropsychological testing. Epilepsy Behavior. 2013;27:107–14.

    Article  CAS  PubMed  Google Scholar 

  64. Riccio CA, Vidrine SM, Cohen MJ, Acosta-Cotte D, Park Y. Neurocognitive and behavioral profiles of children with Landau-Kleffner syndrome. Appl Neuropsychol Child. 2017;6:345–54.

    Article  PubMed  Google Scholar 

  65. Ramanathan RS, Ahluwalia T, Sharma A. Landau-Kleffner syndrome-a rare experience. East J Med. 2012;17:36–9.

    Google Scholar 

  66. Tsai ML, Lo HY, Chaou WT. Clinical and electroencephalographic findings in early and late onset benign childhood epilepsy with occipital paroxysms. Brain and Development. 2001;23:401–5.

    Article  CAS  PubMed  Google Scholar 

  67. Vaughn BV, D’ Cruz OF.  Sleep and epilepsy. Semin Neurol. 2004;24:301–13.

    Article  PubMed  Google Scholar 

  68. Matos G, Andersen ML, do Valle AC, Tufifik S. The relationship between sleep and epilepsy: evidence from clinical trials and animal models. J Neurol Sci. 2010;295:1–7.

    Article  PubMed  Google Scholar 

  69. Mekky JF, Elbhrawy SM, Boraey MF, Omar HM. Sleep architecture in patients with juvenile myoclonic epilepsy. Sleep Med. 2017;38:116–21.

    Article  PubMed  Google Scholar 

  70. Bazil CW, Castro IH, Walczak TS. Reduction of rapid eye movement sleep by diurnal and nocturnal seizures in temporal lobe epilepsy. Arch Neurol. 2000;57:363–8.

    Article  CAS  PubMed  Google Scholar 

  71. Becker DA, Fennell EB, Carney PR. Sleep disturbance in children with epilepsy. Epilepsy Behav. 2003;4:651–8.

    Article  PubMed  Google Scholar 

  72. De Paolis F, Colizzi E, Milioli G, Grassi A, Riccardi S, Puligheddu M, et al. Effects of antiepileptic treatment on sleep and seizures in nocturnal frontal lobe epilepsy. Sleep Med. 2013;14:597–604.

    Article  PubMed  Google Scholar 

  73. Serafini A, Kuate C, Gelisse P, Velizarova R, Gigli GL, Coubes P, et al. Sleep before and after temporal lobe epilepsy surgery. Seizure. 2012;21:260–5.

    Article  PubMed  Google Scholar 

  74. McCormick DA. Cholinergic and noradrenergic modulation of thalamocortical processing. Trends Neurosci. 1989;12:215–21.

    Article  CAS  PubMed  Google Scholar 

  75. Kalume F, Oakley JC, Westenbroek RE, Gile J, de la Iglesia HO, Scheuer T, et al. Sleep impairment and reduced interneuron excitability in a mouse model of Dravet syndrome. Neurobiol Dis. 2015;77:141–54.

    Article  PubMed  PubMed Central  Google Scholar 

  76. Bastlund JF, Jennum P, Mohapel P, Penschuck S, Watson WP. Spontaneous epileptic rats show changes in sleep architecture and hypothalamic pathology. Epilepsia. 2005;46:934–8.

    Article  PubMed  Google Scholar 

  77. Zgodzinski W, Rubaj A, Kleinrok Z, Sieklucka-Dziuba M. Effect of adenosine A1 and A2 receptor stimulation on hypoxia-induced convulsions in adult mice. Pol J Pharmacol. 2001;53:83–92.

    CAS  PubMed  Google Scholar 

  78. Grigg-Damberger MM, Ralls F. Sleep disorders in adults with epilepsy: past, present, and future directions. Curr Opin Pulm Med. 2014;20:542–9.

    Article  PubMed  Google Scholar 

  79. Wallace E, Wright S, Schoenike B, Roopra A, Rho JM, Maganti RK. Altered circadian rhythms and oscillation of clock genes and sirtuin 1 in a model of sudden unexpected death in epilepsy. Epilepsia. 2018;59:1527–39.

    Article  CAS  PubMed  Google Scholar 

  80. Kennaway DJ. Clock genes at the heart of depression. J Psychopharmacol. 2010;24:5–14.

    Article  CAS  PubMed  Google Scholar 

  81. Honma S. The mammalian circadian system: a hierarchical multi-oscillator structure for generating circadian rhythm. J Physiol Sci. 2018;68:207–19.

    Article  PubMed  Google Scholar 

  82. Bunney WE, Bunney BG. Molecular clock genes in man and lower animals: possible implications for circadian abnormalities in depression. Neuropsychopharmacol. 2000;22:335–45.

    Article  CAS  Google Scholar 

  83. Xu C, Yu J, Ruan Y, Wang Y, Chen Z. Decoding circadian rhythm and epileptic activities: clues from animal studies. Front Neurol. 2020;11:751.

    Article  PubMed  PubMed Central  Google Scholar 

  84. Shearman LP, Sriram S, Weaver DR, Maywood ES, Chaves I, Zheng B, et al. Interacting molecular loops in the mammalian circadian clock. Science. 2000;288:1013–9.

    Article  CAS  PubMed  Google Scholar 

  85. Gachon F, Fonjallaz P, Damiola F, Gos P, Kodama T, Zakany J, et al. The loss of circadian PAR bZip transcription factors results in epilepsy. Genes Dev. 2014;18:1397–412.

    Article  Google Scholar 

  86. Li P, Fu X, Smith NA, Ziobro J, Curiel J, Tenga MJ, et al. Loss of CLOCK results in dysfunction of brain circuits underlying focal epilepsy. Neuron. 2017;96:387.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Leite Goes Gitai D, de Andrade TG, Dos Santos YDR, Attaluri S, Shetty AK. Chronobiology of limbic seizures: potential mechanisms and prospects of chronotherapy for mesial temporal lobe epilepsy. Neurosci Biobehav Rev. 2019;98:122–34.

    Article  PubMed  Google Scholar 

  88. Wu H, Liu Y, Liu L, Meng Q, Du C, Li K, et al. Decreased expression of the clock gene Bmal1 is involved in the pathogenesis of temporal lobe epilepsy. Mol Brain. 2021;14:113.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Matos HC, Koike BDV, Pereira WDS, de Andrade TG, Castro OW, Duzzioni M, et al. Rhythms of core clock genes and spontaneous locomotor activity in post-status epilepticus model of mesial temporal lobe epilepsy. Front Neurol. 2018;9:632.

    Article  PubMed  PubMed Central  Google Scholar 

  90. Takahashi JS. Transcriptional architecture of the mammalian circadian clock. Nat Rev Genet. 2017;18:164–79.

    Article  CAS  PubMed  Google Scholar 

  91. Rocha AKA, de A, de Lima E, do Amaral FG, Peres R, Cipolla-Neto J, et al. Pilocarpine-induced epilepsy alters the expression and daily variation of the nuclear receptor RORα in the hippocampus of rats. Epilepsy Behav. 2016;55:38–46.

  92. Solt LA, Burris TP. Action of RORs and their ligands in (patho)physiology. Trends Endocrinol Metab. 2012;23:619–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Laplante M, Sabatini DM. mTOR signaling in growth control and disease. Cell. 2012;149:274–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Lipton J, Yuan E, Boyle L, Ebrahimi-Fakhari D, Kwiatkowski E, Nathan A, et al. The circadian protein BMAL1 regulates translation in response to S6K1-mediated phosphorylation. Cell. 2015;161:1138–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Cao R, Li A, Cho H, Lee B, Obrietan K. Mammalian target of rapamycin signaling modulates photic entrainment of the suprachiasmatic circadian clock. J Neurosci. 2010;30:6302–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Ramanathan C, Kathale ND, Liu D, Lee C, Freeman DA, Hogenesch JB, et al. mTOR signaling regulates central and peripheral circadian clock function. PLoS Genet. 2018;14:e1007369.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  97. Scheffer IE, Heron SE, Regan BM, Mandelstam S, Crompton DE, Hodgson BL, et al. Mutations in mammalian target of rapamycin regulator DEPDC5 cause focal epilepsy with brain malformations. Ann Neurol. 2014;75:782–7.

    Article  CAS  PubMed  Google Scholar 

  98. Ricos MG, Hodgson BL, Pippucci T, Saidin A, Ong YS, Heron SE, et al. Mutations in the mammalian target of rapamycin pathway regulators NPRL2 and NPRL3 cause focal epilepsy. Ann Neurol. 2016;79:120–31.

    Article  CAS  PubMed  Google Scholar 

  99. French JA, Lawson JA, Yapici Z, Ikeda H, Polster T, Nabbout R, et al. Adjunctive everolimus therapy for treatment-resistant focal-onset seizures associated with tuberous sclerosis (EXIST-3): a phase 3, randomised, double-blind, placebo-controlled study. Lancet. 2016;388:2153–63.

    Article  CAS  PubMed  Google Scholar 

  100. Curatolo P. Mechanistic target of rapamycin (mTOR) in tuberous sclerosis complex-associated epilepsy. Pediatr Neurol. 2015;52:281–9.

    Article  PubMed  Google Scholar 

  101. Govek EE, Hatten ME, Van Aelst L. The role of Rho GTPase proteins in CNS neuronal migration. Dev Neurobiol. 2011;71:528–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Fang Y, Park IH, Wu AL, Du G, Huang P, Frohman MA, et al. PLD1 regulates mTOR signaling and mediates Cdc42 activation of S6K1. Curr Biol. 2003;13:2037–44.

    Article  CAS  PubMed  Google Scholar 

  103. Arendt J, Rajaratnam SM. Melatonin and its agonists: an update. Br J Psychiatry. 2008;193:267–9.

    Article  PubMed  Google Scholar 

  104. Hardeland R, Madrid JA, Tan DX, Reiter RJ. Melatonin, the circadian multioscillator system and health: the need for detailed analyses of peripheral melatonin signaling. J Pineal Res. 2012;52:139–66.

    Article  CAS  PubMed  Google Scholar 

  105. Comai S, Gobbi G. Unveiling the role of melatonin MT2 receptors in sleep, anxiety and other neuropsychiatric diseases: a novel target in psychopharmacology. J Psychiatry Neurosci. 2014;39:6–21.

    Article  PubMed  PubMed Central  Google Scholar 

  106. Salva MA, Hartley S. Mood disorder, circadian rhythms, melatonin and melatonin agonists. J Central Nerv Syst Disord. 2012;4:15–26.

    Google Scholar 

  107. Yildirim M, Aydin-Abidin S, Abidin I, Akca M, Canpolat S, Cansu A. Evaluation of the role of chronic daily melatonin administration and pinealectomy on penicillin-induced focal epileptiform activity and spectral analysis of ECoG in rats: an in vivo electrophysiological study. Neurochem Res. 2013;38:1672–85.

    Article  CAS  PubMed  Google Scholar 

  108. Goldberg-Stern H, Oren H, Peled N, Garty BZ. Effect of melatonin on seizure frequency in intractable epilepsy: a pilot study. J Child Neurol. 2012;27:1524–15.

    Article  PubMed  Google Scholar 

  109. Uberos J, Augustin-Morales MC, Molina Carballo A, Florido J, Narbona E, Muñoz-Hoyos A. Normalization of the sleep-wake pattern and melatonin and 6-sulphatoxy-melatonin levels after a therapeutic trial with melatonin in children with severe epilepsy. J Pineal Res. 2011;50:192–6.

    CAS  PubMed  Google Scholar 

  110. Banach M, Gurdziel E, Jedrych M, Borowicz KK. Melatonin in experimental seizures and epilepsy. Pharmacol Rep. 2011;63:1–11.

    Article  CAS  PubMed  Google Scholar 

  111. Vimala PV, Bhutada PS, Patel FR. Therapeutic potential of agomelatine in epilepsy and epileptic complications. Med Hypotheses. 2014;82:105–10.

    Article  CAS  PubMed  Google Scholar 

  112. Golombek DA, Duque DF, De Brito SMG, Burin L, Cardinali DP. Time-dependent anticonvulsant activity of melatonin in hamsters. Eur J Pharmacol. 1992;210:253–8.

    Article  CAS  PubMed  Google Scholar 

  113. Demirel EA, Erdogan MA, Cinar BP, Erbas O. The reducing effect of agomelatine on pentylenetetrazol-induced convulsion. Biol Fut. 2019;70:336–40.

    CAS  Google Scholar 

  114. Khan S, Khurana M, Vyas P, Vohora D. The role of melatonin and its analogues in epilepsy. Rev Neurosci. 2021;32:49–67.

    Article  CAS  Google Scholar 

  115. Tchekalarova J, Atanasova D, Nenchovska Z, Atanasova M, Kortenska L, Gesheva R, et al. Agomelatine protects against neuronal damage without preventing epileptogenesis in the kainate model of temporal lobe epilepsy. Neurobiol Dis. 2017;104:1–14.

    Article  CAS  PubMed  Google Scholar 

  116. Dickmeis T. Glucocorticoids and the circadian clock. J Endocrinol. 2009;200:3–22.

    Article  CAS  PubMed  Google Scholar 

  117. Young EA, Abelson J, Lightman SL. Cortisol pulsatility and its role in stress regulation and health. Front Neuroendocrinol. 2004;25:69–76.

    Article  CAS  PubMed  Google Scholar 

  118. Yang G, Zou LP, Wang J, Ding YX. Epigenetic regulation of glucocorticoid receptor and infantile spasms. Med Hypotheses. 2011;76:187–9.

    Article  CAS  PubMed  Google Scholar 

  119. Hussain SA, Treatment of infantile spasms. Epilepsia Open. 2018;3:143–54.

    Article  PubMed  PubMed Central  Google Scholar 

  120. Zhang SW, Liu YX. Changes of serum adrenocorticotropic hormone and cortisol levels during sleep seizures. Neurosci Bull. 2008;24:84–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Mitsugi N, Arita J, Kimura F. Effects of intracerebroventricular administration of growth for hormone releasing factor on somatostatin secretion into rat hypophysial portal blood. Neuroendocrinol. 1990;51:93–6.

    Article  CAS  Google Scholar 

  122. Delorey TM, Olsen RW. Gamma-aminobutyric acid A receptor structure and function. Biol Chem. 1992;267:16747–50.

    Article  CAS  Google Scholar 

  123. Karoly PJ, Ung H, Grayden DB, Kuhlmann L, Leyde K, Cook MJ, et al. The circadian profile of epilepsy improves seizure forecasting. Brain. 2017;140:2169–82.

    Article  PubMed  Google Scholar 

  124. Devinsky O, Dilley C, Ozery-Flato M, Aharonov R, Goldschmidt Y, Rosen-Zvi M, et al. Changing the approach to treatment choice in epilepsy using big data. Epilepsy Behav. 2016;56:32–7.

    Article  PubMed  Google Scholar 

  125. Ulate-Campos A, Coughlin F, Gaínza-Lein M, Fernández IS, Pearl PL, Loddenkemper T. Automated seizure detection systems and their effectiveness for each type of seizure. Seizure. 2016;40:88–101.

    Article  CAS  PubMed  Google Scholar 

  126. Manganaro S, Loddenkemper T, Rotenberg A. The need for antiepileptic drug chronotherapy to treat selected childhood epilepsy syndromes and avert the harmful consequences of drug resistance. J Cent Nerv Syst Dis. 2017;9:1179573516685883.

    Article  PubMed  PubMed Central  Google Scholar 

  127. Thome-Souza S, Klehm J, Jackson M, Kadish NE, Manganaro S, Fernández IS, et al. Clobazam higher-evening differential dosing as an add-on therapy in refractory epilepsy. Seizure. 2016;40:1–6.

    Article  PubMed  Google Scholar 

  128. Yegnanarayan R, Mahesh SD, Sangle S. Chronotherapeutic dose schedule of phenytoin and carbamazepine in epileptic patients. Chronobiol Int. 2006;23:1035–46.

    Article  CAS  PubMed  Google Scholar 

  129. Guilhoto LM, Loddenkemper T, Vendrame M, Bergin A, Bourgeois BF, Kothare SV. Higher evening antiepileptic drug dose for nocturnal and early-morning seizures. Epilepsy Behav. 2011;20:334–7.

    Article  CAS  PubMed  Google Scholar 

  130. van Andel J, Thijs RD, de Weerd A, Arends J, Leijten F. Non-EEG based ambulatory seizure detection designed for home use: what is available and how will it influence epilepsy care? Epilepsy Behav. 2016;57:82–9.

    Article  PubMed  Google Scholar 

  131. Kaladchibachi S, Fernandez F. Precision light for the treatment of psychiatric disorders. Neural Plast. 2018;2018:5868570.

    Article  PubMed  PubMed Central  Google Scholar 

  132. Baxendale S, O’Sullivan J, Heaney D. Bright light therapy for symptoms of anxiety and depression in focal epilepsy: randomized controlled trial. Br J Psychiatry. 2013;202:352–6.

    Article  PubMed  Google Scholar 

  133. Baxendale S, O’Sullivan J, Heaney D. Bright light therapy as an add on treatment for medically intractable epilepsy. Epilepsy Behav. 2012;24:359–64.

    Article  PubMed  Google Scholar 

  134. Marsh EB, Freeman JM, Kossoff EH, Vining EP, Rubenstein JE, Pyzik PL, et al. The outcome of children with intractable seizures: a 3- to 6-year follow-up of 67 children who remained on the ketogenic diet less than one year. Epilepsia. 2006;47:425–30.

    Article  PubMed  Google Scholar 

  135. Fenoglio-Simeone KA, Wilke JC, Milligan HL, Allen CN, Rho JM, Maganti RK. Ketogenic diet treatment abolishes seizure periodicity and improves diurnal rhythmicity in epileptic Kcna1-null mice. Epilepsia. 2009;50:2027–34.

    Article  PubMed  PubMed Central  Google Scholar 

  136. McDaniel SS, Rensing NR, Thio LL, Yamada KA, Wong M. The ketogenic diet inhibits the mammalian target of rapamycin (mTOR) pathway. Epilepsia. 2011;52:e7–e11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Ribeiro RFN, Cavadas C, Silva MMC. Small-molecule modulators of the circadian clock: pharmacological potentials in circadian-related diseases. Drug Discov Today. 2021;26:1620–41.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work was supported by the Clinical Research Plan of SHDC (SHDC2020CR3066B).

Author information

Authors and Affiliations

Authors

Contributions

JD and XW conceptualized the review. JD revised and made supplements to the manuscript. MSL searched and read the related articles. MSL wrote and revised the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Jing Ding.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no financial or other conflicts of interest.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liu, M., Ding, J. & Wang, X. The interaction between circadian rhythm and epilepsy. Acta Epileptologica 4, 28 (2022). https://doi.org/10.1186/s42494-022-00094-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s42494-022-00094-0

Keywords