Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2025 Feb 1.
Published in final edited form as: Neurosci Biobehav Rev. 2023 Dec 22;157:105523. doi: 10.1016/j.neubiorev.2023.105523

Circadian neurogenetics and its implications in neurophysiology, behavior, and chronomedicine

Muhammad Naveed a,b, Owen Y Chao a, Jennifer W Hill b, Yi-Mei Yang a,c, Joseph P Huston d, Ruifeng Cao e,f,*
PMCID: PMC10872425  NIHMSID: NIHMS1958085  PMID: 38142983

Abstract

The circadian rhythm affects multiple physiological processes, and disruption of the circadian system can be involved in a range of disease-related pathways. The genetic underpinnings of the circadian rhythm have been well-studied in model organisms. Significant progress has been made in understanding how clock genes affect the physiological functions of the nervous system. In addition, circadian timing is becoming a key factor in improving drug efficacy and reducing drug toxicity. The circadian biology of the target cell determines how the organ responds to the drug at a specific time of day, thus regulating pharmacodynamics. The current review brings together recent advances that have begun to unravel the molecular mechanisms of how the circadian clock affects neurophysiological and behavioral processes associated with human brain diseases. We start with a brief description of how the ubiquitous circadian rhythms are regulated at the genetic, cellular, and neural circuit levels, based on knowledge derived from extensive research on model organisms. We then summarize the latest findings from genetic studies of human brain disorders, focusing on the role of human clock gene variants in these diseases. Lastly, we discuss the impact of common dietary factors and medications on human circadian rhythms and advocate for a broader application of the concept of chronomedicine.

Keywords: circadian rhythm, clock gene, behavior, neurophysiology, psychiatric disorders, chronomedicine

1. Introduction

The circadian clocks are endogenous timekeeping mechanisms that allow organisms to synchronize their physiology with the ever-changing environment and perform biological processes at the most appropriate time of the day (Koronowski and Sassone-Corsi, 2021). Circadian rhythms control various biological processes in living organisms, spanning from bacteria to humans (Bell-Pedersen et al., 2005). Perhaps the most apparent manifestation of circadian rhythms is the daily sleep and wake cycle of animals; however, many other fundamental neurophysiological processes are also under regulation by the circadian clock, from body temperature control and hormone secretion to cognition and mood (Alvord et al., 2021). In mammals, the suprachiasmatic nucleus (SCN) in the hypothalamus acts as the “master clock” for the brain and body, regulating their physiological rhythms (Hastings et al., 2018). The rhythmic outputs from the SCN are responsible for orchestrating rhythms in various body systems. The complex interaction between the core clock gene and its protein byproducts is responsible for circadian rhythms at the cellular level.

Disruption of circadian rhythms by genetic or environmental insults can lead to disorders of various physiological processes (Hastings et al., 2003; Maywood et al., 2006). Genetic variations in circadian clock genes are increasingly associated with a variety of common human diseases. Therefore, the link between circadian physiology and other physiological processes has implications for human physiology and medicine. Notably, humans in modern society are increasingly disregarding natural environmental cues that regulate our body’s internal clock, resulting in aberrant body rhythms and an increased risk of developing certain diseases. It is therefore important to understand how these perturbations in circadian rhythms may affect our physiology and susceptibility to disease, to identify pathogenic mechanisms underlying circadian disruption so that novel therapeutics can be developed based on that knowledge.

The current review assembles recent findings that have begun to untangle the molecular mechanisms by which circadian biology influences neurophysiological and behavioral processes connected with human brain illnesses. We begin with a quick overview of how the ubiquitous circadian rhythms are regulated at the genetic, cellular, and brain circuit levels, based on substantial research on model species. We then discuss the most recent discoveries from genetic investigations of human brain illnesses, with a focus on the role of human clock gene variations in these diseases. Finally, we address the effects of common dietary variables and drugs on human circadian rhythms and argue for a broader use of the chronomedicine concept.

2. Circadian timekeeping mechanisms in mammalian cells

Circadian rhythms can be found within cells. Rhythmic cellular processes are driven by rhythmic gene expression, which involves several interlocking transcription-translation feedback loops (TTFLs) consisting of approximately a dozen so-called “clock genes” in mammals. Recent work has found that the intracellular signaling network is coupled to the core clock gene feedback loops, establishing an interphase between the clock machinery and a myriad of intracellular processes. Thus, these cellular processes are regulated by the circadian clock and, in turn, provide feedback to the circadian timing process, synchronizing the cellular clock with cell physiology and metabolism.

2.1. The TTFLs that drive rhythmic gene expression in mammalian cells

The circadian clock system encompasses a cell-autonomous transcript feedback loop comprising a basic set of evolutionarily conserved genes in animals (Takahashi, 2017). During the daytime, the basic helix-loop-helix PAS domain, involving the transcription element NPAS2 (or CLOCK), interacts with aryl hydrocarbon receptor nuclear translocator-like (ARNTL1/BMAL1) to form heterodimers (Vanselow and Kramer, 2010). The CLOCK/BMAL1 heterodimer binds to the E-box enhancers and activates the transcription of the Per1/2/3 and Cry1/2 genes. The Per and Cry1 mRNAs are translated into proteins, which form multiple protein complexes with CSNKIe/d in the cytosol. Once these protein complexes accumulate to a certain level, they translocate into the cell nucleus, interact with Clock/Bmal1 complexes, and inhibit gene transcription, thereby forming a negative feedback loop (Lowrey and Takahashi, 2000). At night, the PER and CRY proteins are phosphorylated and degraded through ubiquitination-mediated mechanisms, allowing BMAL1/CLOCK to initiate a new repeating cycle (Lee et al., 2001). In addition to this feedback loop, the CLOCK/BMAL1 heterodimer also activates transcription of nuclear receptor subfamily 1 group D members 1/2 (NR1D1/2 or REV-ERBα/β) and the retinoic acid receptor-associated orphan receptors (RORs) α/β/γ. REV-ERBs inhibit, and RORs activate Bmal1 transcription, respectively, creating additional feedback loops (Preitner et al., 2002). REV-ERBs and the retinoic acid RORs α/β/γ inhibit or activate Bmal1 transcription through the ROR response elements (RREs), respectively, creating additional feedback loops (Preitner et al., 2002). Apart from these core feedback loops, many other proteins modulate the clock proteins at post-transcriptional levels, including protein kinases, phosphatases, ubiquitin ligases, etc. Together, these mechanisms collaboratively fine-tune the speed and robustness of the circadian clock (Figure 1).

Figure 1.

Figure 1.

The circadian clock gene network in mammals is composed of overlapping transcriptional-translational feedback loops. A collection of fundamental clock genes form transcription-translation feedback loops in the circadian clock system. The circadian clock in mammals is made up of a major negative feedback loop that involves the genes Clock (and its paralog Npas2), Bmal1, Per1/2, and Cry1/2. CLOCK (or NPAS2) and BMAL1 are transcription factors with basic helix-loop-helix PAS-domains that stimulate transcription of the Per and Cry genes. The resultant PER and CRY proteins heterodimerize, translocate to the nucleus, and suppress transcription by interacting with the CLOCK-BMAL1 complex. After some time, the PER-CRY repressor complex degrades, allowing CLOCK-BMAL1 to initiate a new transcriptional cycle. Rev-Erb is a direct target of the CLOCK-BMAL1 transcription activator complex in the secondary autoregulatory feedback loop. REV-ERB binds RREs in the Bmal1 promoter and competes with an ROR to suppress Bmal1 transcription.

2.2. Intracellular signaling pathways that regulate clock gene expression

Cells respond to external stimuli and maintain intracellular homeostasis via various signal transduction pathways that couple extracellular and intracellular signals to gene expression. These signaling pathways regulate many aspects of cell biology and are essential for proper cell function. The cellular clock is constantly regulated by signals originating from outside the cell, which are transmitted through intracellular signaling pathways to regulate gene expression in the cell. Moreover, these signaling pathways can also interact with one another, forming a complex intracellular signaling network that forms an interface between the circadian timing process and cell physiology and metabolism. The known cell signaling pathways that regulate clock gene expression are summarized in Figure 2. Most of these discoveries were made in SCN neurons.

Figure 2.

Figure 2.

Important signaling pathways and protein kinases have been demonstrated to function within the SCN. Several neurotransmitters and neuropeptides that affect SCN neurons can activate plasma membrane ion channels and receptors, resulting in intracellular signaling events. GPCR activation can signal through Gs, Gi, and Gq proteins to activate adenylyl cyclase (AC), inhibit AC, or activate PLC, respectively. AC increases cAMP synthesis, which activates PKA and EPAC (exchange proteins activated by cAMP). PKA and cAMP can both stimulate CREB-mediated transcription. The hydrolysis of PI (4,5) P2 into diacylglycerol (DAG) and IP3 is catalyzed by PLC.

DAG activates PKC at the plasma membrane, whereas IP3 diffuses into the cytosol and causes the ER to release intracellular Ca2+ reserves. Ionotropic receptors, G-protein-gated ion channels, voltage-gated ion channels, and RTKs can all cause an increase in cytosolic Ca2+ levels. The phosphorylation of the receptor by GRK2 results in the downregulation of GPCR signaling. The RAF, MEK1/2, and ERK1/2 pathways are activated at the plasma membrane via receptor-mediated Ras activation. PKC promotes Ras/MAPK signaling by activating Ras or deactivating RKIP-mediated Raf inhibition via phosphorylation of RKIP. MAPK/ERK stimulates RSK1 and MSK1, which induce CREB-mediated transcription. MAPK/ERK also acts as an upstream activator of mTOR, which increases translation by activating p70S6K and blocking 4E-BP1-mediated eIF4E repression. Ca2+-induced CaMKII activation can connect to cGMP synthesis via the NOS/GC pathway in terms of Ca2+ signaling. PKG is activated by cGMP, promoting CREB-mediated transcription. The NOS pathway also activates Dexras1, a small G protein that inhibits GPCR-mediated Gi activation and indirectly inhibits AC via ligand-independent Gi activation. Many upstream signaling events converge on the MAPK/ERK pathway, which plays a critical role in photic entrainment through its effects on CREB. SIK1 is a CRE-inducible gene that works as a feedback inhibitor of CREB signaling by suppressing CREB-dependent transcription coactivator 1 (CRTC1). Finally, various protein kinases, including CK1 and CK2, GSK3, SIK3, and GRK2, have been found to phosphorylate clock proteins PER and CRY.

Photic information from the retina is transmitted to the SCN and triggers the release of neurotransmitters such as glutamate and pituitary AC cyclase-activating peptide (PACAP), which in turn activates a series of intracellular signaling cascades. Among these signaling pathways, the MAPK/ERK pathway is a key signaling pathway that couples photic stimulation to gene expression in the SCN (Obrietan et al., 1998). This MAPK cascade involves sequential phosphorylation-induced activation of Raf, MEK1/2, and ERK1/2. During the early or late subjective night, a brief burst of light triggers a significant increase in ERK1/2 phosphorylation in the SCN (Coogan and Piggins, 2003; Obrietan et al., 1998), which is essential for coupling light to SCN clock entrainment (Butcher et al., 2002). Phosphorylated ERK1/2, in turn, activates MSKs and ribosomal S6 kinases (RSKs) to regulate gene expression and clock entrainment (Butcher et al., 2005; Cao et al., 2013a). MSK1 regulates the photic entrainment of the circadian clock by facilitating light-induced histone phosphorylation, CREB phosphorylation, and Per1 expression (Cao et al., 2013a). Dexras1, an upstream modulator of ERK1/2 and a Ras-like protein, enhances photic responses and inhibits the non-photic response of the circadian clock (Cheng et al., 2004). The Raf-1 kinase inhibitor protein (RKIP) also plays a role in modifying MAPK/ERK signaling in the suprachiasmatic circadian clock (Antoun et al., 2012). Furthermore, light at night triggers phosphorylation of cAMP response element-binding protein (CREB) at Ser133 and Ser142, activating CRE-dependent gene expression in the SCN (Gau et al., 2002; Ginty et al., 1993). The JNK cascade represents another major MAPK signaling pathway that is activated by light at night (but not during the day), indicating that JNK1 may also be associated with photic entrainment (Pizzio et al., 2003). Indeed, the knockdown of JNK3 has been shown to alter behavioral rhythms, resulting in longer free-running periods and impaired phase shifts in response to light (Yoshitane et al., 2012).

Various studies have established a connection between CaMKII and the photic entrainment process. Light pulses delivered during the subjective night rapidly induced CaMKII phosphorylation in the SCN under entrainment and free-running conditions (Agostino et al., 2004). The activation of the CaMKII upstream modulator by light plays an important role in the phase-shifting and induction of Per1/2 genes in the hamster suprachiasmatic circadian clock (Yokota et al., 2001). Administration of a centralized CaMKII inhibitor significantly attenuated phase delays and advances caused by bright light pulses in hamsters (Golombek and Ralph, 1994). Consistent with these effects, CaMKII antagonists attenuated glutamate-induced delays in SCN firing rates, and systemic administration of calmodulin antagonists inhibited light-induced expression of c-Fos in the rat SCN (Fukushima et al., 1997). GSK3 is also associated with photic entrainment in the SCN. Late-night light pulses activate GSK3β activity across the SCN, and inhibiting GSK3 eliminates the continuous increase in the electrical activity of the SCN in response to late-night light pulses (Paul et al., 2017).

In addition to transcriptional regulation, light can regulate mRNA translation in the SCN through translational control pathways. The mechanistic/mammalian target of the rapamycin (mTORC1) pathway is activated by light at night as a downstream event of ERK MAPK activation (Cao et al., 2008; Cao et al., 2010). Importantly, the clock-resetting effects of light are sensitive to rapamycin regulation, probably through the regulation of PER1/2 protein levels in the SCN. The mTORC1 pathway also regulates the intracellular coupling of SCN neurons through translational control of vasoactive intestinal peptide (VIP) via its target 4E-BP1(Cao et al., 2013b). VIP is an important neuropeptide that couples SCN cells. In mice lacking 4E-BP1, SCN exhibits larger Per2::Luc rhythms, and mice demonstrate enhanced entrainment to a shifted light/dark cycle. Conversely, in Mtor+/− mice, VIP expression is decreased, and SCN cell synchrony is impaired (Liu et al., 2018). The mTOR signaling cascade also regulates cellular clock properties, such as period length and amplitude, in various cells and tissues, indicating a significant role of translational control in regulating autonomous clock properties. Another important translational control pathway tat is regulated by light is the MNK/eIF4E phosphorylation pathway, which is downstream of the ERK MAPK pathway. Light at night triggers MNK activation and eIF4E phosphorylation at Ser 209 in the SCN. In mice, where eIF4E Ser209 is mutated to alanine and thus cannot be phosphorylated, light pulse-induced clock resetting and the capability of clock entrainment to T-cycles are impaired. eIF4E phosphorylation facilitates mRNA translation of Per1 and Per2, which may underlie the photic entertainment of the SCN clock (Cao et al., 2015).

3. The master circadian pacemaker in mammals: the suprachiasmatic nucleus (SCN)

In mammals, the circadian system consists of a central pacemaker and peripheral oscillators found in almost all cell types in the body (Albrecht, 2012). The SCN’s primary input pathway is the retinohypothalamic tract (RHT), which originates from intrinsically photosensitive retinal ganglion cells. The SCN rhythmically regulates other brain regions as well as peripheral organs, through neuronal and humoral outputs. The pituitary gland is the only known portal system in the mammalian brain. In the mouse brain, a second portal pathway connects the capillary beds of the central brain clock in the SCN and the organum vasculosum of the lamina terminalis (OVLT) (Yao et al., 2021). Numerous physiological and behavioral processes exhibit daily rhythms, such as body temperature, blood pressure, hormone secretion, learning and memory, mood, etc.

3.1. The SCN neurons are heterogeneous and express different neuronal peptides

The SCN can be roughly divided into the ventrolateral “core” and the dorsomedial “shell” regions (Antle and Silver, 2005) (Figure 3). All SCN neurons are GABAergic; however, they express different neuropeptides. The ventrolateral SCN expresses VIP and gastrin-releasing peptide (GRP) and is directly responsive to photic input (Card et al., 1981; Mikkelsen et al., 1991). In contrast, the dorsolateral SCN expresses arginine vasopressin (AVP), responds secondarily to photic stimulation, and plays a crucial role in maintaining circadian rhythmicity (Antle and Silver, 2005). Extensive intracellular communication occurs within and between different SCN regions, allowing heterogeneous clock cells to synchronize with each other. VIP, expressed in the ventral SCN, plays an important role in the photic resetting of SCN oscillations (Aïoun et al., 1998). Application of VIP to ex-vivo SCN slices during the late subjective night advanced the phase of neuronal firing, stimulated the expression of Per1/2, and mimicked the effects of light at late night (Meyer-Spasche and Piggins, 2004; Nielsen et al., 2002). VIP also mediates circadian rhythmicity and synchrony in SCN neurons (Aton et al., 2005). The synchronization of SCN cells is disrupted in mice lacking Vip or Vpac2 (the gene encoding the receptor for VIP in the SCN), and animals show arrhythmic behavior in constant darkness. GRP transmits photic phase-shifting signals in the SCN by increasing the neurophysiological activity of SCN neurons (Kallingal and Mintz, 2014). Additionally, during the early and late nights, microinjection of GRP into the SCN induces Per1 expression throughout the dorsal and ventral SCN and behavioral phase shifts (Gamble et al., 2007). In contrast to VIP, AVP is thought to primarily function as an output signal of the SCN clock. Deletion of AVP receptors (V1a and V1b receptors) does not affect mouse free-running rhythms in constant darkness. However, these mice exhibit accelerated re-entrainment to shifted light/dark cycles, indicating the AVP signaling confers the SCN an intrinsic resistance to external perturbations (Yamaguchi, 2018). Prokineticin 2 (PK2) and its receptor (PKR2) are highly expressed in the SCN and are crucial for regulating circadian behavior (Li et al., 2018). However, overexpression of PK2 leads to a slight but significant reduction in transcript levels of Bmal1, Per1, Per2, and Cry2 in the SCN. Thus, PK2 signaling is necessary for the amplitude of rhythmic behavior, but it has minimal impact on the entrainment of the SCN and circadian pacemaking (Li et al., 2006; Li et al., 2018; Prosser et al., 2007).

Figure 3.

Figure 3.

Neurochemical composition of the mammalian SCN. The SCN is divided into dorsal (shell) and ventral (core) regions. Most, if not all, of the ~20,000 SCN cells are GABAergic but differ in their neuropeptide content. AVP and VIP delineate the shell and core regions, respectively. GRP-expressing neurons are primarily localized in the medial core. PK2 is another SCN neuropeptide that is highly expressed by shell and core SCN neurons.

The regulation of SCN by glutamatergic signaling has been extensively studied. Both ionic and metabotropic glutamate receptors have been shown to modulate central pacemaker activity (Gannon and Rea, 1994). Glutamate application to SCN tissue explants has been shown to induce neuronal firing (Bos and Mirmiran, 1993), and optic nerve stimulation results in glutamate release (Liou et al., 1986). Similarly, the application of selective receptor antagonists can reversibly reduce SCN neuronal activity induced by RHT stimulation (Cahill and Menaker, 1989). A non-selective antagonist of excitatory amino acid receptors, cis-2,3-piperidine-dicarboxylic acid, can reversibly block the postsynaptic response evoked in the SCN by stimulation of the optic nerve. In addition, treatment with NMDA receptor antagonists perturbed light-induced nocturnal phase shifts in wheel-running activity as well as SCN neuronal activity (Colwell and Menaker, 1992). The administration of NMDA or AMPA receptor agonists during the subjective night recapped the phase-shifting effects of light in vitro (Shibata et al., 1994). Early-night glutamate treatment leading to phase delays involves the activation of cAMP/PKA signaling and increasing Per1 expression (Tischkau et al., 2000). In contrast, late-night glutamate stimulation-induced phase advances are mediated through cGMP/PKG signaling and are attenuated by cAMP/PKA activation (Tischkau et al., 2003). Brain-derived neurotrophic factor (BDNF) has also been shown to increase glutamate-induced phase shifts of neuronal activity in the SCN (Kim et al., 2006).

Besides glutamate, SCN neurons are regulated by PACAP and GABA. The application of PACAP to SCN tissue explants amplified the intensity of glutamate-induced phase delays of neuronal firing rhythms during the early night but inhibited glutamate-induced phase advances during the late night (Chen et al., 1999). While GABA is primarily recognized as an inhibitory neurotransmitter in the adult brain, it can also exert excitatory effects. GABA modulates photic and non-photic phase-shifting of the central clock by GABAA and GABAB receptors (Gillespie et al., 1997; Mintz et al., 2002). Almost all the SCN neurons are GABAergic (Moore and Speh, 1993). Studies have found that SCN neurons display excitatory GABAergic signaling to modulate light input and block GABAergic signaling to attenuate phase delays during the subjective early night (McNeill et al., 2018). In addition, GABAergic signaling is important for the synchrony of SCN neurons (Liu and Reppert, 2000). GABA can both phase shift and synchronize clock cells through GABAA receptors. The GABAergic network comprises both excitatory and inhibitory connections (Freeman et al., 2013). Endogenous GABAA signaling decreases the precision of circadian oscillations in SCN neuronal networks.

Besides receiving glutamatergic input from the RHT, the SCN, specifically its VIP-containing regions, is also innervated by serotonergic projections from the median raphe nuclei, via which serotonin facilitates the glutamate input during the daytime and inhibits it during the nighttime (Reghunandanan and Reghunandanan, 2006). Serotonin can modulate both photic and non-photic effects, including phase-shifting of SCN neuronal firing, locomotor activity rhythms, melatonin secretion, body temperature rhythms, and gene expression (Horikawa and Shibata, 2004; Horikawa et al., 2000; Paulus and Mintz, 2013; Smith et al., 2008). Acetylcholine (ACh) was one of the first neuromodulators presumed to play a key role in circadian rhythm regulation. Cholinergic signaling can modulate both photic and non-photic phase-shifting of the central clock (Basu et al., 2016; Yang et al., 2010). Glycine, the primary inhibitory neurotransmitter in the brain, can also exhibit excitatory transmission through allosteric activation of NMDA receptors (Johnson and Ascher, 1987). Glycine is capable of regulating the activity of clock neurons and resetting their rhythmic activity according to the stage of the daily cycle (Mordel et al., 2011). Glycine can also affect other aspects of SCN function by affecting NMDA receptors, particularly in the regulation of sleep and body temperature (Kawai et al., 2015).

3.2. An emerging role of glial cells in circadian timekeeping

Glial cells constitute a significant population of cells in the human nervous system. The main classes of glial cells include astrocytes, oligodendrocytes, and microglia. Among these, astrocytes are the most abundant and well-characterized type of glial cell in the mammalian brain, closely involved in the metabolic, circulatory, and neuromodulatory control of neurons (Ben Haim and Rowitch, 2017). Glial cells, in addition to neurons, also display circadian rhythms (Prolo et al., 2005). Among glial cells, astrocytes have been extensively studied in the context of circadian rhythms. Several studies have shown that the clocks within mammalian astrocytes can influence neuronal oscillators and subsequent behavioral rhythms, suggesting a harmonious interaction between SCN neurons and glial cells in shaping circadian rhythms. During the dark phase as opposed to the light phase, there is a reduced number of astrocytic processes connecting with synapses in the hippocampus, resulting in a slower removal of glutamate (McCauley et al., 2020). Cultured cortical astrocytes from mice exhibited Per1/2 expression, indicating that glial cells can show circadian rhythms in clock-gene expression (Prolo et al., 2005). A recent study revealed that SCN astrocytes constituted the other “half’ of the circadian clock by releasing glutamate and elevating extracellular glutamate levels during the subjective night (Brancaccio et al., 2017). Boosting extracellular glutamate levels activates pre-synaptic NR2C complexes on dorsomedial SCN neurons, increasing the inhibitory tone within the circuit and suppressing SCN neuronal activity (Brancaccio et al., 2017). During the subjective day, the decrease in glutamate release by astrocytes and the resulting decrease in extracellular glutamate levels attenuate the repression of the spontaneous activity of SCN neurons (Brancaccio et al., 2017). Co-culture of SCN neurons provides astrocytes with a slightly more stable rhythm, providing evidence for the mutually supportive effect between neurons and astrocytes on SCN rhythms (Brancaccio et al., 2017; Tso et al., 2017). Another study revealed that the knockdown of Bmal1 in astrocytes alters daily locomotor activity and cognitive function through GABA signaling (Barca-Mayo et al., 2017). One of the most impressive demonstrations of astrocyte potency is that it can autonomously encode circadian information and instruct its neuronal partners to initiate and indefinitely maintain circadian patterns of neuronal activity and behavior (Brancaccio et al., 2019).

The circadian clock regulates the activities of other glial cells, such as microglia and oligodendrocytes. Microglia, specialized immune cells in the brain, monitor the central nervous system through their intricate branching and dynamic projections (Wake et al., 2009). Deletion of Bmal1, a transcriptional activator of the circadian clock, in both neuronal and glial cells, induces severe age-dependent astrogliosis in the cortex and hippocampus, while preserving intact circadian and sleep-wake rhythms (Musiek et al., 2013). Deletion of Rev-erbα, a nuclear receptor and circadian clock component, leads to spontaneous activation of microglia in the hippocampus, increased expression of pro-inflammatory transcripts, and secondary astrogliosis (Griffin et al., 2019). Mice exposed to dim light at night exacerbate the inflammatory response of microglia with lipopolysaccharides (LPS) (Fonken et al., 2013). Moreover, SD rats injected with LPS in the light phase show robust hippocampal microglia production compared to rats injected in the dark phase (Fonken et al., 2015), indicating circadian regulation of microglia activation. Oligodendrocytes are crucial contributors to neuronal function and provide metabolic support to neurons (Lee et al., 2012). Bellesi and colleagues found that about 2% of oligodendrocyte genes fluctuate as a function of circadian time (Bellesi et al., 2013). Preclinical and clinical studies of oligodendrocytes are warranted to understand how circadian rhythms affect the maturation and function of oligodendrocytes (Colwell and Ghiani, 2020) and to aid in the development of new therapeutic strategies and standards of patient care.

3.3. The free-running property of the circadian clock

As aforementioned, the rhythmic expression of clock genes in cells is driven by self-sustained TTFLs (Rosbash, 2021). Thus, an important property of the circadian clock is that its daily oscillations are autonomous and do not rely on the input of environmental cues. This was originally demonstrated in the studies by French astronomer Jean-Jacques d’Ortous de Mairan, who found that the leaves of the Mimosa plant moved with a periodicity of 24-h, even when the plant was moved to a room of constant darkness (Roenneberg and Merrow, 2005). To demonstrate the existence of a functional circadian clock, constant environment is always used as an experimental condition. The term “zeitgeber” (time giver or time cue) refers to environmental variables that can act as circadian time cues. Zeitgeber time (ZT) means the time based on environmental cues, such as the 12h/12h light/dark cycles. Circadian time (CT), in contrast, means the time based on the endogenous biological clock when organisms are kept in free-running, constant conditions. The onset of activity of day-active organisms is defined as circadian time zero (CT0), and the onset of activity of night-active organisms is CT12 (Eelderink-Chen et al., 2015). For mouse experiments, constant darkness is usually required to reveal differences caused by the circadian clock. However, housing animals in constant darkness can cause difficulties in animal husbandry, and thus, ZTs are used in a lot of animal studies (Bittman et al., 2013).

4. Circadian rhythms in extra-SCN brain regions

The rhythmic expression of clock genes throughout the brain plays a critical role in regulating everyday brain functions (Adamantidis et al., 2007; Mohawk et al., 2012). The master pacemaker located in the SCN synchronizes circadian oscillators in extra-SCN brain regions, allowing neurophysiological processes to be orchestrated with external environmental cues, such as the light-dark cycle. The environmental signal is relayed to other brain regions through the SCN. Some of these receive direct innervation from the SCN pacemaker, while others get indirect input from the SCN through poorly understood neural or humoral mechanisms. The extra-SCN regions can modulate the activity of the SCN, contributing to the adaptability and flexibility of the circadian system (Begemann et al., 2020). To date, most studies on non-SCN clocks have focused on peripheral tissues (liver, muscle, tissue culture, etc.), and circadian oscillators in other brain regions are inadequately studied. Several extra-SCN brain oscillators have been characterized in terms of circadian regulation and output. However, their specific physiological functions and their roles in circadian clock regulation remain elusive.

4.1. Rhythmic gene expression in various brain regions

Although the SCN serves as the master pacemaker, rhythmic gene expression is also observed in many brain regions. The neuronal activity rhythms within the SCN modulate the rhythms in different brain areas via direct and indirect projections and the modulation of hormones, such as cortisol and melatonin (Logan et al., 2022a). However, not all brain regions share the same rhythmic gene expression pattern. The phase, robustness, and perhaps the circadian period differs among different brain regions, reflecting their distinct functions and anatomical connections.

The olfactory bulb (OB) is responsible for processing odor-related information. The SCN entrains but does not sustain the circadian rhythm of the OB (Granados-Fuentes et al., 2004). The mammalian brain includes a diverse set of Per1 circadian oscillators that are either SCN-dependent or SCN-independent and present in specific brain regions (Abraham et al., 2005). The OB contains a master circadian pacemaker that drives rhythms in the piriform cortex and interacts with SCNs to coordinate other daily behaviors in mice (Granados-Fuentes et al., 2006). VIP signaling regulates the output of OB to maintain circadian rhythms in the olfactory system of mammals (Miller et al., 2014).

The striatum is a part of the brain that is involved in dopamine (DA) signaling, movement, motivation, and reward. In psychiatric disorders, disruptions in the rhythmic and different expression of genes can lead to striatal dysfunction and psychosis(Ketchesin et al., 2023). These findings are consistent with the evidence of the regulation of core circadian clock genes in the striatum by DA (Hood et al., 2010; Imbesi et al., 2009). Variable restriction feeding disrupts the daily oscillations of Per2 expression in the limbic forebrain and striatum in rats (Verwey and Amir, 2012).

The prefrontal cortex (PFC), responsible for cognitive and executive functions, and the hippocampus, involved in learning and memory, show pronounced rhythmic patterns. The molecular rhythms of PFC undergo significant changes with age, which can impact cognition, sleep, and mood in later stages of life (Chen et al., 2016). The dorsolateral PFC plays a critical role in the core symptoms of psychiatric disorders and exhibits diurnal rhythms in gene expression, while most of the identified transcripts in the dorsolateral PFC are not rhythmic in schizophrenic patients (Seney et al., 2019). Thus, rhythmic genes might indirectly modulate PFC-dependent functions in schizophrenia. Robust sex differences in PFC shed light on the mechanisms underlying how neurotransmission and synaptic function are regulated in a sex-specific, circadian rhythm-dependent manner (Logan et al., 2022b).

Although the hippocampus and piriform cortex exhibit the highest levels of Per1 gene expression, they do not display intrinsic rhythmicity. In contrast, tissues with comparable (e.g., SCN and OB) or lower (e.g., arcuate nucleus) expression can oscillate in vitro with a cyclic pattern (Abe et al., 2002). The limbic forebrain, involved in the regulation of motivation and emotional states, shows rhythmic patterns of Per2 protein expression over time (Amir and Stewart, 2009). The Per2 gene in the hippocampus regulates long-term potential (LTP) and the recall of certain forms of learned behavior (Wang et al., 2009b) and might be modulated by the DA D1 receptor-PKA-CREB signaling pathway (Kim et al., 2022). Stress induces changes in Per1 protein expression in the limbic forebrain and hypothalamus of rats (Al-Safadi et al., 2014). The cAMP/MAPK/CREB signaling pathway in the hippocampus undergoes circadian oscillations, which correlate with the ability of mice to consolidate and maintain contextual memory (Eckel-Mahan et al., 2008b). The hippocampal-dependent long-term memory (LTM) depends on the SCN-regulated circadian oscillations of the MAPK/AC pathway in the hippocampus (Phan et al., 2011). The intrinsic excitability of the dentate gyrus (DG) and synaptic recruitment mediated by G-protein signaling are subject to circadian regulation (Gonzalez et al., 2023). Overall, rhythmic gene expression in various brain regions plays an important role in regulating the timing of physiological and behavioral processes, and its dysregulation is associated with various neurological and psychiatric disorders.

The circadian rhythm of Per2 protein expression reveals a novel SCN-controlled oscillator in the oval nucleus of the bed nucleus of the stria terminalis (BNST-OV) (Amir et al., 2004). The central (CeA) and basolateral nuclei of the amygdala (BLA) exhibit contrasting circadian rhythms of Per2 protein expression (Lamont et al., 2005). Moreover, thyroidectomy alters the daily pattern of Per2 protein expression in the BNST-OV and CeA in rats (Amir and Robinson, 2006). Circadian neurons in the paraventricular nucleus of the hypothalamus (PVN) synchronize and sustain the daily glucocorticoid rhythm (Jones et al., 2021). The differential transcriptome rhythms observed in the human striatum help to understand the normal functioning of circadian rhythms and the pathological consequences of their disruption (Ketchesin et al., 2021). Overall, the field of rhythmic gene expression in the brain is rapidly evolving, and further research is needed to fully understand the underlying mechanisms and the functional significance of these patterns in different brain regions.

4.2. Diurnal oscillation of neurophysiological processes

Rhythmic expression drives rhythmic neuronal activity. The activity of neuronal and glial populations changes throughout the day/night cycle, regulating sleep/wake rhythms, cognition, hunger, reward, and other functions (Frank, 2019; Prevot, 2022). Overall, the circadian system controls a diurnal rhythm that governs various neurophysiological processes in the body. Many neurophysiological processes exhibit robust diurnal oscillations, which are thought to be important for the normal functionality of the brain and mental well-being. It is assumed that circadian rhythms affect long-term memory formation. Memory is strongly influenced by the time of day of formation and retrieval (Chaudhury and Colwell, 2002; Eckel-Mahan et al., 2008a), and operations that disrupt the circadian system can significantly damage memory processes (Smarr et al., 2014). Memory oscillates in a diurnal cycle (Gerstner and Yin, 2010; Rawashdeh et al., 2014; Urban et al., 2021), usually peaking during the active phase in rodents, although this varies by task and laboratory (Liu et al., 2022; Tsao et al., 2022; Winocur and Hasher, 2004). Disruptions to the circadian rhythm, such as those caused by shift work or jet lag, can impair memory formation and impact cognitive performance. The time of day represents a crucial biological variable in both rodent and human studies (Nelson et al., 2021).

We provide a detailed list of findings related to the role of circadian cycles in rodents, with a particular focus on learning and memory (Table 1). In brief, the circadian rhythm plays a key role in processing learning and memory, but its interaction with different types of learning and memory can yield inconsistent results. For instance, classic studies using the passive avoidance task have demonstrated that nocturnal rodents performed better during daylight times compared to the nighttime (Davies et al., 1973, 1974; Sandman et al., 1971; Stephens et al., 1967). In contrast, nocturnal rodents showed peak performance in the Morris water maze (MWM) test of spatial memory when they were active in the dark phase (Valentinuzzi et al., 2004b). Similarly, diurnal grass rats demonstrated optimal performance in the MWM during the light period (active phase) (Martin-Fairey and Nunez, 2014b). These findings suggest that the circadian rhythm differentially affects key brain circuits, such as the hippocampus and amygdala, which play crucial roles in different tasks. McCauley and his colleagues have shown that the circadian rhythm affects the surface NMDA receptors and astrocyte-synapse interactions of hippocampal pyramidal neurons, which in turn affects learning that depends on the hippocampus (McCauley et al., 2020). The impact of the circadian rhythm on these circuits depends on the specific nature and demands of the tasks being performed. For instance, visually guided behavior in mice reaches its peak sensitivity and stability at night without affecting the retinal ganglion cells, demonstrating how the day/night cycle mediates search strategies for visual cues (Koskela et al., 2020). The expression of core clock genes in the SCN is similar between nocturnal and diurnal rodents but differs in downstream regions such as the hippocampus and amygdala (Martin-Fairey and Nunez, 2014b; Ramanathan et al., 2010; Wang et al., 2009a). The persistence of memory depends on the time of training; nocturnal rodents showed better retention when trained at night in spatial learning and operant learning tasks (Gritton et al., 2012a). This effect appears to be dependent on the daily corticosterone rhythm, as adrenalectomy eliminates circadian changes in conditioned fear extinction learning (Woodruff et al., 2015a). The circadian rhythm has a complex impact on the corticosterone level, which is an important factor in controlling synaptic strength within the hippocampus (McCauley et al., 2020). In contrast, mice trained during their inactive phase in cued fear conditioning acquired the conditioning faster and had longer recall compared to mice trained during the active phase (Price and Obrietan, 2018). However, extinction was more likely to occur during the dark phase (Chaudhury and Colwell, 2002). A 24-h rhythm is observed in passive avoidance memory, with performance being better at a 24-h interval compared to the intervals in between (Holloway and Wansley, 1973a; Holloway and Wansley, 1973b). The effects of motor activity and risk assessment behaviors on passive avoidance memory formation and applied behavioral strategies are temporal and sex-dependent (Meseguer Henarejos et al., 2020).

Table 1.

The circadian regulation of rodent behaviors.

Task Animal Treatment Experimental Time Intertrial interval Results References
Passive avoidance Swiss Webster mouse ZT time 12:12h LD, ZT9 or 17 24h Light phase groups performed better than dark ones. (Stephens et al., 1967)
Passive avoidance Rat ZT time; melanocyte - stimulating hormone 12:12h LD, ZT6 or 13 48h Light phase groups performed better than dark ones. Melanocytes improved their performance in the dark. (Sandman et al., 1971)
Passive avoidance Sprague-Dawley rat ZT time 12:12h LD, every 4 h 48h Light phase groups performed better than dark ones. (Davies et al., 1973)
Passive avoidance Rat ZT time; ITI 12:12h LD, ZT2, 8, 14, 20 15min, 6h, 12h, 18h, 24h, 30h, 36h, 48h, 72h 15min and all 24h groups performed better than others. (Holloway and Wansley, 1973b)
Passive avoidance Rat ZT time; ITI 12:12h LD, ZT1-4.5 or 5.5-8 30sec-36h 30 sec, 15min, 12h, 24h, and 36h groups performed better than others. (Wansley and Holloway, 1976)
Passive avoidance Sprague-Dawley rat ZT time; ITI 12:12h LD; ZT1-4 or 7-10 15min or every 3-h up to 24h 15min, 9h, 12h, and 24h groups performed better than others. (Hunsicker and Mellgren, 1977)
Passive avoidance Sprague-Dawley rat SCN lesions; ITI 12:12h LD; ZT6-7 18h, 24h, or 30h 24h performance was better than others; SCN lesions did not follow this trend. (Stephan and Kovacevic, 1978)
Passive avoidance Rat +12h, +6h, or −6h shifts 12:12h LD, ZT0.5-2.5 48-h or 7 days Phase-shifted groups had impaired performance. (Tapp and Holloway, 1981)
Passive avoidance Wistar rat +4h or −6h shifts 14:10h LD→DL or DD 24h, 48h, 96h, 144h Phase-shifted groups had impaired performance. (Fekete et al., 1985)
Passive avoidance Wistar rat −6h shifts; ACTH analog ORG-2766 or DGAVP 14:10h LD→DL 24h, 48h,168h Phase-shifted groups had impaired performance; ACTH and DGAVP attenuated the deficits. (Fekete et al., 1986)
Passive avoidance Long-Evans rat ZT time; aging 12:12h DL, ZT13 or 23 24h or 6 weeks In the early dark phase, performance was better than in the late dark phase tested at 6 weeks, but not 24 h, later in the aged, but not young, group (Winocur and Hasher, 1999)
Fear conditioning C57BL/6J mouse ZT time 12:12h LD; ZT2 or 14 Every 24-h up to 5 days Dark phase groups performed better than light ones. (Valentinuzzi et al., 2001)
Fear conditioning (tone-cued) C-57/6J; C-3H; mouse ZT time 12:12h LD or DL or DD, ZT3 or 15 N.A. Light phase groups performed better than dark ones. (Chaudhury and Colwell, 2002)
Fear conditioning C57Bl/6 mouse ZT time and time cues 12:12h LD, ZT4-8 or 16-20 (tested at the same or different times) 24h, 36h, or 48h ZT4 groups had better performances than others. (Eckel-Mahan et al., 2008b)
Fear conditioning (tone-cued) Long-Evans rat Every +3h shift for 6 days(one session) or 4 sessions 12:12h LD, ZT3 24h No effects of phase shifts. (Craig and McDonald, 2008)
Fear conditioning C57Bl/6 mouse Phase shift before or after the LD 12:12h LD, ZT3 or 6 1 to 7days The phase-shifted group had impaired performance. (Loh et al., 2010)
Fear conditioning C57Bl/6 mouse SCN lesions 12:12h LD 2weeks SCN lesions impaired performance. (Phan et al., 2011)
Fear conditioning (extinction) Sprague-Dawley rat ZT time 12:12h LD, ZT4 or 16 24h Facilitated extinction more in the dark phase group than in the light one. (Woodruff et al., 2015b)
Fear conditioning C57BL/6J mouse +12h shift or bifurcated 6:6h LD 12;12h LD, ZT12 or 24 12h or 24h The Phase-shifted group showed impaired retrieval, while the bifurcated group showed impaired learning. (Harrison et al., 2017)
Fear conditioning C57Bl6/J mouse ZT time; ITI; Bmal1 deletion 12:12h LD→DD, CT4 or 16 30min, 42h or 54h The CT4 group had better long-term retrieval than CT16; Bmal1 deletion impaired performance. (Price and Obrietan, 2018)
Fear conditioning (extinction) Sprague-Dawley rat ZT time; Per1/2 knockdown in PFC 12:12h LD, ZT4 or 16 24h Facilitated extinction in the dark phase group; PFC Per1/2 knockdown eliminated this facilitation. (Woodruff et al., 2018)
Fear conditioning (extinction) C57Bl/6J mouse +8h shifted across 1 or 2 days 12:12h LD, ZT13 24h or 48h The phase-shifted group showed impaired fear of extinction. (Clark et al., 2020)
Fear conditioning Wistar rat ZT time; DD 12:12h LD or DD, ZT The DD group showed facilitated fear extinction, specifically during the light phase. (Asadian et al., 2022)
Fear conditioning Siberian hamster −3 h shift 16:8h LD, ZT16-20 48h The phase-shifted group showed impaired fear memory. (Steiger et al., 2022)
MWM Long-Evans rat +3h shift 12:12h LD, ZT11 2 or 10 days (probe) There were no effects on learning; the phase-shifted group showed impaired retention. (Devan et al., 2001)
MWM (delayed non-match-to-sample) Long-Evans rat ZT time and time cues; aging 12:12h DL, ZT13 or 23 0-80sec The early dark phase group performed better than the late dark phase group in the old animals, whereas the reverse was true for the young. (Winocur and Hasher, 2004)
MWM Wistar rat ZT time 12:12h LD, ZT3 or 15 30 min Dark phase groups performed better than light ones. (Valentinuzzi et al., 2004a)
MWM C57BL/6 mouse LL 12:12h LD or LL N.A. (no probe) The LL group showed impaired learning more than the LD one. (Fujioka et al., 2011)
MWM Diurnal grass rat ZT time 12:12h LD, ZT4 or 16 24h There were no effects on learning; the light phase groups performed better than the dark ones. (Martin-Fairey and Nunez, 2014a)
MWM ICR mouse Abnormal LD 12:12h, 3:5h or 5:3h LD N.A. Abnormal LD impaired the performance. (Li et al., 2017)
MWM Wistar Kyoto and SHR rat LD disturbance 12:12h LD, ZT1-2 24h (probe) The LD disturbance impaired the performance. (Wang et al., 2020)
MWM Wistar rat LL or DD; fluoxetine 12:12h LD→ LL or DD, ZT0-3 24h The LL group showed impaired performance, while fluoxetine reversed the deficits. (Sharma et al., 2021)
MWM Long-Evans rat 12:9h LD for 6 days 12:12h LD→12:9h LD, ZT10.5 18 days (probe) The 21-day group showed impaired memory, but not learning. (Deibel et al., 2022)
MWM Wistar rat ZT time; DD 12:12h LD or DD, ZT The DD group showed impaired spatial memory. (Asadian et al., 2022)
MWM Non-Tg and 3xTg-AD mice ZT time; 3xTg-AD 12:12h LD, ZT4 or 16 24h (probe) The light group learned faster than the dark one, while the dark group had better reversal learning than the light one. AD impaired probe and reversal memory. (Carvalho da Silva et al., 2022)
NOR Siberian hamster ZT time; arrhythmic circadian; PTZ (1 mg/kg, i.p. daily) 16:8h LD, ZT3, 7, 11, 15, 19, 23 60min Early light groups had impaired performance in controls; the arrhythmic group showed impaired performance; PTZ facilitated performance (Ruby et al., 2008)
NOR Wistar rat ZT time 12:12h LD, ZT2, 12, 14 or 20 Min or 24h No effects in time-of-day. (Takahashi et al., 2013)
NOR Djungarian hamster ZT time; arrhythmic circadian 14:10h LD, ZT4, 7, 13, 16, 19 1h ZT13, 16, and 19 groups performed better than ZT4 and 7 groups, and arrhythmic circadian impaired performance. (Müller et al., 2015)
NOR Djungarian hamster Arrhythmic circadian 14:10h LD 1h Arrhythmic circadian impaired performance (Weinert et al., 2016)
NOR C57BL/6 mouse ZT time; DD; SON lesions 12:12h LD, ZT0, 4, 8, 12, 16, 20 24h Early light and SCN lesioned groups showed impaired performance. (Shimizu et al., 2016)
NOR & object-displacement recognition C57BL/6J mouse ZT time; LL 12:12h LD, ZT6 or 18; 12:12h LD→LL, CT6 or 18 5min or 24h The light phase group performed better than the dark phase one. (Tam et al., 2017)
NOR B6/C3H mouse Ts65Dn; SCN lesions 12:12h LD, ZT6 24h SCN lesions did not affect memory in controls; Ts65Dn mice had impaired performance, while SCN lesions alleviated it in Ts65Dn mice. (Chuluun et al., 2020)
NOR (one object during the learning and removal in the test) C57BL/6N Crl mouse ZT time 12:12h LD or DD, ZT3.5 or 15.5 N.A. The light phase groups performed better than the dark ones in learning, but not memory. (McCauley et al., 2020)
NOR C57BL/6 mouse ZT time; dim light (ZT12-16) 12:12h LD, ZT2 or 14 12h The light phase group performed better than dark phase one; dim light exposure reversed this pattern. (Tam et al., 2021)
NOR Wistar rat LL or DD; fluoxetine 12:12h LD→ LL or DD, ZT0-3 24h The LL group showed impaired memory, while fluoxetine reversed the deficits. (Sharma et al., 2021)
NOR C57BL/6J mouse Cry1, Cry2 KO, AAV-Cry1 12:12h LD→DD, ZT20-22 24h AAV-Cry1 expression in the SCN attenuated the deficits of Cry1-Cry2 KO. (Maywood et al., 2021)
NOR C57BL/6 mouse ZT time 12:12h LD, ZT0, 6, 12, 18 12h or 24h No effects of time of day (the dark phase groups seemed to perform better in complex object recognition). (Gessner et al., 2022)
NLR C57BL/6J mouse ZT time; sex 12:12h LD→DD, CT4 or 16 30min The dark phase group performed better than the light one in males, while the opposite effect was found in females. (Goode et al., 2022)
Active avoidance Sprague-Dawley rat ZT time; ACTH 12:12h LD, ZT1.5-2.5 or 9-10 Variable ratio: 90 sec The late light phase group performed better than the early light phase one; ACTH facilitated performance (Pagano and Lovely, 1972)
Active avoidance (extinction) Rat ZT time; ITI 12:12h LD, ZT0-4, 7-10, 12-15, 18-21 15min, 6h, 12h, 18h, 24h 15min, 12h, and 24h groups had stronger resistance to extinction than the others. (Holloway and Sturgis, 1976)
Active avoidance (extinction) Rat ZT time; pinealectomy 12;12h LD, ZT3 or 15 60sec The dark phase group learned faster than the light one. (Catalá et al., 1985)
Shock avoidance operant task Sprague-Dawley rat ZT time 12:12h LD or DL, ZT2, 6 or 10, 14, 18, 22 N.A. Dark phase groups performed better in shock avoidance, but not responses, than light ones. (Ghiselli and Patton, 1976)
CPA (foot shock) Syrian hamster ITI 14:10h LD, 24, 48, 32, or 40h 24 and 48h groups were more effective than the 32 and 40h groups. (Cain et al., 2004a)
CPA (foot shock) Syrian hamster ZT time and time cues 14:10h LD, ZT4 or 13 24h Non-matched LD cues impaired performance (Cain et al., 2008)
CPA (foot shock) Syrian hamster ZT time and time cues; SCN lesions 14:10h LD, ZT2 or 11 24h Non-matched LD cues impaired performance: SCN lesions did not affect it. (Cain et al., 2012)
CPP (wheel running) Syrian hamster ZT time 14:10h LD, ZT4 or 13; LL, CT4 or 13 24h The effects were paired with LD cues. (Ralph et al., 2002)
CPP (food) Rat Time cue 12:12h LD, CT2 or 11 (tested at the same or different times) 24h No effects of time cues. (McDonald et al., 2002)
CPP (wheel running) Syrian hamster ZT time; SCN lesions 14:10h LD, ZT3 or 14 24h The effects were paired with LD cues, and SCN was not needed. (Ko et al., 2003)
CPP (food) Wistar and Long-Evans rats ZT time and time cues 12:12h LD, ZT3 or 11 (training at ZT11) 24h No effects were observed in the Long-Evan rats; non-matched LD cues affected the Wistar rats. (Cain et al., 2004b)
CPP Syrian hamster +6h shift every 3 days for 25 days 14:10h LD, ZT18 8 days The phase-shifted group showed impaired performance. (Gibson et al., 2010)
CPP (food) Long-Evans rat Every +3h shift for 6 days→LD 10days, repeated 4 times 12:12h LD 24h The phase-shifted group showed impaired performance. (McDonald et al., 2013)
CPP (morphine) Wistar rat ZT time and time cue 12:12h LD, ZT1-2 or 11-12 24h The late-light phase group showed a stronger preference than the early-light ones. (Khaksari et al., 2022)
Radial-arm maze Sprague-Dawley rat DL; theophylline (15mg/kg, i.p. daily) 12:12h LD or DL 7 or 14 days The dark phase group performed better than the light one; theophylline facilitated performance only in the light phase. (Hauber and Bareiß, 2001)
Radial-arm maze C3H/HeN mouse ZT time; Per1 depletion 12:12h LD, ZT2 or 14 24h The light phase group performed better than the dark ones; Per1 depletion reversed this pattern. (Rawashdeh et al., 2014)
T-maze operant task Wistar rat ZT time; SCN ablation 12:12h LD→DD, ZT3-4, or 10-11 Variable ratio 10 sec Intact learning without LD cues and SCN. (Mistlberger et al., 1996)
Spontaneous T-maze Siberian hamster ZT time; arrhythmic circadian; PTZ (0.3 or 1 mg/kg, i.p. daily) 16:8h LD, ZT3, 7, 11, 15, 19, 23 N.A. Late light and dark phase groups performed better than early light ones; PTZ improved performanc e in the arrhythmic circadian group. (Ruby et al., 2013)
Spontaneous T-maze Siberian hamster Arrhythmic circadian; SCN lesions 16:8h LD, N.A. Arrhythmic circadian lesions, but not SCN lesions, impaired performance. (Fernandez et al., 2014)
Spontaneous T-maze C57BL/6J mouse ZT time; diet 12:12h LD, ZT2 or 14 N.A. The dark phase group performed better than the light one; high fat and sucrose diets impaired it. (Davis et al., 2020)
Spontaneous T-maze C57BL/6J mouse ZT time; diet 12:12h LD, ZT2 or 14 N.A. The dark phase group performed better than the light one; a high-fat diet impaired it. (Davis et al., 2021)
Spontaneous T-maze C57BL/6J mouse ZT time; diet 12:12h LD, ZT2 or 14 N.A. The dark phase group performed better than the light one; a high-fat diet impaired it. (Davis et al., 2021)
Spontaneous Y-maze Wistar rat LL or DD; fluoxetine 12:12h LD→ LL or DD, ZT0-3 N.A. LL group showed impaired memory, while fluoxetine reversed the deficits. (Sharma et al., 2021)
Spontaneous T-maze C57BL/6J mouse ZT time; Tg-SwDI 12:12h LD, ZT2 or 14 N.A. The dark phase group performed better than the light one; the Tg-SwDI mouse had impaired performance. (Fusilier et al., 2021)
6-point alley T-maze C57Bl/6 mouse ZT time 12:12h LD, ZT2, 6, 10, 14, 18, 22 N.A. Dark phase groups performed better than light ones. (Hoffmann and Balschun, 1992)
Spontaneous Y-maze C57BL/6J mouse SCN lesions 12:12h LD N.A. SCN lesions did not affect performance. (Mulder et al., 2014)
Sustained attention operant task Sprague-Dawley rat ZT time 12:12h LD, ZT4 or 16 N.A. The dark phase group performed better than the light one. (Gritton et al., 2012b)
Sustained attention operant task Sprague-Dawley rat ZT time; SCN lesions 12:12h LD, ZT 4 or 16 N.A. The dark phase group performed better than the light one; SCN lesions impaired performance at ZT4. (Gritton et al., 2013)
Barnes maze Mouse ZT time; Bmal1 forebrain deletion 12:12h LD, ZT4 or 16 17days (probe) Bmal1 deletion led to learning and recall deficits at both times. (Price et al., 2016)
Barnes maze C57BL/6J mouse ZT time; elF4E or MNK depletion 12:12h LD, ZT6 or 18 24h (probe) Dark phase groups performed better than light ones; elF4E and MNK depletions affected learning but not memory. (Liu et al., 2022)
Social recognition Djungarian hamster Arrhythmic circadian; social ranking 14:10h LD 2min or 24h The social subordinate, but not dominant, animals had intact 24-h memory; arrhythmic circadian impairment impaired the memory (Weinert et al., 2016)
Social recognition C57BL/6N mouse ZT time; time cues; Bmal1 deletion 12:12h LD, ZT4 or 10 6h or 24h ZT10 showed impaired retrieval of weak social memory in WT; Bmal1 deletion impaired memory at both times. (Hasegawa et al., 2019)
Licking latency task Rat ITI 12:12h LD, ZT1.5-4 or 5.5-8 15min, 1-h or every 6h up to 36h 12h, 24h, and 36h groups performed better than others. (Wansley and Holloway, 1975)
Circadian time-place three-arms task C57BL/6J mouse ZT time; Cry depletion 12:12h LD, ZT2-3.5, 5-6.5 or 8-9.5 3h Animals were able to learn specific time-place rules, but the Cry gene knockout impaired it. (Van der Zee et al., 2008)

Abbreviations: ZT: zeitgeber time; CT: circadian time; LD or DL: light-dark cycle; DD: dark-dark cycle; MWM: Morris water maze; N.A. not applicable or within seconds; CTA: conditioned taste aversion; CPP: conditioned place preference; CPA: conditioned place avoidance; ACTH: adrenocorticotropic hormone; DGAVP: desglycinamide−9-(Arg8)-vasopressin; PTZ: pentylenetetrazol; SHR: spontaneously hypertensive rat; PFC; prefrontal cortex; Tg-SwDI: the human APP gene with K670N/M671L, E693Q and D694N mutations; 3xTg AD: the TauP301L, APPSwe, and γ-secretase (PS1M146V); NOR, novel object recognition with 24 h training-testing interval; TMh, time memory to heat exposure; cTPL, circadian time place memory; CFC, contextual fear conditioning; NLR, novel location recognition; CR, conditioned reflex; Ts65Dn: segmental trisomic 16 on the B6/C3h background

Circadian rhythms regulate mating behavior in many mammals (Antle and Silver, 2016). Male rodents prioritize sexual behavior during the dark phase (Mahoney and Smale, 2005), and female rodents are most receptive to sexual behavior during this time. Sensitivity and perception also exhibit daily rhythms, with high pain thresholds often observed during the dark phase, when nocturnal animals are most likely to encounter painful stimuli (Palada et al., 2020). There is also a relationship between attention and the time of day (Paolone et al., 2012). The time of day can influence drug-seeking behavior and behavioral sensitivity to drugs, which are influenced by the circadian system. Additionally, drugs of abuse can provide feedback to regulate circadian rhythms (Sleipness et al., 2005, 2007; Uz et al., 2005; Webb, 2017). Similarly, nocturnal rodents are more active at night, and aggressive behavior shows predictable patterns of daily occurrence (Todd and Machado, 2019). Natural variation in maternal behavioral expression in laboratory rodents favors increased pup contact and pup-directed behavior during the light phase (Jensen Peña and Champagne, 2013). These processes can significantly impact sleep, cognition, hormone secretion, appetite, and digestion.

4.3. The role of biological sex in the circadian clock

Biological sex can impact multiple aspects of physiological functioning, such as circadian rhythm and clock regulation (Krizo and Mintz, 2015). In most laboratory rodent species, the estrous cycle is a 4-5-day rhythm with cyclical changes in vaginal cytology and reproductive hormones (McQuillan et al., 2019). Circulating estrogen and progesterone levels in female rodents alter circadian rhythms of physiology and behavior (Nakamura et al., 2008a). Estrogen directly influences the circadian rhythms of Per2 expression in the rat uterus (Nakamura et al., 2008b). Circadian rhythms in humans, rodents, and some non-human primates change during adolescence and puberty (Logan et al., 2018). Sleep-wake cycles and melatonin rhythms begin to phase delay with the onset of puberty (Roenneberg et al., 2004) and coincide with sexual maturity (Hagenauer and Lee, 2013). Sex-specific structural and functional variations are present in brain regions pertinent to circadian systems. For example, the sexually dimorphic role of the circadian clock genes in the stratum has been demonstrated in regulating alcohol-drinking behavior. Deletion of Bmal1 from medium spiny neurons of the striatum leads to increased alcohol intake in male mice but decreased intake in females (de Zavalia et al., 2023; de Zavalia et al., 2021). Also, sex-dependent characteristics, such as sexually dimorphic physiology, are linked to circadian rhythm disorders in male-dominant brain disorders (Alachkar et al., 2022). Female mice respond to light pulses with more significant phase shifts than male mice (Kuljis et al., 2013), but genetic ablation of ERα in female mice results in an amplified phase-shifting response to light (Blattner and Mahoney, 2013). Similar results have been found for testosterone, with gonadectomized male mice exhibiting more significant phase-shifting responses than non-gonadectomized male mice (Karatsoreos et al., 2011). The SCN is sexually dimorphic and enriched with estrogen receptors (Hatcher et al., 2020), but male mice have significantly higher SCN activation during the day (Bailey and Silver, 2014). In most studies, the gross anatomy of the SCN is similar in females and males, but there are sex differences in general SCN morphology. The shape of the human SCN differs by gender, with the anteroposterior axis 40% longer in women and the mediolateral axis 35% more spherical in men (Fernández-Guasti et al., 2000). The developmental patterns of VIP and AVP neurons in the SCN also show sex differences in mice (Carmona-Alcocer et al., 2023). AVP signaling resets SCN molecular rhythms in a sexually dimorphic manner (Rohr et al., 2022). Androgen, estrogen, and progesterone receptors are all expressed in the human SCN, with men having more androgen receptors and women having more ESR1 (Fernández-Guasti et al., 2000). In ovariectomized female mice, estrogen receptor subtype 1 (ESR1) and ESR2 agonists can mimic the period-shortening effects of estradiol, with ESR1 agonists effective at lower doses (Royston et al., 2014). In contrast, androgen activation does not affect the period in rodents (McGinnis et al., 2007). Executing the influence of biological sex on circadian rhythm is essential for multiple domains, such as sleep medicine, chronobiology, and customized healthcare.

5. Circadian neurogenetics and brain disorders

Circadian rhythm disruption is associated with an increased risk of developing brain diseases. Circadian disruption can occur due to various factors, such as shift work, jet lag, irregular sleep schedules, exposure to artificial light at night, and certain medical conditions. The high prevalence of circadian rhythm disruption in neurodegenerative and psychiatric illnesses provides evidence for the brain’s susceptibility to clock dysfunction and its direct involvement in disease. Different diseases vary in severity, medications differ in specificity and sensitivity, and the instruments used to measure circadian rhythm disruptions in clinical settings also vary, making it challenging to characterize and understand these disruptions. Sleep disorders and circadian disruptions are linked to several neurodevelopmental disorders, including attention deficit hyperactivity disorder (ADHD), Alzheimer’s disease (AD), schizophrenia, and autism spectrum disorders (ASDs) (Logan and McClung, 2019). The investigation of molecular clock mechanisms in psychiatric diseases is gaining momentum, as evidence suggests that misalignment between the endogenous circadian rhythm and the sleep-wake cycle may contribute to various psychiatric disorders. Much of the data relating circadian disruption to human brain illnesses is correlational, given the challenges in establishing causal relationships in human studies. Thus, animal models are frequently used to study circadian disruption in the pathogenesis of disease. Obtaining such knowledge is important to advance our understanding of the disease’s pathogenic mechanisms and develop new therapeutic strategies.

5.1. Human sleep disorders associated with clock gene mutations

Rare gene variants have been reported to cause sleep conditions characterized by extreme sleep duration or timing (Weedon et al., 2022). Familial advanced sleep phase syndrome (FASPS) is a sleep disorder in which a person’s sleep-wake cycle is earlier than the typical sleep-wake cycle (Ashbrook et al., 2020). FASPS is a familial disorder that appears to be isolated as a single gene with an autosomal dominant pattern of inheritance (Reid et al., 2001). A variation in human sleep behavior can be attributed to a missense mutation in the human PER2 gene, which alters the circadian period (Toh et al., 2001). A human PER3 variant causes a circadian phenotype and is associated with a seasonal mood trait (Zhang et al., 2016). A mutation in the human CRY2 gene produces an advanced sleep phase (Hirano et al., 2016), while TIM mutations alter the phase response and lead to an advanced sleep phase as well (Kurien et al., 2019). The stability of PER2 is regulated by casein kinase (CK)1-dependent phosphorylation of residues with FASPS (Shanware et al., 2011). Another study has identified a missense mutation (T44A) in the human CKIdelta gene in the population of FASPS (Xu et al., 2005). Transgenic Drosophila carrying the human CKIdelta-T44A mutation showed a lengthened circadian period, but transgenic mice carrying the same mutation have a shorter circadian period.

Familial delayed sleep phase syndrome (FDSPS) is a delay in a person’s circadian rhythm. FDSPS is associated with a dominant coding variation in the human circadian clock gene CRY1 (Patke et al., 2017). The CRY1 variant is a gain-of-function mutation that causes lengthening of the period of molecular circadian rhythms. Structural polymorphisms of the human PER3 gene have been related to the pathogenesis of FDSPS (Archer et al., 2003; Ebisawa et al., 2001). The human CLOCK gene variation has also been associated with FDSPS (Ebisawa, 2007). Functional alterations induced by missense variants in the human CKIɛ gene have been observed and exhibit an inverse correlation with circadian rhythm sleep disorders (Takano et al., 2004).

Natural short sleep is a rare genetic trait in which individuals sleep less than average hours without daytime sleepiness or other sleep deprivation consequences. A rare mutation in the human β1AR gene (ADRB1) has been found to be associated with natural short sleep (Shi et al., 2019), and mutations in the GRM1 gene have been identified as contributors to the natural short sleep trait (Shi et al., 2021). Short sleep duration has also been reported to be caused by mutations in the DEC2/BHLHE41 gene (He et al., 2009). Mutations in the NPSR1 gene reduce sleep duration while preserving memory consolidation (Xing et al., 2019). Treatment for this condition may involve modifying sleep habits, such as maintaining a regular sleep schedule and avoiding exposure to bright light in the evening.

5.2. Clock gene disruption and other brain disorders

A mood disorder is a mental health condition that primarily affects emotional state. Circadian rhythm disturbances are common in individuals with mood disorders (Vadnie and McClung, 2017). Disruptions to the circadian rhythm, such as those caused by irregular sleep patterns, shift work, or jet lag, can lead to a range of negative health outcomes, including depression and bipolar disorder (Logan and McClung, 2019; Walker et al., 2020). Kumari and colleagues have shown that alterations in the light-dark cycle in diurnal and nocturnal rodents have different effects on the endocrine, inflammatory, and antioxidant systems associated with depressive-like behavior. Constant light adversely affects both species, but constant darkness primarily negatively affects diurnal squirrels (Kumari et al., 2021). Sleep disorders in major depression disorder (MDD) include increased sleep latency, decreased latency to the first REM sleep episode, and early morning wakeup, although little is known about the role of circadian clock genes in MDD. In individuals with depression, circadian dysfunction may manifest as sleep disturbances, including insomnia or hypersomnia, and changes in appetite and energy levels (Crouse et al., 2021; Germain and Kupfer, 2008). Similarly, individuals with bipolar disorder may experience disruptions to their sleep-wake cycles during episodes of mania or depression (McClung, 2013; Singla et al., 2022). Evening scores are higher in bipolar disorder (BPD) and schizophrenia/schizoaffective patients than in controls. This finding appears to be connected to age in BPD patients (i.e., younger BPD patients are more intense “evening types”), whereas schizophrenia/schizoaffective people tend to show stronger eveningness at all ages. Being identified as an “evening type” could explain some of the sleep difficulties described by BPD patients, as well as the increased severity of BPD, as demonstrated by a younger age at treatment initiation, a higher incidence of self-reported fast mood swings, and rapid-cycling mood changes (Küng et al., 2019). Research suggests that interventions aiming to regulate circadian rhythms, such as chronotherapy, may be effective in treating mood disorders (Ruan et al., 2021). In addition, maintaining regular sleep patterns and avoiding disruptions in circadian rhythms may help prevent the onset or worsening of these diseases. The circadian pattern of the human brain in MDD identifies different genes that may be involved in important circadian events, including sleep/wake cycles and metabolism (Li et al., 2013). Diurnal variation and sex differences in hippocampal physiology are associated with neurological disorders and present differently in men and women (Goode et al., 2022).

Schizophrenia is a chronic mental illness characterized by positive symptoms (distortions of normal function, such as delusions and hallucinations), negative symptoms (absence of normal behaviors, such as anhedonia, asocialness, and blunted affect), and cognitive deficits. Some studies have shown disruptions of circadian rhythms in people with schizophrenia, including disturbances in the sleep-wake cycle (Kaskie et al., 2017), changes in the timing of melatonin secretion (Duan et al., 2021; Monteleone et al., 1997), and altered activity patterns (Qi et al., 2022; Zhang et al., 2021). These disturbances are thought to contribute to the cognitive and emotional symptoms of schizophrenia, such as attention and memory deficits, as well as mood disorders (Mansour et al., 2006). The dorsolateral prefrontal cortex (PFC), known to be crucial for schizophrenia’s fundamental symptoms (Smucny et al., 2022), is believed to have molecular cycles that are disrupted by circadian rhythms. Research has shown that the evening chronotype may increase the risk of psychiatric problems (Zou et al., 2022). Overall, while the relationship between circadian dysfunction and schizophrenia is still being studied, there is growing evidence that disruptions in internal clocks may play a role in its development and progression.

Neurodevelopmental disorders affect the development of the nervous system, leading to abnormal brain function that may affect emotion, memory, self-control, and learning ability. One of the most common neurodevelopmental disorders of childhood, ADHD, is characterized by inattentiveness, impulsivity, and hyperactivity and is associated with high rates of co-occurring sleep problems and circadian alterations (Philipsen et al., 2006). Reductions in sleep quality, delays in the circadian phase, and evening preference are consistently reported in children and adults with ADHD (Coogan and McGowan, 2017), and these may be correlated with the severity of ADHD symptoms (Rybak et al., 2007). Abnormal melatonin rhythms are also reported in children with ADHD (Coogan and McGowan, 2017). In adults with ADHD, a loss of circadian gene expression rhythms in the oral mucosa accompanies delays in cortisol rhythm and reduced amplitude melatonin rhythms (Baird et al., 2012). Advancing melatonin rhythms using bright light therapy improves ADHD symptoms of hyperactivity and impulsivity in adults, independent of changes in sleep time, sleep efficiency, wake time, or waking during sleep (Fargason et al., 2017). There is growing evidence to suggest that individuals with ASD may also experience circadian dysfunction (Lorsung et al., 2021). It is well-recognized that individuals with ASD are more likely to have sleep problems compared to the general population, such as difficulty falling and staying asleep, and waking up prematurely (Koo et al., 2021; Lorsung et al., 2021). This may be due to an imbalance in melatonin secretion, which is regulated by the circadian rhythm system. Sleep problems are very common in children with ASD (Cortesi et al., 2010; Richdale and Schreck, 2009), although identifying underlying circadian rhythm abnormalities has proven challenging (Takase et al., 1998). The most consistent correlation found in prepubertal children, pubertal adolescents, and young adults with ASD is a decrease in melatonin levels at night (Kulman et al., 2000; Tordjman et al., 2005). It is hypothesized that this decrease in melatonin during early development leads to the accumulation of oxidative stress, which is harmful to the developing nervous system and increases the risk of ASD (Jin et al., 2018). Studies have shown that melatonin levels are often disrupted in people with ASD, which can lead to sleep problems and impaired cognitive function (Wu et al., 2020). Additionally, some research suggests that individuals with ASD may have altered patterns of structured activity (Okamoto et al., 2018; Postema et al., 2019). This can contribute to circadian dysfunction and further impact sleep, mood, and behavior in people with ASD. Several studies have investigated the use of timed light or melatonin supplements to help regulate circadian rhythms and improve sleep in people with ASD (Cheng et al., 2021; Fatemeh et al., 2022). Addressing circadian dysfunction in patients with ASD may be an important focus of intervention.

Neurodegenerative diseases involve progressive deterioration in the structure and function of the central or peripheral nervous systems. In healthy older adults, individuals with AD or PD exhibit significantly reduced melatonin rhythm amplitude and excessive sleepiness, as well as other sleep-wake cycle disorders, such as later sleep episodes (Gros and Videnovic, 2017; Uchida et al., 1996; Videnovic and Golombek, 2017). PD is strongly associated with REM sleep behavior disorder, excessive daytime sleepiness, insomnia, and restless legs syndrome (Barone et al., 2009; Chaudhuri et al., 2006). In fact, more than 80% of people with REM sleep behavior disorder are later diagnosed with PD or dementia (Iranzo et al., 2014; Schenck et al., 2013). Transgenic mice overexpressing α-syn exhibit reduced amplitude and greater fragmentation of locomotor activity rhythms during the active phase, which worsens with age (Kudo et al., 2011). In addition, SCN neuronal firing is reduced in early adulthood and gradually progresses with age, suggesting that the circadian pacemaker is weakened (Kudo et al., 2011) and that dopaminergic regulation of circadian rhythms may be lost through the striatum, midbrain, and SCN circuits (Grippo et al., 2017; Korshunov et al., 2017). In a cohort study of 239 patients with PD, sleep latency was significantly increased, sleep efficiency and REM were reduced along with melatonin levels, and there was a loss of BMAL1 serum expression level (Breen et al., 2014). In another study, circadian melatonin rhythm dysfunction led to excessive daytime sleepiness in PD (Videnovic et al., 2014). Extended-release melatonin is an effective treatment option for PD patients with poor sleep quality (Ahn et al., 2020). Dr. Alois Alzheimer described AD as a type of memory loss or dementia (Cipriani et al., 2011). Memory issues may initially manifest as recurrent or chronological experiences, but over time they can affect various aspects of daily functioning. Neurologists have long recognized the existence of different subtypes of AD, such as posterior cortical and logopedic aphasic variations (Elahi and Miller, 2017). Cognitive impairment and dementia are also associated with sleep and circadian rhythm disruption, and the intensity of sleep disruption is linked to the severity of dementia (Winsky-Sommerer et al., 2019). Bmal1 KO mice exhibit premature aging, neurodegeneration, and a reduced lifespan (Musiek et al., 2013). One study found that even a single night of sleep deprivation led to Aβ accumulation in the right hippocampus and thalamus of the human brain (Shokri-Kojori et al., 2018). However, these increases were negatively correlated with mood but not with a specific AD-associated genotype (APOE genotype). In a triple transgenic mouse model of AD, there was a simultaneous alteration of the diurnal variation of hippocampal synaptic plasticity, memory, and metabolism (Carvalho da Silva et al., 2022). A comparative study demonstrated that AD patients had reduced diurnal activity, increased nocturnal activity, and phase delay in body temperature rhythms, and sunset severity was positively correlated with lower amplitude and more phase delay rhythms (Volicer et al., 2001). A clinical study involving 189 elderly residents in group care facilities showed that regular light exposure and melatonin supplementation improved cognitive symptoms of dementia and reduced aggression (Riemersma-van der Lek et al., 2008). However, further research is needed to fully understand the nature and extent of circadian dysfunction in individuals with neurodegenerative disorders and to develop effective interventions.

6. Chronomedicine or Chronotherapy

Chronomedicine, or circadian medicine, is the branch of medicine that studies the effects of biorhythms and time-dependent processes on health and disease. Chronomedicine can be conceptualized as dealing with the prevention, causation, diagnosis, and treatment of human diseases, with a particular focus on the role that time plays in our physiology, endocrinology, metabolism, and behavior at various organizational levels. Circadian medicine is increasingly regarded as an important category in clinical practice, as basic and translational research demonstrates the fundamental significance of circadian rhythms in human health and disease. The function of the local clock supports the temporal optimization of memory processes, elucidating the potential of circadian therapeutic strategies in preventing and treating memory impairment (Hartsock and Spencer, 2020).

6.1. Food supplements that can influence the human circadian clock

Caffeine is the most widely consumed psychoactive drug in the world. Its wakefulness-inducing effects and ability to disrupt sleep have been well-established (Temple et al., 2017). Caffeine has been found to alter circadian clocks in mammalian cells in vitro, in mice ex vivo, and in vivo (Oike et al., 2011). Caffeine exerts its effects by blocking adenosine receptors in the brain (Reichert et al., 2022). Extracellular adenosine levels in different brain areas in rats were shown to increase during wakefulness and decrease during sleep (Huston et al., 1996). While caffeine does not entrain the biological clock, it can improve daytime alertness with non-24-hour rhythms in blind individuals (St Hilaire and Lockley, 2015). In a clinical study, caffeine was shown to reduce sleep efficiency, sleep duration, slow-wave sleep (SWS), and REM sleep during daytime recovery sleep (Carrier et al., 2009). Epidemiological studies have also demonstrated that caffeine affects the timing of circadian rhythms in humans (Burke et al., 2015; Landolt, 2015; Reichert et al., 2022). CLOCK gene polymorphisms play an important role in determining the response to caffeine therapy in premature neonates with apnea (Guo et al., 2021). Resetting the late timing of ‘night owls’ has been shown to have a positive impact on performance, mental health, and sleep timing in real-world settings (Facer-Childs et al., 2019). Similarly, to caffeine, polyphenols found in green and oolong tea have been found to alleviate circadian rhythm disturbances and improve the microbiome and overall health (Guo et al., 2019; Zhang et al., 2020). Probiotics, which are beneficial bacteria (Matenchuk et al., 2020; Takada et al., 2017), vitamin D, an important nutrient (Fallah et al., 2020; Majid et al., 2018), and magnesium, an essential mineral (Abbasi et al., 2012; van Ooijen and O’Neill, 2016), have also been shown to influence various physiological processes in the body, including the regulation of the circadian clock.

The pineal hormone melatonin is known to reset the circadian clock in animal and human studies (Carriedo-Diez et al., 2022). Exogenous melatonin normalizes circadian variation in human physiology and strongly impacts general health, especially in the elderly and shift workers (Kunz et al., 2004). Recently, melatonergic agents such as Circadin, Ramelteon (TAK-375), Tasimelteon, PD-6735 (TIK-301), and combined melatonergic/serotonergic drugs such as Agomelatine have become available in the clinic to treat insomnia and depression (Naveed et al., 2022). Tasimelteon entrained non-24-hour sleep-wake disorder in completely blind people, but continued treatment was necessary to maintain these improvements (Lockley et al., 2015). In addition, Tasimelteon had no clinically significant effect on next-day driving performance in healthy adults (Torres et al., 2022). In an animal study, idiopathic REM sleep behavior disorder was associated with altered clock gene expression and delayed melatonin secretion (Weissová et al., 2018). High-dose melatonin increases sleep duration during nighttime and daytime sleep episodes in older adults (Duffy et al., 2022). Young children may be highly sensitive to light one hour before bedtime, suggesting that the home lighting environment and its effect on circadian rhythm timing should be considered as a possible factor in behavioral sleep difficulties (Hartstein et al., 2023). The combination of bright light exposure in the early morning and exogenous melatonin in the evening may provide the greatest phase-shift treatment response (Burke et al., 2013). Yet small changes in ordinary light in the evening can significantly affect plasma melatonin concentrations and the entrainment phase of human circadian pacemakers (Zeitzer et al., 2000). In one clinical study, melatonin upregulated Bmal1 expression in AD patients, but Per1 levels remained unchanged (Delgado-Lara et al., 2020). Furthermore, the time of administration may influence the physiological response to melatonin (Lok et al., 2019).

6.2. Dosing time matters for pharmacokinetics

Pharmacokinetics (PK) deals with the movement of drugs in the body. The circadian clock significantly regulates PK depending on the time of day, a discovery that led to the concept of chronomedicine. Drugs can affect the body’s clock function, and these side effects may complicate their use for certain diseases. Severe alterations in circadian rhythms were observed in mice and human patients receiving chemotherapeutic drugs (Ortiz-Tudela et al., 2014), which can be divided into two categories: (i) patients who exhibited unintentional side effects or non-specific toxicity, leading to clock resetting; and (ii) targeted timing drugs. Growing evidence indicates that the effects of drugs are dependent on circadian timing (Table 2). In a cohort study of 14,840 patients with self-poisoning in Sri Lanka, a close link between the diurnal variation of poisoning and death was highlighted. If the poisoning occurred in the late morning, the likelihood of death was three times higher than if it happened in the late afternoon and evening. This difference does not appear to be explained by treatment but is thought to be influenced by intestinal P. glycoprotein (P-gp) and cytochrome P450 3A4 (CYP3A4) rhythms (Carroll et al., 2012).

Table 2.

Clinical studies on the morning vs. evening schedule of drug administration (≥ 50 subjects).

Drugs Subjects Type of Study Drug-Delivery Time Conclusions References
Simvastatin, 172 (33 male, 117 female) random Double blind, placebo-controlled study Evening vs Morning Evening is more effective than morning (Saito et al., 1991)
Simvastatin, 57 (27 men and 33 women), mean age 66 Randomized controlled trial Evening vs Morning Evening is more effective than morning (Wallace et al., 2003)
Valsartan, an ARBs 90 subjects (30 men and 60 women), Clinical trial Morning vs Evening No difference (Hermida et al., 2003)
Captopril, an ACEIs 121 (75 males, 46 females) Prospective, randomized, double-blind, placebo-controlled study Evening Restored the diurnal BP rhythm and decreased the elevated night/day BP ratio at bedtime administration (Qiu et al., 2005)
Phenytoin and carbamazepine 148 () 18-65 age Comparative study Evening Improved the response of diurnally active epileptic patients not responding to standard doses at bedtime administration (Yegnanarayan et al., 2006)
Telmisartan, ACEIs, and ARBs 215 (114 men and 101 women) (46.4+/−12.0 years) Randomized, double-blind, placebo-controlled trial Evening vs Morning Reduced BP during sleep at bedtime administration (Hermida et al., 2007)
Prednisone, a corticosteroid 288 (54·6 (11·2) 55·4 (41 men and 247 women), Randomized, double-blind Modified vs immediate Reduced morning stiffness of joints by modified release (Buttgereit et al., 2008)
Torsemide, a diuretic 113 (44 men and 69 women) (51.7+/−10.6 years) Randomized, double-blind Evening vs Morning More effective in lowering BP in patients with uncomplicated essential hypertension at bedtime (Hermida et al., 2008b)
Nifedipine, a CCBs 180 (52.5 +/−10.7 years) (86 men and 94 women) Randomized, double-blind Evening vs Morning Significantly reduced the bedtime edema at bedtime administration (Hermida et al., 2008c)
ACIE, ARB, CCB 250 (136 men and 114 women), 60.1±11.7 years of age Randomized, open-label, blinded endpoint (PROBE), parallel-group trial Evening vs Morning A larger reduction in BP at bedtime administration (Hermida et al., 2008a)
Vaccination (hepatitis A and influenza A) 164 (men and women) Clinical study Morning vs Afternoon Higher antibody response to both the hepatitis A and influenza A vaccines at morning vaccination (Phillips et al., 2008)
Olmesartan ACEIs and ARBs 123 (39 men and 84 women) (46.6+/−12.3 years) Randomized, double-blind Evening vs Morning Improved wake/asleep BP ratio at bedtime administration (Hermida et al., 2009)
Ramipril ACEIs and ARBs 115 (52 men and 63 women) (46.7+/−11.2 years) Randomized, double-blind Evening vs Morning Better nocturnal BP regulation at bedtime (Hermida and Ayala, 2009)
Amlodipine/valsartan, a CCBs and ARBs 203 (92 men/111 women), 56.7 +/− 12.5 years Single/Combined Evening vs Morning Improved the efficacy of lowering sleep BP and increasing sleep duration at bedtime administration (Hermida et al., 2010b)
ARBs, ACEIs, CCBs, β-blockers, and diuretics 2156 (1044 men/1112 women), MAPEC study Evening vs Morning The progressive decrease in asleep BP and increase in sleep-time were best achieved at bedtime therapy. (Hermida et al., 2010c)
Spirapril, an ACEIs 165 (65 men/100 women), 42.5 ± 13.9 [mean ± SD] yrs. of age) Open-label, parallel-group, blinded-endpoint Evening vs Morning Better sleep-time BP regulation at bedtime administration (Hermida et al., 2010a)
Valsartan/Hydrochlorothiazide, a diuretic, and ARBs 204 (95 men/109 women), (49.7 ± 11.1 years) An open-label, blinded endpoint Evening vs Morning Improved efficacy in lowering asleep BP and increased sleep time at bedtime administration (Hermida et al.,2011b)
Ezetimibe/simvastatin, statins 171 (87 in the morning and 84 in the evening administration group) > 18 years Randomized, crossover Morning vs Evening Morning administration was not inferior in lowering LDL-C (Yoon et al., 2011)
Antihypertensive drugs (ARBs e.g., valsartan; the ACEIs e.g., ramipril; and CCBs, e.g., amlodipine, etc.) 661 (396 men/265 women), (59.2 ± 13.5 years) PROBE trial Evening vs Morning Improved BP control and reduced risk of cardiovascular events at bedtime (Hermida et al., 2011a)
Aspirin, COX inhibitor 350 pregnant women (30.7 ± 5.3 years) Randomized, double-blind, placebo-controlled trial Evening vs Morning BP regulation and pregnancy outcome in high-risk pregnant women at bedtime administration (Ayala et al., 2013)
Prednisolone, a steroid 350 Primarily female (18–80 years) Randomized, double-blind, placebo-controlled trial Evening vs Morning Better reduction of morning stiffness of joints at bedtime administration (Buttgereit et al., 2013)
BP lowering drugs 2012 (976 men and 1,036 women), (52.7 ± 13.6 years) Prospective, PROBE trial Evening vs Morning Improved ambulatory BP control and reduced the risk of new-onset diabetes at bedtime administration (Hermida et al., 2016)
Influenza vaccination 276 adults (65+ age) Cluster-randomized trial Morning vs Afternoon Benefitted influenza antibody response at morning vaccination (Long et al., 2016)
ARB, ACEI, CCB, β-blocker, and/or diuretic 19084 (10614 men/8470 women), 60.5 ± 13.7 years of age Multi center, controlled, PROBE study Evening vs Morning Improved asleep ABP control and reduced CVD morbidity and mortality at bedtime administration (Hermida et al., 2019)
Aspirin, an NSAIDs 175 (59 women and 116 men), (59.8 years) Randomized controlled trial Evening vs Morning Reduced platelet aggregation at bedtime administration (Krasińska et al., 2019)
Ibuprofen, an NSAIDs 70 (men and women), (18-35 years) A randomized, double-blind, placebo-controlled trial Morning vs Morning Sufficient for pain management after surgical interventions at daytime administration (Tamimi et al., 2022)

There have been recent advances in understanding the phases in which circadian rhythms control xenobiotic metabolism. This circadian coordination of xenobiotic or drug metabolism helps increase the xenobiotic’s water solubility and efficient excretion, mainly through urine and bile (Dallmann et al., 2014; Lévi et al., 2010). The pharmacokinetic properties of all drugs can be affected by circadian variations, as shown by hundreds of compounds in laboratory rodents and humans (Levi and Schibler, 2007; Yang et al., 2015). To some extent, xenobiotics are influenced by circadian rhythmic resting activity patterns, such as mealtimes, general sleep-wake cycles, and physical activity, all of which regulate blood pressure and flow. The extent of influence depends not only on the route of administration and excretion but also on circadian rhythm regulation of gastric pH and gastrointestinal (GI) motility, both of which affect drug absorption. In contrast, blood flow and capillary perfusion affect the absorption of drugs from the GI tract, their distribution to tissues and target organs, and even their excretion through the glomerular filtration rate in the kidneys (Ballesta et al., 2017; Dallmann et al., 2016). The circadian clock in the blood-brain barrier regulates xenobiotic efflux, which may affect the response to drug treatments targeting the brain (Zhang et al., 2018). Our ability to pharmacologically target circadian time to alleviate or treat certain chronic diseases will provide the next advance in chronomedicine.

On the other hand, efforts are underway to develop new drugs that target the molecular clock. Robust circadian clock function is important for well-being, and enhancing the body’s clock function is considered to have the potential to treat several human diseases where circadian clock function is disrupted. The clock modulator nobiletin regulates circadian rhythms and physiology in female mice with AD (Kim et al., 2021). Since nobiletin can activate various clock-controlled metabolic genes involved in insulin signaling and mitochondrial function, including Igf1, Glut1, Insr, Irs1, Ucp2, and Ucp4, it may present a novel intervention target against AD progression. Some new drugs, such as Rev-Erbα agonists (Ercolani et al., 2015), are currently being developed to target circadian time coordination or molecular clocks in different tissues to enhance the robustness of these components and/or alter circadian phases. Timed dosing, meaning that drugs are administered at specific times of the day, is one approach to mitigating these side effects.

7. Conclusions

Elucidating the molecular mechanisms of the biological clock in model organisms has revealed the surprising conservation of genes that regulate human circadian rhythms. Emerging genetic studies on human clock genes suggest that genetic variation in these genes may be associated with neurological and mood disorders. This genetic variation in the human clock genes may also lead to phenotypic variation associated with disease pathways, such as mood disorders. Considering the time of day in chronotherapy can significantly improve the success rate of clinical trials and ultimately improve patient care. The question remains regarding whether clock genes primarily influence circadian clock function, or if they have functions independent of the circadian system. In addition, the circadian clock is cell-autonomous and distributed throughout the body, making therapeutic interventions targeting peripheral and central circadian oscillators crucial. Future explorations in these research areas, coupled with increased public awareness, may allow us to appropriately diagnose and treat human circadian rhythm disorders and improve lifestyle choices that align our physiological systems with the daily clock.

Highlights.

  • The cellular circadian clock is endogenously driven by transcription-translation feedback loops.

  • The suprachiasmatic nucleus (SCN) is the master circadian pacemaker in mammals.

  • Circadian rhythms are found in various extra-SCN brain regions and regulate neurophysiology and behavior.

  • Disruption of circadian rhythms and clock gene variations are associated with many human brain diseases.

  • The concept of chronomedicine and chronotherapy should be widely applied to promote health and treat diseases.

Acknowledgments

The work was supported by grants from NIH (NS118026, GM143260, DK109714) to R.C. We are grateful to Dr. Sobia for her assistance in designing the figures.

Footnotes

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

Conflicts of interest

The authors declare no conflicts of interest.

Data Availability

The article includes all data used in this work.

References

  1. Abbasi B, Kimiagar M, Sadeghniiat K, Shirazi MM, Hedayati M, Rashidkhani B, 2012. The effect of magnesium supplementation on primary insomnia in elderly: A double-blind placebo-controlled clinical trial. J Res Med Sci 17, 1161–1169. [PMC free article] [PubMed] [Google Scholar]
  2. Abe M, Herzog ED, Yamazaki S, Straume M, Tei H, Sakaki Y, Menaker M, Block GD, 2002. Circadian rhythms in isolated brain regions. Journal of Neuroscience 22, 350–356. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Abraham U, Prior JL, Granados-Fuentes D, Piwnica-Worms DR, Herzog ED, 2005. Independent circadian oscillations of Period1 in specific brain areas in vivo and in vitro. J Neurosci 25, 8620–8626. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Adamantidis AR, Zhang F, Aravanis AM, Deisseroth K, de Lecea L, 2007. Neural substrates of awakening probed with optogenetic control of hypocretin neurons. Nature 450, 420–424. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Agostino PV, Ferreyra GA, Murad AD, Watanabe Y, Golombek DA, 2004. Diurnal, circadian and photic regulation of calcium/calmodulin-dependent kinase II and neuronal nitric oxide synthase in the hamster suprachiasmatic nuclei. Neurochemistry International 44, 617–625. [DOI] [PubMed] [Google Scholar]
  6. Ahn JH, Kim M, Park S, Jang W, Park J, Oh E, Cho JW, Kim JS, Youn J, 2020. Prolonged-release melatonin in Parkinson’s disease patients with a poor sleep quality: A randomized trial. Parkinsonism Relat Disord 75, 50–54. [DOI] [PubMed] [Google Scholar]
  7. Aïoun J, Chambille I, Peytevin J, Martinet L, 1998. Neurons containing gastrin-releasing peptide and vasoactive intestinal polypeptide are involved in the reception of the photic signal in the suprachiasmatic nucleus ofthe Syrian hamster: an immunocytochemical ultrastructural study. Cell Tissue Res 291, 239–253. [DOI] [PubMed] [Google Scholar]
  8. Al-Safadi S, Al-Safadi A, Branchaud M, Rutherford S, Dayanandan A, Robinson B, Amir S, 2014. Stress-induced changes in the expression of the clock protein PERIOD1 in the rat limbic forebrain and hypothalamus: role of stress type, time of day, and predictability. PLoS One 9, e111166. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Alachkar A, Lee J, Asthana K, Vakil Monfared R, Chen J, Alhassen S, Samad M, Wood M, Mayer EA, Baldi P, 2022. The hidden link between circadian entropy and mental health disorders. Transl Psychiatry 12, 281. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Albrecht U, 2012. Timing to Perfection: The Biology of Central and Peripheral Circadian Clocks. Neuron 74, 246–260. [DOI] [PubMed] [Google Scholar]
  11. Alvord VM, Kantra EJ, Pendergast JS, 2021. Estrogens and the circadian system. Seminars in Cell & Developmental Biology. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Amir S, Lamont EW, Robinson B, Stewart J, 2004. A circadian rhythm in the expression of PERIOD2 protein reveals a novel SCN-controlled oscillator in the oval nucleus of the bed nucleus of the stria terminalis. J Neurosci 24, 781–790. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Amir S, Robinson B, 2006. Thyroidectomy alters the daily pattern of expression of the clock protein, PER2, in the oval nucleus of the bed nucleus of the stria terminalis and central nucleus of the amygdala in rats. Neurosci Lett 407, 254–257. [DOI] [PubMed] [Google Scholar]
  14. Amir S, Stewart J, 2009. Motivational modulation of rhythms of the expression of the clock protein PER2 in the limbic forebrain. Biological psychiatry 65, 829–834. [DOI] [PubMed] [Google Scholar]
  15. Antle MC, Silver R, 2005. Orchestrating time: arrangements of the brain circadian clock. Trends in Neurosciences 28, 145–151. [DOI] [PubMed] [Google Scholar]
  16. Antle MC, Silver R, 2016. Circadian insights into motivated behavior. Behavioral neuroscience of motivation, 137–169. [DOI] [PubMed] [Google Scholar]
  17. Antoun G, Cannon PB, Cheng H-YM, 2012. Regulation of MAPK/ERK Signaling and Photic Entrainment of the Suprachiasmatic Nucleus Circadian Clock by Raf Kinase Inhibitor Protein. The Journal of Neuroscience 32, 4867–4877. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Archer SN, Robilliard DL, Skene DJ, Smits M, Williams A, Arendt J, von Schantz M, 2003. A length polymorphism in the circadian clock gene Per3 is linked to delayed sleep phase syndrome and extreme diurnal preference. Sleep 26, 413–415. [DOI] [PubMed] [Google Scholar]
  19. Asadian N, Parsaie H, Vafaei AA, Dadkhah M, Omoumi S, Sedaghat K, 2022. Chronic light deprivation induces different effects on spatial and fear memory and hippocampal BDNF/TRKB expression during light and dark phases of rat diurnal rhythm. Behav Brain Res 418, 113638. [DOI] [PubMed] [Google Scholar]
  20. Ashbrook LH, Krystal AD, Fu Y-H, Ptáček LJ, 2020. Genetics of the human circadian clock and sleep homeostat. Neuropsychopharmacology 45, 45–54. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Aton SJ, Colwell CS, Harmar AJ, Waschek J, Herzog ED, 2005. Vasoactive intestinal polypeptide mediates circadian rhythmicity and synchrony in mammalian clock neurons. Nature Neuroscience 8, 476–483. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Ayala DE, Ucieda R, Hermida RC, 2013. Chronotherapy with low-dose aspirin for prevention of complications in pregnancy. Chronobiol Int 30, 260–279. [DOI] [PubMed] [Google Scholar]
  23. Bailey M, Silver R, 2014. Sex differences in circadian timing systems: implications for disease. Frontiers in neuroendocrinology 35, 111–139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Baird A, Coogan A, Siddiqui A, Donev R, Thome J, 2012. Adult attention-deficit hyperactivity disorder is associated with alterations in circadian rhythms at the behavioural, endocrine and molecular levels. Molecular psychiatry 17, 988–995. [DOI] [PubMed] [Google Scholar]
  25. Ballesta A, Innominato PF, Dallmann R, Rand DA, Lévi FA, 2017. Systems Chronotherapeutics. Pharmacological Reviews 69, 161–199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Barca-Mayo O, Pons-Espinal M, Follert P, Armirotti A, Berdondini L, De Pietri Tonelli D, 2017. Astrocyte deletion of Bmall alters daily locomotor activity and cognitive functions via GABA signalling. Nature Communications 8, 14336. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Barone P, Antonini A, Colosimo C, Marconi R, Morgante L, Avarello TP, Bottacchi E, Cannas A, Ceravolo G, Ceravolo R, Cicarelli G, Gaglio RM, Giglia RM, Iemolo F, Manfredi M, Meco G, Nicoletti A, Pederzoli M, Petrone A, Pisani A, Pontieri FE, Quatrale R, Ramat S, Scala R, Volpe G, Zappulla S, Bentivoglio AR, Stocchi F, Trianni G, Dotto PD, 2009. The PRIAMO study: A multicenter assessment of nonmotor symptoms and their impact on quality of life in Parkinson’s disease. Mov Disord 24, 1641–1649. [DOI] [PubMed] [Google Scholar]
  28. Basu P, Wensel AL, McKibbon R, Lefebvre N, Antle MC, 2016. Activation of Ml/4 receptors phase advances the hamster circadian clock during the day. Neurosci Lett 621, 22–27. [DOI] [PubMed] [Google Scholar]
  29. Begemann K, Neumann A-M, Oster H, 2020. Regulation and function of extra-SCN circadian oscillators in the brain. Acta Physiologica 229, e13446. [DOI] [PubMed] [Google Scholar]
  30. Bell-Pedersen D, Cassone VM, Earnest DJ, Golden SS, Hardin PE, Thomas TL, Zoran MJ, 2005. Circadian rhythms from multiple oscillators: lessons from diverse organisms. Nature Reviews Genetics 6, 544–556. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Bellesi M, Pfister-Genskow M, Maret S, Keles S, Tononi G, Cirelli C, 2013. Effects of sleep and wake on oligodendrocytes and their precursors. Journal of Neuroscience 33, 14288–14300. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Ben Haim L, Rowitch DH, 2017. Functional diversity of astrocytes in neural circuit regulation. Nat Rev Neurosci 18, 31–41. [DOI] [PubMed] [Google Scholar]
  33. Bittman EL, Kilduff TS, Kriegsfeld LJ, Szymusiak R, Toth LA, Turek FW, 2013. Animal care practices in experiments on biological rhythms and sleep: report of the Joint Task Force of the Society for Research on Biological Rhythms and the Sleep Research Society. J Am Assoc Lab Anim Sci 52, 437–443. [PMC free article] [PubMed] [Google Scholar]
  34. Blattner MS, Mahoney MM, 2013. Photic phase-response curve in 2 strains of mice with impaired responsiveness to estrogens. J Biol Rhythms 28, 291–300. [DOI] [PubMed] [Google Scholar]
  35. Bos NP, Mirmiran M, 1993. Effects of excitatory and inhibitory amino acids on neuronal discharges in the cultured suprachiasmatic nucleus. Brain Res Bull 31, 67–72. [DOI] [PubMed] [Google Scholar]
  36. Brancaccio M, Edwards MD, Patton AP, Smyllie NJ, Chesham JE, Maywood ES, Hastings MH, 2019. Cell-autonomous clock of astrocytes drives circadian behavior in mammals. Science 363, 187–192. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Brancaccio M, Patton AP, Chesham JE, Maywood ES, Hastings MH, 2017. Astrocytes Control Circadian Timekeeping in the Suprachiasmatic Nucleus via Glutamatergic Signaling. Neuron 93, 1420–1435.e1425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Breen DP, Vuono R, Nawarathna U, Fisher K, Shneerson JM, Reddy AB, Barker RA, 2014. Sleep and circadian rhythm regulation in early Parkinson disease. JAMA Neurol 71, 589–595. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Burke TM, Markwald RR, Chinoy ED, Snider JA, Bessman SC, Jung CM, Wright KP Jr., 2013. Combination of light and melatonin time cues for phase advancing the human circadian clock. Sleep 36, 1617–1624. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Burke TM, Markwald RR, McHill AW, Chinoy ED, Snider JA, Bessman SC, Jung CM, O’Neill JS, Wright KP Jr., 2015. Effects of caffeine on the human circadian clock in vivo and in vitro. Sci Transl Med 7, 305ra146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Butcher GQ, Doner J, Dziema H, Collamore M, Burgoon PW, Obrietan K, 2002. The p42/44 mitogen-activated protein kinase pathway couples photic input to circadian clock entrainment. J Biol Chem 277, 29519–29525. [DOI] [PubMed] [Google Scholar]
  42. Butcher GQ, Lee B, Cheng H-YM, Obrietan K, 2005. Light stimulates MSK1 activation in the suprachiasmatic nucleus via a PACAP-ERK/MAP kinase-dependent mechanism. J Neurosci 25, 5305–5313. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Buttgereit F, Doering G, Schaeffler A, Witte S, Sierakowski S, Gromnica-Ihle E, Jeka S, Krueger K, Szechinski J, Alten R, 2008. Efficacy of modified-release versus standard prednisone to reduce duration of morning stiffness of the joints in rheumatoid arthritis (CAPRA-1): a double-blind, randomised controlled trial. Lancet 371, 205–214. [DOI] [PubMed] [Google Scholar]
  44. Buttgereit F, Mehta D, Kirwan J, Szechinski J, Boers M, Alten RE, Supronik J, Szombati I, Romer U, Witte S, Saag KG, 2013. Low-dose prednisone chronotherapy for rheumatoid arthritis: a randomised clinical trial (CAPRA-2). Ann Rheum Dis 72, 204–210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Cahill GM, Menaker M, 1989. Effects of excitatory amino acid receptor antagonists and agonists on suprachiasmatic nucleus responses to retinohypothalamic tract volleys. Brain Res 479, 76–82. [DOI] [PubMed] [Google Scholar]
  46. Cain SW, Chalmers JA, Ralph MR, 2012. Circadian modulation of passive avoidance is not eliminated in arrhythmic hamsters with suprachiasmatic nucleus lesions. Behav Brain Res 230, 288–290. [DOI] [PubMed] [Google Scholar]
  47. Cain SW, Chou T, Ralph MR, 2004a. Circadian modulation of performance on an aversion-based place learning task in hamsters. Behav Brain Res 150, 201–205. [DOI] [PubMed] [Google Scholar]
  48. Cain SW, Ko CH, Chalmers JA, Ralph MR, 2004b. Time of day modulation of conditioned place preference in rats depends on the strain of rat used. Neurobiology of Learning and Memory 81, 217–220. [DOI] [PubMed] [Google Scholar]
  49. Cain SW, McDonald RJ, Ralph MR, 2008. Time stamp in conditioned place avoidance can be set to different circadian phases. Neurobiol Learn Mem 89, 591–594. [DOI] [PubMed] [Google Scholar]
  50. Cao R, Butcher GQ, Karelina K, Arthur JS, Obrietan K, 2013a. Mitogen- and stress-activated protein kinase 1 modulates photic entrainment of the suprachiasmatic circadian clock. Eur J Neurosci 37, 130–140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Cao R, Gkogkas CG, de Zavalia N, Blum ID, Yanagiya A, Tsukumo Y, Xu H, Lee C, Storch KF, Liu AC, Amir S, Sonenberg N, 2015. Light-regulated translational control of circadian behavior by eIF4E phosphorylation. Nat Neurosci 18, 855–862. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Cao R, Lee B, Cho HY, Saklayen S, Obrietan K, 2008. Photic regulation of the mTOR signaling pathway in the suprachiasmatic circadian clock. Mol Cell Neurosci 38, 312–324. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Cao R, Li A, Cho HY, Lee B, Obrietan K, 2010. Mammalian target of rapamycin signaling modulates photic entrainment of the suprachiasmatic circadian clock. J Neurosci 30, 6302–6314. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Cao R, Robinson B, Xu H, Gkogkas C, Khoutorsky A, Alain T, Yanagiya A, Nevarko T, Liu AC, Amir S, Sonenberg N, 2013b. Translational control of entrainment and synchrony of the suprachiasmatic circadian clock by mTOR/4E-BP1 signaling. Neuron 79, 712–724. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Card JP, Brecha N, Karten HJ, Moore RY, 1981. Immunocytochemical localization of vasoactive intestinal polypeptide-containing cells and processes in the suprachiasmatic nucleus of the rat: light and electron microscopic analysis. J Neurosci 1, 1289–1303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Carmona-Alcocer V, Brown LS, Anchan A, Rohr KE, Evans JA, 2023. Developmental patterning of peptide transcription in the central circadian clock in both sexes. Front Neurosci 17, 1177458. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Carriedo-Diez B, Tosoratto-Venturi JL, Cantón-Manzano C, Wanden-Berghe C, Sanz-Valero J, 2022. The Effects of the Exogenous Melatonin on Shift Work Sleep Disorder in Health Personnel: A Systematic Review. Int J Environ Res Public Health 19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Carrier J, Paquet J, Fernandez-Bolanos M, Girouard L, Roy J, Selmaoui B, Filipini D, 2009. Effects of caffeine on daytime recovery sleep: A double challenge to the sleep-wake cycle in aging. Sleep Med 10, 1016–1024. [DOI] [PubMed] [Google Scholar]
  59. Carroll R, Metcalfe C, Gunnell D, Mohamed F, Eddleston M, 2012. Diurnal variation in probability of death following self-poisoning in Sri Lanka—evidence for chronotoxicity in humans. International Journal of Epidemiology 41, 1821–1828. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Carvalho da Silva AM, Lemos C, Silva HB, Ferreira IL, Tomé AR, Rego AC, Cunha RA, 2022. Simultaneous Alteration of the Circadian Variation of Memory, Hippocampal Synaptic Plasticity, and Metabolism in a Triple Transgenic Mouse Model of Alzheimer’s Disease. Front Aging Neurosci 14, 835885. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Catalá MD, Pallardo F, Roman A, Villanueva P, Viña Giner JM, 1985. Effect of pinealectomy and circadian rhythm on avoidance behavior in the male rat. Physiology & Behavior 34, 327–333. [DOI] [PubMed] [Google Scholar]
  62. Chaudhuri KR, Martinez-Martin P, Schapira AH, Stocchi F, Sethi K, Odin P, Brown RG, Koller W, Barone P, MacPhee G, Kelly L, Rabey M, MacMahon D, Thomas S, Ondo W, Rye D, Forbes A, Tluk S, Dhawan V, Bowron A, Williams AJ, Olanow CW, 2006. International multicenter pilot study of the first comprehensive self-completed nonmotor symptoms questionnaire for Parkinson’s disease: the NMSQuest study. Mov Disord 21, 916–923. [DOI] [PubMed] [Google Scholar]
  63. Chaudhury D, Colwell CS, 2002. Circadian modulation of learning and memory in fear-conditioned mice. Behavioural Brain Research 133, 95–108. [DOI] [PubMed] [Google Scholar]
  64. Chen C-Y, Logan RW, Ma T, Lewis DA, Tseng GC, Sibille E, McClung CA, 2016. Effects of aging on circadian patterns of gene expression in the human prefrontal cortex. Proceedings of the National Academy of Sciences 113, 206–211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Chen D, Buchanan GF, Ding JM, Hannibal J, Gillette MU, 1999. Pituitary adenylyl cyclase-activating peptide: a pivotal modulator of glutamatergic regulation of the suprachiasmatic circadian clock. Proc Natl Acad Sci U S A 96, 13468–13473. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Cheng DCY, Ganner JL, Gordon CJ, Phillips CL, Grunstein RR, Comas M, 2021. The efficacy of combined bright light and melatonin therapies on sleep and circadian outcomes: A systematic review. Sleep Medicine Reviews 58, 101491. [DOI] [PubMed] [Google Scholar]
  67. Cheng H-YM, Obrietan K, Cain SW, Lee BY, Agostino PV, Joza NA, Harrington ME, Ralph MR, Penninger JM, 2004. Dexras1 Potentiates Photic and Suppresses Nonphotic Responses of the Circadian Clock. Neuron 43, 715–728. [DOI] [PubMed] [Google Scholar]
  68. Chuluun B, Pittaras E, Hong H, Fisher N, Colas D, Ruby NF, Heller HC, 2020. Suprachiasmatic lesions restore object recognition in down syndrome model mice. Neurobiol Sleep Circadian Rhythms 8, 100049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Cipriani G, Dolciotti C, Picchi L, Bonuccelli U, 2011. Alzheimer and his disease: a brief history. Neurological Sciences 32, 275–279. [DOI] [PubMed] [Google Scholar]
  70. Clark JW, Hoyer D, Cain SW, Phillips A, Drummond SPA, Jacobson LH, 2020. Circadian disruption impairs fear extinction and memory of conditioned safety in mice. Behav Brain Res 393, 112788. [DOI] [PubMed] [Google Scholar]
  71. Colwell CS, Ghiani CA, 2020. Potential Circadian Rhythms in Oligodendrocytes? Working Together Through Time. Neurochem Res 45, 591–605. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Colwell CS, Menaker M, 1992. NMDA as well as non-NMDA receptor antagonists can prevent the phase-shifting effects of light on the circadian system of the golden hamster. J Biol Rhythms 7, 125–136. [DOI] [PubMed] [Google Scholar]
  73. Coogan AN, McGowan NM, 2017. A systematic review of circadian function, chronotype and chronotherapy in attention deficit hyperactivity disorder. ADHD Attention Deficit and Hyperactivity Disorders 9, 129–147. [DOI] [PubMed] [Google Scholar]
  74. Coogan AN, Piggins HD, 2003. Circadian and photic regulation of phosphorylation of ERK1/2 and Elk-1 in the suprachiasmatic nuclei of the Syrian hamster. J Neurosci 23, 3085–3093. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Cortesi F, Giannotti F, Ivanenko A, Johnson K, 2010. Sleep in children with autistic spectrum disorder. Sleep Med 11, 659–664. [DOI] [PubMed] [Google Scholar]
  76. Craig LA, McDonald RJ, 2008. Chronic disruption of circadian rhythms impairs hippocampal memory in the rat. Brain Research Bulletin 76, 141–151. [DOI] [PubMed] [Google Scholar]
  77. Crouse JJ, Carpenter JS, Song YJC, Hockey SJ, Naismith SL, Grunstein RR, Scott EM, Merikangas KR, Scott J, Hickie IB, 2021. Circadian rhythm sleep-wake disturbances and depression in young people: implications for prevention and early intervention. Lancet Psychiatry 8, 813–823. [DOI] [PubMed] [Google Scholar]
  78. Dallmann R, Brown SA, Gachon F, 2014. Chronopharmacology: New Insights and Therapeutic Implications. Annual Review of Pharmacology and Toxicology 54, 339–361. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Dallmann R, Okyar A, Lévi F, 2016. Dosing-Time Makes the Poison: Circadian Regulation and Pharmacotherapy. Trends Mol Med 22, 430–445. [DOI] [PubMed] [Google Scholar]
  80. Davies JA, Navaratnam V, Redfern PH, 1973. A 24-hour rhythm in passive-avoidance behaviour in rats. Psychopharmacologia 32, 211–214. [DOI] [PubMed] [Google Scholar]
  81. Davies JA, Navaratnam V, Redfern PH, 1974. The effect of phase-shift on the passive avoidance response in rats and the modifying action of chlordiazepoxide. Br J Pharmacol 51, 447–451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Davis JA, Paul JR, McMeekin LJ, Nason SR, Antipenko JP, Yates SD, Cowell RM, Habegger KM, Gamble KL, 2020. High-Fat and High-Sucrose Diets Impair Time-of-Day Differences in Spatial Working Memory of Male Mice. Obesity (Silver Spring) 28, 2347–2356. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Davis JA, Paul JR, Yates SD, Cutts EJ, McMahon LL, Pollock JS, Pollock DM, Bailey SM, Gamble KL, 2021. Time-restricted feeding rescues high-fat-diet-induced hippocampal impairment. iScience 24, 102532. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. de Zavalia N, Ferraro S, Amir S, 2023. Sexually dimorphic role of circadian clock genes in alcohol drinking behavior. Psychopharmacology (Berl) 240, 431–440. [DOI] [PubMed] [Google Scholar]
  85. de Zavalia N, Schoettner K, Goldsmith JA, Solis P, Ferraro S, Parent G, Amir S, 2021. Bmal1 in the striatum influences alcohol intake in a sexually dimorphic manner. Commun Biol 4, 1227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Deibel SH, Higdon S, Cassell TTS, House-Denine ML, Giberson E, Webb IC, Thorpe CM, 2022. Impaired Morris water task retention following T21 light dark cycle exposure is not due to reduced hippocampal c-FOS expression. Front Behav Neurosci 16, 1025388. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Delgado-Lara DL, Gonzalez-Enriquez GV, Torres-Mendoza BM, González-Usigli H, Cárdenas-Bedoya J, Macías-Islas MA, de la Rosa AC, Jiménez-Delgado A, Pacheco-Moisés F, Cruz-Serrano JA, Ortiz GG, 2020. Effect of melatonin administration on the PER1 and BMAL1 clock genes in patients with Parkinson’s disease. Biomedicine & Pharmacotherapy 129, 110485. [DOI] [PubMed] [Google Scholar]
  88. Devan BD, Goad EH, Petri HL, Antoniadis EA, Hong NS, Ko CH, Leblanc L, Lebovic SS, Lo Q, Ralph MR, McDonald RJ, 2001. Circadian Phase-Shifted Rats Show Normal Acquisition but Impaired Long-Term Retention of Place Information in the Water Task. Neurobiology of Learning and Memory 75, 51–62. [DOI] [PubMed] [Google Scholar]
  89. Duan C, Jenkins ZM, Castle D, 2021. Therapeutic use of melatonin in schizophrenia: A systematic review. World J Psychiatry 11, 463–476. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Duffy JF, Wang W, Ronda JM, Czeisler CA, 2022. High dose melatonin increases sleep duration during nighttime and daytime sleep episodes in older adults. J Pineal Res 73, e12801. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Ebisawa T, 2007. Circadian Rhythms in the CNS and Peripheral Clock Disorders: Human Sleep Disorders and Clock Genes. Journal of Pharmacological Sciences 103, 150–154. [DOI] [PubMed] [Google Scholar]
  92. Ebisawa T, Uchiyama M, Kajimura N, Mishima K, Kamei Y, Katoh M, Watanabe T, Sekimoto M, Shibui K, Kim K, Kudo Y, Ozeki Y, Sugishita M, Toyoshima R, Inoue Y, Yamada N, Nagase T, Ozaki N, Ohara O, Ishida N, Okawa M, Takahashi K, Yamauchi T, 2001. Association of structural polymorphisms in the human period3 gene with delayed sleep phase syndrome. EMBO Rep 2, 342–346. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Eckel-Mahan KL, Phan T, Han S, Wang H, Chan GC, Scheiner ZS, Storm DR, 2008a. Circadian oscillation of hippocampal MAPK activity and cAmp: implications for memory persistence. Nat Neurosci 11, 1074–1082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Eckel-Mahan KL, Phan T, Han S, Wang H, Chan GCK, Scheiner ZS, Storm DR, 2008b. Circadian oscillation of hippocampal MAPK activity and cAMP: implications for memory persistence. Nat Neurosci 11, 1074–1082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Eelderink-Chen Z, Olmedo M, Bosman J, Merrow M, 2015. Using circadian entrainment to find cryptic clocks. Methods Enzymol 551, 73–93. [DOI] [PubMed] [Google Scholar]
  96. Elahi FM, Miller BL, 2017. A clinicopathological approach to the diagnosis of dementia. Nature Reviews Neurology 13, 457–476. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Ercolani L, Ferrari A, Mei CD, Parodi C, Wade M, Grimaldi B, 2015. Circadian clock: Time for novel anticancer strategies? Pharmacological Research 100, 288–295. [DOI] [PubMed] [Google Scholar]
  98. Facer-Childs ER, Middleton B, Skene DJ, Bagshaw AP, 2019. Resetting the late timing of ‘night owls’ has a positive impact on mental health and performance. Sleep Med 60, 236–247. [DOI] [PubMed] [Google Scholar]
  99. Fallah M, Askari G, Asemi Z, 2020. Is Vitamin D Status Associated with Depression, Anxiety and Sleep Quality in Pregnancy: A Systematic Review. Adv Biomed Res 9, 32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Fargason RE, Fobian AD, Hablitz LM, Paul JR, White BA, Cropsey KL, Gamble KL, 2017. Correcting delayed circadian phase with bright light therapy predicts improvement in ADHD symptoms: A pilot study. Journal of psychiatric research 91, 105–110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Fatemeh G, Sajjad M, Niloufar R, Neda S, Leila S, Khadijeh M, 2022. Effect of melatonin supplementation on sleep quality: a systematic review and meta-analysis of randomized controlled trials. J Neurol 269, 205–216. [DOI] [PubMed] [Google Scholar]
  102. Fekete M, Van Ree JM, De Wied D, 1986. The ACTH-(4-9) analog ORG 2766 and desglycinamide9-(Arg8)-vasopressin reverse the retrograde amnesia induced by disrupting circadian rhythms in rats. Peptides 7, 563–568. [DOI] [PubMed] [Google Scholar]
  103. Fekete M, van Ree JM, Niesink RJM, de Wied D, 1985. Disrupting circadian rhythms in rats induces retrograde amnesia. Physiology & Behavior 34, 883–887. [DOI] [PubMed] [Google Scholar]
  104. Fernández-Guasti A, Kruijver FP, Fodor M, Swaab DF, 2000. Sex differences in the distribution of androgen receptors in the human hypothalamus. J Comp Neurol 425, 422–435. [DOI] [PubMed] [Google Scholar]
  105. Fernandez F, Lu D, Ha P, Costacurta P, Chavez R, Heller HC, Ruby NF, 2014. Dysrhythmia in the suprachiasmatic nucleus inhibits memory processing. Science 346, 854–857. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Fonken LK, Frank MG, Kitt MM, Barrientos RM, Watkins LR, Maier SF, 2015. Microglia inflammatory responses are controlled by an intrinsic circadian clock. Brain Behav Immun 45, 171–179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Fonken LK, Weil ZM, Nelson RJ, 2013. Mice exposed to dim light at night exaggerate inflammatory responses to lipopolysaccharide. Brain Behav Immun 34, 159–163. [DOI] [PubMed] [Google Scholar]
  108. Frank MG, 2019. The Role of Glia in Sleep Regulation and Function. Handb Exp Pharmacol 253, 83–96. [DOI] [PubMed] [Google Scholar]
  109. Freeman GM Jr., Krock RM, Aton SJ, Thaben P, Herzog ED, 2013. GABA networks destabilize genetic oscillations in the circadian pacemaker. Neuron 78, 799–806. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Fujioka A, Fujioka T, Tsuruta R, Izumi T, Kasaoka S, Maekawa T, 2011. Effects of a constant light environment on hippocampal neurogenesis and memory in mice. Neurosci Lett 488, 41–44. [DOI] [PubMed] [Google Scholar]
  111. Fukushima T, Shimazoe T, Shibata S, Watanabe A, Ono M, Hamada T, Watanabe S, 1997. The involvement of calmodulin and Ca2+/calmodulin-dependent protein kinase II in the circadian rhythms controlled by the suprachiasmatic nucleus. Neurosci Lett 227, 45–48. [DOI] [PubMed] [Google Scholar]
  112. Fusilier AR, Davis JA, Paul JR, Yates SD, McMeekin LJ, Goode LK, Mokashi MV, Remiszewski N, van Groen T, Cowell RM, McMahon LL, Roberson ED, Gamble KL, 2021. Dysregulated clock gene expression and abnormal diurnal regulation of hippocampal inhibitory transmission and spatial memory in amyloid precursor protein transgenic mice. Neurobiol Dis 158, 105454. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Gamble KL, Allen GC, Zhou T, McMahon DG, 2007. Gastrin-releasing peptide mediates light-like resetting of the suprachiasmatic nucleus circadian pacemaker through cAMP response element-binding protein and Per1 activation. J Neurosci 27, 12078–12087. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Gannon RL, Rea MA, 1994. In situ hybridization of antisense mRNA oligonucleotides for AMPA, NMDA and metabotropic glutamate receptor subtypes in the rat suprachiasmatic nucleus at different phases of the circadian cycle. Brain Res Mol Brain Res 23, 338–344. [DOI] [PubMed] [Google Scholar]
  115. Gau D, Lemberger T, von Gall C, Kretz O, Le Minh N, Gass P, Schmid W, Schibler U, Korf HW, Schütz G, 2002. Phosphorylation of CREB Ser142 regulates light-induced phase shifts of the circadian clock. Neuron 34, 245–253. [DOI] [PubMed] [Google Scholar]
  116. Germain A, Kupfer DJ, 2008. Circadian rhythm disturbances in depression. Hum Psychopharmacol 23, 571–585. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Gerstner JR, Yin JC, 2010. Circadian rhythms and memory formation. Nat Rev Neurosci 11, 577–588. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Gessner N, Shinbashi M, Chuluun B, Heller C, Pittaras E, 2022. Handling, task complexity, time-of-day, and sleep deprivation as dynamic modulators of recognition memory in mice. Physiol Behav 251, 113803. [DOI] [PubMed] [Google Scholar]
  119. Ghiselli WB, Patton RA, 1976. Diurnal Variation in Performance of Free-Operant Avoidance Behavior of Rats. Psychological Reports 38, 83–90. [DOI] [PubMed] [Google Scholar]
  120. Gibson EM, Wang C, Tjho S, Khattar N, Kriegsfeld LJ, 2010. Experimental ‘jet lag’ inhibits adult neurogenesis and produces long-term cognitive deficits in female hamsters. PLoS One 5, e15267. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Gillespie CF, Mintz EM, Marvel CL, Huhman KL, Albers HE, 1997. GABA(A) and GABA(B) agonists and antagonists alter the phase-shifting effects of light when microinjected into the suprachiasmatic region. Brain Res 759, 181–189. [DOI] [PubMed] [Google Scholar]
  122. Ginty DD, Kornhauser JM, Thompson MA, Bading H, Mayo KE, Takahashi JS, Greenberg ME, 1993. Regulation of CREB phosphorylation in the suprachiasmatic nucleus by light and a circadian clock. Science 260, 238–241. [DOI] [PubMed] [Google Scholar]
  123. Golombek DA, Ralph MR, 1994. KN-62, an inhibitor of Ca2+/calmodulin kinase II, attenuates circadian responses to light. Neuroreport 5, 1638–1640. [DOI] [PubMed] [Google Scholar]
  124. Gonzalez JC, Lee H, Vincent AM, Hill AL, Goode LK, King GD, Gamble KL, Wadiche JI, Overstreet-Wadiche L, 2023. Circadian regulation of dentate gyrus excitability mediated by G-protein signaling. Cell Rep 42, 112039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Goode LK, Fusilier AR, Remiszewski N, Reeves JM, Abiraman K, Defenderfer M, Paul JR, McMahon LL, Gamble KL, 2022. Examination of Diurnal Variation and Sex Differences in Hippocampal Neurophysiology and Spatial Memory. eNeuro 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Granados-Fuentes D, Prolo LM, Abraham U, Herzog ED, 2004. The suprachiasmatic nucleus entrains, but does not sustain, circadian rhythmicity in the olfactory bulb. J Neurosci 24, 615–619. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Granados-Fuentes D, Tseng A, Herzog ED, 2006. A circadian clock in the olfactory bulb controls olfactory responsivity. J Neurosci 26, 12219–12225. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Griffin P, Dimitry JM, Sheehan PW, Lananna BV, Guo C, Robinette ML, Hayes ME, Cedeño MR, Nadarajah CJ, Ezerskiy LA, Colonna M, Zhang J, Bauer AQ, Burris TP, Musiek ES, 2019. Circadian clock protein Rev-erbα regulates neuroinflammation. Proc Natl Acad Sci U S A 116, 5102–5107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Grippo RM, Purohit AM, Zhang Q, Zweifel LS, Güler AD, 2017. Direct Midbrain Dopamine Input to the Suprachiasmatic Nucleus Accelerates Circadian Entrainment. Curr Biol 27, 2465–2475.e2463 [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Gritton HJ, Kantorowski A, Sarter M, Lee TM, 2012a. Bidirectional interactions between circadian entrainment and cognitive performance. Learning & Memory 19, 126–141. [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Gritton HJ, Kantorowski A, Sarter M, Lee TM, 2012b. Bidirectional interactions between circadian entrainment and cognitive performance. Learn Mem 19, 126–141. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Gritton HJ, Stasiak AM, Sarter M, Lee TM, 2013. Cognitive performance as a zeitgeber: cognitive oscillators and cholinergic modulation of the SCN entrain circadian rhythms. PLoS One 8, e56206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Gros P, Videnovic A, 2017. Sleep and circadian rhythm disorders in Parkinson’s disease. Current sleep medicine reports 3, 222–234. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Guo HL, Long JY, Hu YH, Liu Y, He X, Li L, Xia Y, Ding XS, Chen F, Xu J, Cheng R, 2021. Caffeine Therapy for Apnea of Prematurity: Role of the Circadian CLOCK Gene Polymorphism. Front Pharmacol 12, 724145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Guo T, Ho CT, Zhang X, Cao J, Wang H, Shao X, Pan D, Wu Z, 2019. Oolong Tea Polyphenols Ameliorate Circadian Rhythm of Intestinal Microbiome and Liver Clock Genes in Mouse Model. J Agric Food Chem 67, 11969–11976. [DOI] [PubMed] [Google Scholar]
  136. Hagenauer MH, Lee TM, 2013. Adolescent sleep patterns in humans and laboratory animals. Hormones and behavior 64, 270–279. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Harrison EM, Carmack SA, Block CL, Sun J, Anagnostaras SG, Gorman MR, 2017. Circadian waveform bifurcation, but not phase-shifting, leaves cued fear memory intact. Physiology & Behavior 169, 106–113. [DOI] [PubMed] [Google Scholar]
  138. Hartsock MJ, Spencer RL, 2020. Memory and the circadian system: Identifying candidate mechanisms by which local clocks in the brain may regulate synaptic plasticity. Neurosci Biobehav Rev 118, 134–162. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Hartstein LE, Diniz Behn C, Wright KP Jr., Akacem LD, Stowe SR, LeBourgeois MK, 2023. Evening Light Intensity and Phase Delay of the Circadian Clock in Early Childhood. J Biol Rhythms 38, 77–86. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Hasegawa S, Fukushima H, Hosoda H, Serita T, Ishikawa R, Rokukawa T, Kawahara-Miki R, Zhang Y, Ohta M, Okada S, Tanimizu T, Josselyn SA, Frankland PW, Kida S, 2019. Hippocampal clock regulates memory retrieval via Dopamine and PKA-induced GluA1 phosphorylation. Nature Communications 10, 5766. [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Hastings MH, Maywood ES, Brancaccio M, 2018. Generation of circadian rhythms in the suprachiasmatic nucleus. Nature Reviews Neuroscience 19, 453–469. [DOI] [PubMed] [Google Scholar]
  142. Hastings MH, Reddy AB, Maywood ES, 2003. A clockwork web: circadian timing in brain and periphery, in health and disease. Nature Reviews Neuroscience 4, 649–661. [DOI] [PubMed] [Google Scholar]
  143. Hatcher KM, Royston SE, Mahoney MM, 2020. Modulation of circadian rhythms through estrogen receptor signaling. European Journal of Neuroscience 51, 217–228. [DOI] [PubMed] [Google Scholar]
  144. Hauber W, Bareiß A, 2001. Facilitative effects of an adenosine A1/A2 receptor blockade on spatial memory performance of rats: selective enhancement of reference memory retention during the light period. Behavioural Brain Research 118, 43–52. [DOI] [PubMed] [Google Scholar]
  145. He Y, Jones CR, Fujiki N, Xu Y, Guo B, Holder JL, Rossner MJ, Nishino S, Fu Y-H, 2009. The Transcriptional Repressor DEC2 Regulates Sleep Length in Mammals. Science 325, 866–870. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Hermida RC, Ayala DE, 2009. Chronotherapy with the angiotensin-converting enzyme inhibitor ramipril in essential hypertension: improved blood pressure control with bedtime dosing. Hypertension 54, 40–46. [DOI] [PubMed] [Google Scholar]
  147. Hermida RC, Ayala DE, Chayan L, Mojon A, Fernandez JR, 2009. Administration-time-dependent effects of olmesartan on the ambulatory blood pressure of essential hypertension patients. Chronobiol Int 26, 61–79. [DOI] [PubMed] [Google Scholar]
  148. Hermida RC, Ayala DE, Fernández JR, Calvo C, 2007. Comparison of the efficacy of morning versus evening administration of telmisartan in essential hypertension. Hypertension 50, 715–722. [DOI] [PubMed] [Google Scholar]
  149. Hermida RC, Ayala DE, Fernández JR, Calvo C, 2008a. Chronotherapy Improves Blood Pressure Control and Reverts the Nondipper Pattern in Patients With Resistant Hypertension. Hypertension 51, 69–76. [DOI] [PubMed] [Google Scholar]
  150. Hermida RC, Ayala DE, Fontao MJ, Mojón A, Alonso I, Fernández JR, 2010a. Administration-time-dependent effects of spirapril on ambulatory blood pressure in uncomplicated essential hypertension. Chronobiol Int 27, 560–574. [DOI] [PubMed] [Google Scholar]
  151. Hermida RC, Ayala DE, Fontao MJ, Mojón A, Fernández JR, 2010b. Chronotherapy with valsartan/amlodipine fixed combination: improved blood pressure control of essential hypertension with bedtime dosing. Chronobiol Int 27, 1287–1303. [DOI] [PubMed] [Google Scholar]
  152. Hermida RC, Ayala DE, Mojón A, Chayán L, Domínguez MJ, Fontao MJ, Soler R, Alonso I, Fernández JR, 2008b. Comparison of the Effects on Ambulatory Blood Pressure of Awakening versus Bedtime Administration of Torasemide in Essential Hypertension. Chronobiology International 25, 950–970. [DOI] [PubMed] [Google Scholar]
  153. Hermida RC, Ayala DE, Mojón A, Fernández JR, 2008c. Chronotherapy with nifedipine GITS in hypertensive patients: improved efficacy and safety with bedtime dosing. Am J Hypertens 21, 948–954. [DOI] [PubMed] [Google Scholar]
  154. Hermida RC, Ayala DE, Mojón A, Fernández JR, 2010c. Influence of circadian time of hypertension treatment on cardiovascular risk: results of the MAPEC study. Chronobiol Int 27, 1629–1651. [DOI] [PubMed] [Google Scholar]
  155. Hermida RC, Ayala DE, Mojón A, Fernández JR, 2011a. Bedtime Dosing of Antihypertensive Medications Reduces Cardiovascular Risk in CKD. Journal of the American Society of Nephrology 22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Hermida RC, Ayala DE, Mojón A, Fernández JR, 2016. Bedtime ingestion of hypertension medications reduces the risk of new-onset type 2 diabetes: a randomised controlled trial. Diabetologia 59, 255–265. [DOI] [PubMed] [Google Scholar]
  157. Hermida RC, Ayala DE, Mojón A, Fontao MJ, Fernández JR, 2011b. Chronotherapy with valsartan/hydrochlorothiazide combination in essential hypertension: improved sleep-time blood pressure control with bedtime dosing. Chronobiol Int 28, 601–610. [DOI] [PubMed] [Google Scholar]
  158. Hermida RC, Calvo C, Ayala DE, Domínguez MJ, Covelo M, Fernández JR, Mojón A, López JE, 2003. Administration time-dependent effects of valsartan on ambulatory blood pressure in hypertensive subjects. Hypertension 42, 283–290. [DOI] [PubMed] [Google Scholar]
  159. Hermida RC, Crespo JJ, Domínguez-Sardiña M, Otero A, Moyá A, Ríos MT, Sineiro E, Castiñeira MC, Callejas PA, Pousa F, Salgado JL, Durán C, Sánchez JJ, Fernández JR, Mojón A, Ayala DE, Investigators, f.t.H.P., 2019. Bedtime hypertension treatment improves cardiovascular risk reduction: the Hygia Chronotherapy Trial. European Heart Journal 41, 4565–4576. [DOI] [PubMed] [Google Scholar]
  160. Hirano A, Shi G, Jones CR, Lipzen A, Pennacchio LA, Xu Y, Hallows WC, McMahon T, Yamazaki M, Ptáček FJ, Fu YH, 2016. A Cryptochrome 2 mutation yields advanced sleep phase in humans. Elife 5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Hoffmann HJ, Balschun D, 1992. Circadian differences in maze performance of C57BI/6 Ola mice. Behav Processes 27, 77–83. [DOI] [PubMed] [Google Scholar]
  162. Holloway FA, Sturgis RD, 1976. Periodic decrements in retrieval of the memory of nonreinforcement as reflected in resistance to extinction. Journal of Experimental Psychology: Animal Behavior Processes 2, 335–341. [Google Scholar]
  163. Holloway FA, Wansley R, 1973a. Multiphasic Retention Deficits at Periodic Intervals after Passive-Avoidance Learning. Science 180, 208–210. [DOI] [PubMed] [Google Scholar]
  164. Holloway FA, Wansley RA, 1973b. Multiple retention deficits at periodic intervals after active and passive avoidance learning. Behavioral Biology 9, 1–14. [DOI] [PubMed] [Google Scholar]
  165. Hood S, Cassidy P, Cossette M-P, Weigl Y, Verwey M, Robinson B, Stewart J, Amir S, 2010. Endogenous Dopamine Regulates the Rhythm of Expression of the Clock Protein PER2 in the Rat Dorsal Striatum via Daily Activation of D2 Dopamine Receptors. The Journal of Neuroscience 30, 14046–14058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Horikawa K, Shibata S, 2004. Phase-resetting response to (+)8-OH-DPAT, a serotonin 1A/7 receptor agonist, in the mouse in vivo. Neurosci Lett 368, 130–134. [DOI] [PubMed] [Google Scholar]
  167. Horikawa K, Yokota S.-i., Fuji K, Akiyama M, Moriya T, Okamura H, Shibata S, 2000. Nonphotic Entrainment by 5-HT1A/7 Receptor Agonists Accompanied by Reduced Per1 and Per2 mRNA Levels in the Suprachiasmatic Nuclei. The Journal of Neuroscience 20, 5867–5873. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Hunsicker JP, Mellgren RF, 1977. Multiple deficits in the retention of an appetitively motivated behavior across a 24-h period in rats. Animal Learning & Behavior 5, 14–16. [Google Scholar]
  169. Huston JP, Haas HL, Boix F, Pfister M, Decking U, Schrader J, Schwarting RK, 1996. Extracellular adenosine levels in neostriatum and hippocampus during rest and activity periods of rats. Neuroscience 73, 99–107. [DOI] [PubMed] [Google Scholar]
  170. Imbesi M, Yildiz S, Dirim Arslan A, Sharma R, Manev H, Uz T, 2009. Dopamine receptor-mediated regulation of neuronal “clock” gene expression. Neuroscience 158, 537–544. [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Iranzo A, Fernández-Arcos A, Tolosa E, Serradell M, Molinuevo JL, Valldeoriola F, Gelpi E, Vilaseca E, Sánchez-Valle R, Lladó A, Gaig C, Santamaría J, 2014. Neurodegenerative disorder risk in idiopathic REM sleep behavior disorder: study in 174 patients. PLoS One 9, e89741. [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Jensen Peña C, Champagne FA, 2013. Implications of temporal variation in maternal care for the prediction of neurobiological and behavioral outcomes in offspring. Behavioral neuroscience 127, 33. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Jin Y, Choi J, Won J, Hong Y, 2018. The Relationship between Autism Spectrum Disorder and Melatonin during Fetal Development. Molecules 23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Johnson JW, Ascher P, 1987. Glycine potentiates the NMDA response in cultured mouse brain neurons. Nature 325, 529–531. [DOI] [PubMed] [Google Scholar]
  175. Jones JR, Chaturvedi S, Granados-Fuentes D, Herzog ED, 2021. Circadian neurons in the paraventricular nucleus entrain and sustain daily rhythms in glucocorticoids. Nature Communications 12, 5763. [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Kallingal GJ, Mintz EM, 2014. Site-specific effects of gastrin-releasing peptide in the suprachiasmatic nucleus. Eur J Neurosci 39, 630–639. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Karatsoreos IN, Butler MP, LeSauter J, Silver R, 2011. Androgens Modulate Structure and Function of the Suprachiasmatic Nucleus Brain Clock. Endocrinology 152, 1970–1978. [DOI] [PMC free article] [PubMed] [Google Scholar]
  178. Kaskie RE, Graziano B, Ferrarelli F, 2017. Schizophrenia and sleep disorders: links, risks, and management challenges. Nature and science of sleep 9, 227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. Kawai N, Sakai N, Okuro M, Karakawa S, Tsuneyoshi Y, Kawasaki N, Takeda T, Bannai M, Nishino S, 2015. The sleep-promoting and hypothermic effects of glycine are mediated by NMDA receptors in the suprachiasmatic nucleus. Neuropsychopharmacology 40, 1405–1416. [DOI] [PMC free article] [PubMed] [Google Scholar]
  180. Ketchesin KD, Zong W, Hildebrand MA, Scott MR, Seney ML, Cahill KM, Shankar VG, Glausier JR, Lewis DA, Tseng GC, McClung CA, 2023. Diurnal Alterations in Gene Expression Across Striatal Subregions in Psychosis. Biol Psychiatry 93, 137–148. [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. Ketchesin KD, Zong W, Hildebrand MA, Seney ML, Cahill KM, Scott MR, Shankar VG, Glausier JR, Lewis DA, Tseng GC, McClung CA, 2021. Diurnal rhythms across the human dorsal and ventral striatum. Proc Natl Acad Sci U S A 118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Khaksari M, Nakhaei P, Khastar H, Bakhtazad A, Rahimi K, Garmabi B, 2022. Circadian fluctuation in curiosity is a risk factor for morphine preference. Biological Rhythm Research 53, 122–134. [Google Scholar]
  183. Kim E, Nohara K, Wirianto M, Escobedo G, Lim JY, Morales R, Yoo S-H, Chen Z, 2021. Effects of the Clock Modulator Nobiletin on Circadian Rhythms and Pathophysiology in Female Mice of an Alzheimer’s Disease Model, Biomolecules. [DOI] [PMC free article] [PubMed] [Google Scholar]
  184. Kim M, Custodio RJ, Lee HJ, Sayson LV, Ortiz DM, Kim BN, Kim HJ, Cheong JH, 2022. Per2 Expression Regulates the Spatial Working Memory of Mice through DRD1-PKA-CREB Signaling. Mol Neurobiol 59, 4292–4303. [DOI] [PubMed] [Google Scholar]
  185. Kim YI, Choi HJ, Colwell CS, 2006. Brain-derived neurotrophic factor regulation of N-methyl-D-aspartate receptor-mediated synaptic currents in suprachiasmatic nucleus neurons. J Neurosci Res 84, 1512–1520. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Ko CH, McDonald RJ, Ralph MR, 2003. The Suprachiasmatic Nucleus is not Required for Temporal Gating of Performance on a Reward-based Learning and Memory Task. Biological Rhythm Research 34, 177–192. [Google Scholar]
  187. Koo HW, Ismail J, Yang WW, Syed Zakaria SZ, 2021. Sleep Disturbances in Children With Autism Spectrum Disorder at a Malaysian Tertiary Hospital. Frontiers in Pediatrics 8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Koronowski KB, Sassone-Corsi P, 2021. Communicating clocks shape circadian homeostasis. Science 371, eabd0951. [DOI] [PMC free article] [PubMed] [Google Scholar]
  189. Korshunov KS, Blakemore LJ, Trombley PQ, 2017. Dopamine: A Modulator of Circadian Rhythms in the Central Nervous System. Front Cell Neurosci 11, 91. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Koskela S, Turunen T, Ala-Laurila P, 2020. Mice Reach Higher Visual Sensitivity at Night by Using a More Efficient Behavioral Strategy. Curr Biol 30, 42–53.e44. [DOI] [PubMed] [Google Scholar]
  191. Krasińska B, Paluszkiewicz L, Miciak-Lawicka E, Krasiński M, Rzymski P, Tykarski A, Krasiński Z, 2019. The effect of acetylsalicylic acid dosed at bedtime on the anti-aggregation effect in patients with coronary heart disease and arterial hypertension: A randomized, controlled trial. Cardiol J 26, 727–735. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. Krizo JA, Mintz EM, 2015. Sex Differences in Behavioral Circadian Rhythms in Laboratory Rodents. Frontiers in Endocrinology 5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Kudo T, Loh DH, Truong D, Wu Y, Colwell CS, 2011. Circadian dysfunction in a mouse model of Parkinson’s disease. Exp Neurol 232, 66–75. [DOI] [PubMed] [Google Scholar]
  194. Kuljis DA, Loh DH, Truong D, Vosko AM, Ong ML, McClusky R, Arnold AP, Colwell CS, 2013. Gonadal-and Sex-Chromosome-Dependent Sex Differences in the Circadian System. Endocrinology 154, 1501–1512. [DOI] [PMC free article] [PubMed] [Google Scholar]
  195. Kulman G, Lissoni P, Rovelli F, Roselli MG, Brivio F, Sequeri P, 2000. Evidence of pineal endocrine hypofunction in autistic children. Neuro Endocrinol Lett 21, 31–34. [PubMed] [Google Scholar]
  196. Kumari R, Verma V, Kronfeld-Schor N, Singaravel M, 2021. Differential response of diurnal and nocturnal mammals to prolonged altered light-dark cycle: a possible role of mood associated endocrine, inflammatory and antioxidant system. Chronobiol Int 38, 1618–1630. [DOI] [PubMed] [Google Scholar]
  197. Küng AL, Pham E, Cordera P, Hasler R, Aubry JM, Dayer A, Perroud N, Piguet C, 2019. Psychiatric disorders among offspring of patients with Bipolar and Borderline Personality Disorder. Journal of clinical psychology 75, 1810–1819. [DOI] [PMC free article] [PubMed] [Google Scholar]
  198. Kunz D, Mahlberg R, Müller C, Tilmann A, Bes F, 2004. Melatonin in patients with reduced REM sleep duration: two randomized controlled trials. J Clin Endocrinol Metab 89, 128–134. [DOI] [PubMed] [Google Scholar]
  199. Kurien P, Hsu PK, Leon J, Wu D, McMahon T, Shi G, Xu Y, Lipzen A, Pennacchio LA, Jones CR, Fu YH, Ptáček LJ, 2019. TIMELESS mutation alters phase responsiveness and causes advanced sleep phase. Proc Natl Acad Sci U S A 116, 12045–12053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  200. Lamont EW, Robinson B, Stewart J, Amir S, 2005. The central and basolateral nuclei of the amygdala exhibit opposite diurnal rhythms of expression of the clock protein Period2. Proc Natl Acad Sci U S A 102, 4180–4184. [DOI] [PMC free article] [PubMed] [Google Scholar]
  201. Landolt HP, 2015. CIRCADIAN RHYTHMS. Caffeine, the circadian clock, and sleep. Science 349,1289. [DOI] [PubMed] [Google Scholar]
  202. Lee C, Etchegaray JP, Cagampang FR, Loudon AS, Reppert SM, 2001. Posttranslational mechanisms regulate the mammalian circadian clock. Cell 107, 855–867. [DOI] [PubMed] [Google Scholar]
  203. Lee Y, Morrison BM, Li Y, Lengacher S, Farah MH, Hoffman PN, Liu Y, Tsingalia A, Jin L, Zhang P-W, Pellerin L, Magistretti PJ, Rothstein JD, 2012. Oligodendroglia metabolically support axons and contribute to neurodegeneration. Nature 487, 443–448. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Lévi F, Okyar A, Dulong S, Innominate PF, Clairambault J, 2010. Circadian Timing in Cancer Treatments. Annual Review of Pharmacology and Toxicology 50, 377–421. [DOI] [PubMed] [Google Scholar]
  205. Levi F, Schibler U, 2007. Circadian Rhythms: Mechanisms and Therapeutic Implications. Annual Review of Pharmacology and Toxicology 47, 593–628. [DOI] [PubMed] [Google Scholar]
  206. Li J-D, Hu W-P, Boehmer L, Cheng MY, Lee AG, Jilek A, Siegel JM, Zhou Q-Y, 2006. Attenuated Circadian Rhythms in Mice Lacking the Prokineticin 2 Gene. The Journal of Neuroscience 26, 11615–11623. [DOI] [PMC free article] [PubMed] [Google Scholar]
  207. Li JZ, Bunney BG, Meng F, Hagenauer MH, Walsh DM, Vawter MP, Evans SJ, Choudary PV, Cartagena P, Barchas JD, Schatzberg AF, Jones EG, Myers RM, Watson SJ Jr., Akil H, Bunney WE, 2013. Circadian patterns of gene expression in the human brain and disruption in major depressive disorder. Proc Natl Acad Sci U S A 110, 9950–9955. [DOI] [PMC free article] [PubMed] [Google Scholar]
  208. Li S, Li S, Li Y, Yan H, Huang C, Liu Q, 2017. Influence of circadian disorder on structures and functions of neurons in hippocampus of mice. Biological Rhythm Research 48, 639–645. [Google Scholar]
  209. Li X, Zhang C, Zhou QY, 2018. Overexpression of Prokineticin 2 in Transgenic Mice Leads to Reduced Circadian Behavioral Rhythmicity and Altered Molecular Rhythms in the Suprachiasmatic Clock. J Circadian Rhythms 16, 13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  210. Liou SY, Shibata S, Iwasaki K, Ueki S, 1986. Optic nerve stimulation-induced increase of release of 3H-glutamate and 3H-aspartate but not 3H-GABA from the suprachiasmatic nucleus in slices of rat hypothalamus. Brain Res Bull 16, 527–531. [DOI] [PubMed] [Google Scholar]
  211. Liu C, Reppert SM, 2000. GABA synchronizes clock cells within the suprachiasmatic circadian clock. Neuron 25, 123–128. [DOI] [PubMed] [Google Scholar]
  212. Liu D, Li J, Lin H, Lorsung E, Le N, Singla R, Mishra A, Fukunaga R, Cao R, 2022. Circadian activities of the brain MNK-eIF4E signalling axis contribute to diurnal rhythms of some cognitive functions. Eur J Neurosci 56, 3553–3569. [DOI] [PMC free article] [PubMed] [Google Scholar]
  213. Lockley SW, Dressman MA, Licamele L, Xiao C, Fisher DM, Flynn-Evans EE, Hull JT, Torres R, Lavedan C, Polymeropoulos MH, 2015. Tasimelteon for non-24-hour sleep-wake disorder in totally blind people (SET and RESET): two multicentre, randomised, double-masked, placebo-controlled phase 3 trials. Lancet 386, 1754–1764. [DOI] [PubMed] [Google Scholar]
  214. Logan RW, Hasler BP, Forbes EE, Franzen PL, Torregrossa MM, Huang YH, Buysse DJ, Clark DB, McClung CA, 2018. Impact of sleep and circadian rhythms on addiction vulnerability in adolescents. Biological psychiatry 83, 987–996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  215. Logan RW, McClung CA, 2019. Rhythms of life: circadian disruption and brain disorders across the lifespan. Nature Reviews Neuroscience 20, 49–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
  216. Logan RW, Xue X, Ketchesin KD, Hoffman G, Roussos P, Tseng G, McClung CA, Seney ML, 2022a. Sex differences in molecular rhythms in the human cortex. Biological psychiatry 91, 152–162. [DOI] [PMC free article] [PubMed] [Google Scholar]
  217. Logan RW, Xue X, Ketchesin KD, Hoffman G, Roussos P, Tseng G, McClung CA, Seney ML, 2022b. Sex Differences in Molecular Rhythms in the Human Cortex. Biol Psychiatry 91, 152–162. [DOI] [PMC free article] [PubMed] [Google Scholar]
  218. Loh DH, Navarro J, Hagopian A, Wang LM, Deboer T, Colwell CS, 2010. Rapid Changes in the Light/Dark Cycle Disrupt Memory of Conditioned Fear in Mice. PLoS One 5, e12546. [DOI] [PMC free article] [PubMed] [Google Scholar]
  219. Lok R, van Koningsveld MJ, Gordijn MCM, Beersma DGM, Hut RA, 2019. Daytime melatonin and light independently affect human alertness and body temperature. J Pineal Res 67, e12583. [DOI] [PMC free article] [PubMed] [Google Scholar]
  220. Long JE, Drayson MT, Taylor AE, Toellner KM, Lord JM, Phillips AC, 2016. Morning vaccination enhances antibody response over afternoon vaccination: A cluster-randomised trial. Vaccine 34, 2679–2685. [DOI] [PMC free article] [PubMed] [Google Scholar]
  221. Lorsung E, Karthikeyan R, Cao R, 2021. Biological Timing and Neurodevelopmental Disorders: A Role for Circadian Dysfunction in Autism Spectrum Disorders. Front Neurosci 15, 642745. [DOI] [PMC free article] [PubMed] [Google Scholar]
  222. Lowrey PL, Takahashi JS, 2000. GENETICS OF THE MAMMALIAN CIRCADIAN SYSTEM: Photic Entrainment, Circadian Pacemaker Mechanisms, and Posttranslational Regulation. Annual Review of Genetics 34, 533–562. [DOI] [PubMed] [Google Scholar]
  223. Mahoney MM, Smale L, 2005. A daily rhythm in mating behavior in a diurnal murid rodent Arvicanthis niloticus. Hormones and behavior 47, 8–13. [DOI] [PubMed] [Google Scholar]
  224. Majid MS, Ahmad HS, Bizhan H, Hosein HZM, Mohammad A, 2018. The effect of vitamin D supplement on the score and quality of sleep in 20-50 year-old people with sleep disorders compared with control group. Nutr Neurosci 21, 511–519. [DOI] [PubMed] [Google Scholar]
  225. Mansour HA, Wood J, Logue T, Chowdari KV, Dayal M, Kupfer DJ, Monk TH, Devlin B, Nimgaonkar VL, 2006. Association study of eight circadian genes with bipolar I disorder, schizoaffective disorder and schizophrenia. Genes, Brain and Behavior 5, 150–157. [DOI] [PubMed] [Google Scholar]
  226. Martin-Fairey CA, Nunez AA, 2014a. Circadian modulation of memory and plasticity gene products in a diurnal species. Brain Res 1581, 30–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
  227. Martin-Fairey CA, Nunez AA, 2014b. Circadian modulation of memory and plasticity gene products in a diurnal species. Brain research 1581, 30–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
  228. Matenchuk BA, Mandhane PJ, Kozyrskyj AL, 2020. Sleep, circadian rhythm, and gut microbiota. Sleep Medicine Reviews 53, 101340. [DOI] [PubMed] [Google Scholar]
  229. Maywood ES, Chesham JE, Winsky-Sommerer R, Hastings MH, 2021. Restoring the Molecular Clockwork within the Suprachiasmatic Hypothalamus of an Otherwise Clockless Mouse Enables Circadian Phasing and Stabilization of Sleep-Wake Cycles and Reverses Memory Deficits. J Neurosci 41, 8562–8576. [DOI] [PMC free article] [PubMed] [Google Scholar]
  230. Maywood ES, O’Neill J, Wong GKY, Reddy AB, Hastings MH, 2006. Circadian timing in health and disease, in: Kalsbeek A, Fliers E, Hofman MA, Swaab DF, van Someren EJW, Buijs RM (Eds.), Progress in Brain Research. Elsevier, pp. 253–269. [DOI] [PubMed] [Google Scholar]
  231. McCauley JP, Petroccione MA, D’Brant LY, Todd GC, Affinnih N, Wisnoski JJ, Zahid S, Shree S, Sousa AA, De Guzman RM, Migliore R, Brazhe A, Leapman RD, Khmaladze A, Semyanov A, Zuloaga DG, Migliore M, Scimemi A, 2020. Circadian Modulation of Neurons and Astrocytes Controls Synaptic Plasticity in Hippocampal Area CA1. Cell Rep 33, 108255. [DOI] [PMC free article] [PubMed] [Google Scholar]
  232. McClung CA, 2013. How Might Circadian Rhythms Control Mood? Let Me Count the Ways. Biological Psychiatry 74, 242–249. [DOI] [PMC free article] [PubMed] [Google Scholar]
  233. McDonald RJ, Hong NS, Ray C, Ralph MR, 2002. No Time of Day Modulation or Time Stamp on Multiple Memory Tasks in Rats. Learning and Motivation 33, 230–252. [Google Scholar]
  234. McDonald RJ, Zelinski EL, Keeley RJ, Sutherland D, Fehr L, Hong NS, 2013. Multiple effects of circadian dysfunction induced by photoperiod shifts: Alterations in context memory and food metabolism in the same subjects. Physiology & Behavior 118, 14–24. [DOI] [PubMed] [Google Scholar]
  235. McGinnis MY, Lumia AR, Tetel MJ, Molenda-Figueira HA, Possidente B, 2007. Effects of anabolic androgenic steroids on the development and expression of running wheel activity and circadian rhythms in male rats. Physiology & Behavior 92, 1010–1018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  236. McNeill JK, Walton JC, Albers HE, 2018. Functional Significance ofthe Excitatory Effects of GABA in the Suprachiasmatic Nucleus. J Biol Rhythms 33, 376–387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  237. McQuillan HJ, Han SY, Cheong I, Herbison AE, 2019. GnRH Pulse Generator Activity Across the Estrous Cycle of Female Mice. Endocrinology 160, 1480–1491. [DOI] [PubMed] [Google Scholar]
  238. Meseguer Henarejos AB, Popović N, Bokonjić D, Morales-Delgado N, Alonso A, Caballero Bleda M, Popović M, 2020. Sex and Time-of-Day Impact on Anxiety and Passive Avoidance Memory Strategies in Mice. Front Behav Neurosci 14, 68. [DOI] [PMC free article] [PubMed] [Google Scholar]
  239. Meyer-Spasche A, Piggins HD, 2004. Vasoactive intestinal polypeptide phase-advances the rat suprachiasmatic nuclei circadian pacemaker in vitro via protein kinase A and mitogen-activated protein kinase. Neurosci Lett 358, 91–94. [DOI] [PubMed] [Google Scholar]
  240. Mikkelsen JD, Larsen PJ, O’Hare MMT, Wiegand SJ, 1991. Gastrin releasing peptide in the rat suprachiasmatic nucleus: An immunohistochemical, chromatographic and radioimmunological study. Neuroscience 40, 55–66. [DOI] [PubMed] [Google Scholar]
  241. Miller JE, Granados-Fuentes D, Wang T, Marpegan L, Holy TE, Herzog ED, 2014. Vasoactive intestinal polypeptide mediates circadian rhythms in mammalian olfactory bulb and olfaction. J Neurosci 34, 6040–6046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  242. Mintz EM, Jasnow AM, Gillespie CF, Huhman KL, Albers HE, 2002. GABA interacts with photic signaling in the suprachiasmatic nucleus to regulate circadian phase shifts. Neuroscience 109, 773–778. [DOI] [PubMed] [Google Scholar]
  243. Mistlberger RE, de Groot MHM, Bossert JM, Marchant EG, 1996. Discrimination of circadian phase in intact and suprachiasmatic nuclei-ablated rats. Brain Research 739, 12–18. [DOI] [PubMed] [Google Scholar]
  244. Mohawk JA, Green CB, Takahashi JS, 2012. Central and peripheral circadian clocks in mammals. Annu Rev Neurosci 35, 445–462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  245. Monteleone P, Natale M, La Rocca A, Maj M, 1997. Decreased nocturnal secretion of melatonin in drug-free schizophrenics: no change after subchronic treatment with antipsychotics. Neuropsychobiology 36, 159–163. [DOI] [PubMed] [Google Scholar]
  246. Moore RY, Speh JC, 1993. GABA is the principal neurotransmitter of the circadian system. Neurosci Lett 150, 112–116. [DOI] [PubMed] [Google Scholar]
  247. Mordel J, Karnas D, Inyushkin A, Challet E, Pévet P, Meissl H, 2011. Activation of glycine receptor phase-shifts the circadian rhythm in neuronal activity in the mouse suprachiasmatic nucleus. The Journal of Physiology 589, 2287–2300. [DOI] [PMC free article] [PubMed] [Google Scholar]
  248. Mulder CK, Papantoniou C, Gerkema MP, Van Der Zee EA, 2014. Neither the SCN nor the adrenals are required for circadian time-place learning in mice. Chronobiology International 31, 1075–1092. [DOI] [PMC free article] [PubMed] [Google Scholar]
  249. Müller L, Fritzsche P, Weinert D, 2015. Novel object recognition of Djungarian hamsters depends on circadian time and rhythmic phenotype. Chronobiol Int 32, 458–467. [DOI] [PubMed] [Google Scholar]
  250. Musiek ES, Lim MM, Yang G, Bauer AQ, Qi L, Lee Y, Roh JH, Ortiz-Gonzalez X, Dearborn JT, Culver JP, Herzog ED, Hogenesch JB, Wozniak DF, Dikranian K, Giasson BI, Weaver DR, Holtzman DM, Fitzgerald GA, 2013. Circadian clock proteins regulate neuronal redox homeostasis and neurodegeneration. J Clin Invest 123, 5389–5400. [DOI] [PMC free article] [PubMed] [Google Scholar]
  251. Nakamura TJ, Sellix MT, Menaker M, Block GD, 2008a. Estrogen directly modulates circadian rhythms of PER2 expression in the uterus. Am J Physiol Endocrinol Metab 295, E1025–1031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  252. Nakamura TJ, Sellix MT, Menaker M, Block GD, 2008b. Estrogen directly modulates circadian rhythms of PER2 expression in the uterus. American Journal of Physiology-Endocrinology and Metabolism 295, E1025–E1031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  253. Naveed M, Li L-D, Sheng G, Du Z-W, Zhou Y-P, Nan S, Zhu M-Y, Zhang J, Zhou Q-G, 2022. Agomelatine: An Astounding Sui-generis Antidepressant? Current Molecular Pharmacology 15, 943–961. [DOI] [PubMed] [Google Scholar]
  254. Nelson RJ, Bumgarner JR, Walker WH, DeVries AC, 2021. Time-of-day as a critical biological variable. Neuroscience & Biobehavioral Reviews 127, 740–746. [DOI] [PMC free article] [PubMed] [Google Scholar]
  255. Nielsen HS, Hannibal J, Fahrenkrug J, 2002. Vasoactive intestinal polypeptide induces per1 and per2 gene expression in the rat suprachiasmatic nucleus late at night. Eur J Neurosci 15, 570–574. [DOI] [PubMed] [Google Scholar]
  256. Obrietan K, Impey S, Storm DR, 1998. Light and circadian rhythmicity regulate MAP kinase activation in the suprachiasmatic nuclei. Nat Neurosci 1, 693–700. [DOI] [PubMed] [Google Scholar]
  257. Oike H, Kobori M, Suzuki T, Ishida N, 2011. Caffeine lengthens circadian rhythms in mice. Biochem Biophys Res Commun 410, 654–658. [DOI] [PubMed] [Google Scholar]
  258. Okamoto Y, Kitada R, Miyahara M, Kochiyama T, Naruse H, Sadato N, Okazawa H, Kosaka H, 2018. Altered perspective-dependent brain activation while viewing hands and associated imitation difficulties in individuals with autism spectrum disorder. Neuroimage Clin 19, 384–395. [DOI] [PMC free article] [PubMed] [Google Scholar]
  259. Ortiz-Tudela E, Iurisci I, Beau J, Karaboue A, Moreau T, Rol MA, Madrid JA, Lévi F, Innominato PF, 2014. The circadian rest-activity rhythm, a potential safety pharmacology endpoint of cancer chemotherapy. International Journal of Cancer 134, 2717–2725. [DOI] [PubMed] [Google Scholar]
  260. Pagano RR, Lovely RH, 1972. Diurnal cycle and ACTH facilitation of shuttlebox avoidance. Physiology & Behavior 8, 721–723. [DOI] [PubMed] [Google Scholar]
  261. Palada V, Gilron I, Canlon B, Svensson CI, Kalso E, 2020. The circadian clock at the intercept of sleep and pain. Pain 161, 894–900. [DOI] [PubMed] [Google Scholar]
  262. Paolone G, Lee TM, Sarter M, 2012. Time to pay attention: attentional performance time-stamped prefrontal cholinergic activation, diurnality, and performance. Journal of neuroscience 32, 12115–12128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  263. Patke A, Murphy PJ, Onat OE, Krieger AC, Özçelik T, Campbell SS, Young MW, 2017. Mutation of the Human Circadian Clock Gene CRY1 in Familial Delayed Sleep Phase Disorder. Cell 169, 203–215.e213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  264. Paul JR, McKeown AS, Davis JA, Totsch SK, Mintz EM, Kraft TW, Cowell RM, Gamble KL, 2017. Glycogen synthase kinase 3 regulates photic signaling in the suprachiasmatic nucleus. Eur J Neurosci 45, 1102–1110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  265. Paulus EV, Mintz EM, 2013. Photic and nonphotic responses of the circadian clock in serotonin-deficient Pet-1 knockout mice. Chronobiol Int 30, 1251–1260. [DOI] [PubMed] [Google Scholar]
  266. Phan TX, Chan GC, Sindreu CB, Eckel-Mahan KL, Storm DR, 2011. The diurnal oscillation of MAP (mitogen-activated protein) kinase and adenylyl cyclase activities in the hippocampus depends on the suprachiasmatic nucleus. J Neurosci 31, 10640–10647. [DOI] [PMC free article] [PubMed] [Google Scholar]
  267. Philipsen A, Hornyak M, Riemann D, 2006. Sleep and sleep disorders in adults with attention deficit/hyperactivity disorder. Sleep medicine reviews 10, 399–405. [DOI] [PubMed] [Google Scholar]
  268. Phillips AC, Gallagher S, Carroll D, Drayson M, 2008. Preliminary evidence that morning vaccination is associated with an enhanced antibody response in men. Psychophysiology 45, 663–666. [DOI] [PubMed] [Google Scholar]
  269. Pizzio GA, Hainich EC, Ferreyra GA, Coso OA, Golombek DA, 2003. Circadian and photic regulation of ERK, JNK and p38 in the hamster SCN. Neuroreport 14, 1417–1419. [DOI] [PubMed] [Google Scholar]
  270. Postema MC, van Rooij D, Anagnostou E, Arango C, Auzias G, Behrmann M, Filho GB, Calderoni S, Calvo R, Daly E, Deruelle C, Di Martino A, Dinstein I, Duran FLS, Durston S, Ecker C, Ehrlich S, Fair D, Fedor J, Feng X, Fitzgerald J, Floris DL, Freitag CM, Gallagher L, Glahn DC, Gori I, Haar S, Hoekstra L, Jahanshad N, Jalbrzikowski M, Janssen J, King JA, Kong XZ, Lazaro L, Lerch JP, Luna B, Martinho MM, McGrath J, Medland SE, Muratori F, Murphy CM, Murphy DGM, O’Heam K, Oranje B, Parellada M, Puig O, Retico A, Rosa P, Rubia K, Shook D, Taylor MJ, Tosetti M, Wallace GL, Zhou F, Thompson PM, Fisher SE, Buitelaar JK, Francks C, 2019. Altered structural brain asymmetry in autism spectrum disorder in a study of 54 datasets. Nature Communications 10, 4958. [DOI] [PMC free article] [PubMed] [Google Scholar]
  271. Preitner N, Damiola F, Lopez-Molina L, Zakany J, Duboule D, Albrecht U, Schibler U, 2002. The orphan nuclear receptor REV-ERBalpha controls circadian transcription within the positive limb of the mammalian circadian oscillator. Cell 110, 251–260. [DOI] [PubMed] [Google Scholar]
  272. Prevot V, 2022. Glial control of neuronal function. Nature Reviews Endocrinology 18, 195–195. [DOI] [PubMed] [Google Scholar]
  273. Price KH, Dziema EL, Aten S, Loeser J, Norona FE, Hoyt K, Obrietan K, 2016. Modulation of learning and memory by the targeted deletion of the circadian clock gene Bmal 1 in forebrain circuits. Behav Brain Res 308, 222–235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  274. Price KH, Obrietan K, 2018. Modulation of learning and memory by the genetic disruption of circadian oscillator populations. Physiology & Behavior 194, 387–393. [DOI] [PMC free article] [PubMed] [Google Scholar]
  275. Prolo LM, Takahashi JS, Herzog ED, 2005. Circadian rhythm generation and entrainment in astrocytes. J Neurosci 25, 404–408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  276. Prosser HM, Bradley A, Chesham JE, Ebling FJP, Hastings MH, Maywood ES, 2007. Prokineticin receptor 2 (Prokr2) is essential for the regulation of circadian behavior by the suprachiasmatic nuclei. Proceedings of the National Academy of Sciences 104, 648–653. [DOI] [PMC free article] [PubMed] [Google Scholar]
  277. Qi Z, Wang J, Gong J, Su T, Fu S, Huang L, Wang Y, 2022. Common and specific patterns of functional and structural brain alterations in schizophrenia and bipolar disorder: a multimodal voxel-based meta-analysis. J Psychiatry Neurosci 47, E32–e47. [DOI] [PMC free article] [PubMed] [Google Scholar]
  278. Qiu Y-G, Zhu J-H, Tao Q-M, Zheng P, Chen J-Z, Hu S-J, Zhang L-R, Zheng L-R, Zhao L-L, Yao X-Y, 2005. Captopril Administered at Night Restores the Diurnal Blood Pessure Rhythm in Adequately Controlled, Nondipping Hypertensives. Cardiovascular Drugs and Therapy 19, 189–195. [DOI] [PubMed] [Google Scholar]
  279. Ralph MR, Ko CH, Antoniadis EA, Seco P, Irani F, Presta C, McDonald RJ, 2002. The significance of circadian phase for performance on a reward-based learning task in hamsters. Behavioural Brain Research 136, 179–184. [DOI] [PubMed] [Google Scholar]
  280. Ramanathan C, Stowie A, Smale L, Nunez A, 2010. PER2 rhythms in the amygdala and bed nucleus of the stria terminalis of the diurnal grass rat (Arvicanthis niloticus). Neuroscience letters 473, 220–223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  281. Rawashdeh O, Jilg A, Jedlicka P, Slawska J, Thomas L, Saade A, Schwarzacher SW, Stehle JH, 2014. PERIOD 1 coordinates hippocampal rhythms and memory processing with daytime. Hippocampus 24, 712–723. [DOI] [PubMed] [Google Scholar]
  282. Reghunandanan V, Reghunandanan R, 2006. Neurotransmitters of the suprachiasmatic nuclei. J Circadian Rhythms 4, 2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  283. Reichert CF, Deboer T, Landolt HP, 2022. Adenosine, caffeine, and sleep-wake regulation: state of the science and perspectives. J Sleep Res 31, e13597. [DOI] [PMC free article] [PubMed] [Google Scholar]
  284. Reid KJ, Chang A-M, Dubocovich ML, Turek FW, Takahashi JS, Zee PC, 2001. Familial Advanced Sleep Phase Syndrome. Archives of Neurology 58, 1089–1094. [DOI] [PubMed] [Google Scholar]
  285. Richdale AL, Schreck KA, 2009. Sleep problems in autism spectrum disorders: prevalence, nature, & possible biopsychosocial aetiologies. Sleep Med Rev 13, 403–411. [DOI] [PubMed] [Google Scholar]
  286. Riemersma-van der Lek RF, Swaab DF, Twisk J, Hol EM, Hoogendijk WJ, Van Someren EJ, 2008. Effect of bright light and melatonin on cognitive and noncognitive function in elderly residents of group care facilities: a randomized controlled trial. Jama 299, 2642–2655. [DOI] [PubMed] [Google Scholar]
  287. Roenneberg T, Kuehnle T, Pramstaller PP, Ricken J, Havel M, Guth A, Merrow M, 2004. A marker for the end of adolescence. Current biology 14, R1038–R1039. [DOI] [PubMed] [Google Scholar]
  288. Roenneberg T, Merrow M, 2005. Circadian clocks - the fall and rise of physiology. Nat Rev Mol Cell Biol 6, 965–971. [DOI] [PubMed] [Google Scholar]
  289. Rohr KE, Inda T, Evans JA, 2022. Vasopressin Resets the Central Circadian Clock in a Manner Influenced by Sex and Vasoactive Intestinal Polypeptide Signaling. Neuroendocrinology 112, 904–916. [DOI] [PMC free article] [PubMed] [Google Scholar]
  290. Rosbash Μ, 2021. Circadian Rhythms and the Transcriptional Feedback Loop (Nobel Lecture)*. Angew Chem Int Ed Engl 60, 8650–8666. [DOI] [PubMed] [Google Scholar]
  291. Royston SE, Yasui N, Kondilis AG, Lord SV, Katzenellenbogen JA, Mahoney MM, 2014. ESR1 and ESR2 Differentially Regulate Daily and Circadian Activity Rhythms in Female Mice. Endocrinology 155, 2613–2623. [DOI] [PMC free article] [PubMed] [Google Scholar]
  292. Ruan W, Yuan X, Eltzschig HK, 2021. Circadian rhythm as a therapeutic target. Nature Reviews Drug Discovery 20, 287–307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  293. Ruby NF, Fernandez F, Garrett A, Klima J, Zhang P, Sapolsky R, Heller HC, 2013. Spatial memory and long-term object recognition are impaired by circadian arrhythmia and restored by the GABAAAntagonist pentylenetetrazole. PLoS One 8, e72433. [DOI] [PMC free article] [PubMed] [Google Scholar]
  294. Ruby NF, Hwang CE, Wessells C, Fernandez F, Zhang P, Sapolsky R, Heller HC, 2008. Hippocampal-dependent learning requires a functional circadian system. Proc Natl Acad Sci U S A 105, 15593–15598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  295. Rybak YE, McNeely HE, Mackenzie BE, Jain UR, Levitan RD, 2007. Seasonality and circadian preference in adult attention-deficit/hyperactivity disorder: clinical and neuropsychological correlates. Comprehensive psychiatry 48, 562–571. [DOI] [PubMed] [Google Scholar]
  296. Saito Y, Yoshida S, Nakaya N, Hata Y, Goto Y, 1991. Comparison between morning and evening doses of simvastatin in hyperlipidemic subjects. A double-blind comparative study. Arterioscler Thromb 11, 816–826. [DOI] [PubMed] [Google Scholar]
  297. Sandman CA, Kastin AJ, Schally AV, 1971. Behavioral inhibition as modified by melanocyte-stimulating hormone (MSH) and light-dark conditions. Physiology & Behavior 6, 45–48. [DOI] [PubMed] [Google Scholar]
  298. Schenck CH, Boeve BF, Mahowald MW, 2013. Delayed emergence of a parkinsonian disorder or dementia in 81% of older men initially diagnosed with idiopathic rapid eye movement sleep behavior disorder: a 16-year update on a previously reported series. Sleep Med 14, 744–748. [DOI] [PubMed] [Google Scholar]
  299. Seney ML, Cahill K, Enwright JF, Logan RW, Huo Z, Zong W, Tseng G, McClung CA, 2019. Diurnal rhythms in gene expression in the prefrontal cortex in schizophrenia. Nature Communications 10, 3355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  300. Shanware NP, Hutchinson JA, Kim SH, Zhan L, Bowler MJ, Tibbetts RS, 2011. Casein Kinase 1-dependent Phosphorylation of Familial Advanced Sleep Phase Syndrome-associated Residues Controls PERIOD 2 Stability*. Journal of Biological Chemistry 286, 12766–12774. [DOI] [PMC free article] [PubMed] [Google Scholar]
  301. Sharma A, Mohammad A, Saini AK, Goyal R, 2021. Neuroprotective Effects of Fluoxetine on Molecular Markers of Circadian Rhythm, Cognitive Deficits, Oxidative Damage, and Biomarkers of Alzheimer’s Disease-Like Pathology Induced under Chronic Constant Light Regime in Wistar Rats. ACS Chem Neurosci 12, 2233–2246. [DOI] [PubMed] [Google Scholar]
  302. Shi G, Xing L, Wu D, Bhattacharyya BJ, Jones CR, McMahon T, Chong SYC, Chen JA, Coppola G, Geschwind D, Krystal A, Ptáček LJ, Fu YH, 2019. A Rare Mutation of β(1)-Adrenergic Receptor Affects SleepAVake Behaviors. Neuron 103, 1044–1055.e1047. [DOI] [PMC free article] [PubMed] [Google Scholar]
  303. Shi G, Yin C, Fan Z, Xing L, Mostovoy Y, Kwok PY, Ashbrook LH, Krystal AD, Ptáček LJ, Fu YH, 2021. Mutations in Metabotropic Glutamate Receptor 1 Contribute to Natural Short Sleep Trait. Curr Biol 31, 13–24.el4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  304. Shibata S, Watanabe A, Hamada T, Ono M, Watanabe S, 1994. N-methyl-D-aspartate induces phase shifts in circadian rhythm of neuronal activity of rat SCN in vitro. Am J Physiol 267, R360–364. [DOI] [PubMed] [Google Scholar]
  305. Shimizu K, Kobayashi Y, Nakatsuji E, Yamazaki M, Shimba S, Sakimura K, Fukada Y, 2016. SCOP/PHLPP1β mediates circadian regulation of long-term recognition memory. Nat Commun 7, 12926. [DOI] [PMC free article] [PubMed] [Google Scholar]
  306. Shokri-Kojori E, Wang GJ, Wiers CE, Demiral SB, Guo M, Kim SW, Lindgren E, Ramirez V, Zehra A, Freeman C, Miller G, Manza P, Srivastava T, De Santi S, Tomasi D, Benveniste H, Volkow ND, 2018. β-Amyloid accumulation in the human brain after one night of sleep deprivation. Proc Natl Acad Sci U S A 115, 4483–4488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  307. Singla R, Mishra A, Cao R, 2022. The trilateral interactions between mammalian target of rapamycin (mTOR) signaling, the circadian clock, and psychiatric disorders: an emerging model. Translational Psychiatry 12, 355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  308. Sleipness EP, Sorg BA, Jansen HT, 2005. Time of day alters long-term sensitization to cocaine in rats. Brain research 1065, 132–137. [DOI] [PubMed] [Google Scholar]
  309. Sleipness EP, Sorg BA, Jansen HT, 2007. Contribution of the suprachiasmatic nucleus to day: night variation in cocaine-seeking behavior. Physiology & behavior 91, 523–530. [DOI] [PubMed] [Google Scholar]
  310. Smarr BL, Jennings KJ, Driscoll JR, Kriegsfeld LJ, 2014. A time to remember: the role of circadian clocks in learning and memory. Behav Neurosci 128, 283–303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  311. Smith VM, Sterniczuk R, Phillips CI, Antle MC, 2008. Altered photic and non-photic phase shifts in 5-HT1A receptor knockout mice. Neuroscience 157, 513–523. [DOI] [PubMed] [Google Scholar]
  312. Smucny J, Dienel SJ, Lewis DA, Carter CS, 2022. Mechanisms underlying dorsolateral prefrontal cortex contributions to cognitive dysfunction in schizophrenia. Neuropsychopharmacology 47, 292–308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  313. St Hilaire MA, Lockley SW, 2015. Caffeine does not entrain the circadian clock but improves daytime alertness in blind patients with non-24-hour rhythms. Sleep Med 16, 800–804. [DOI] [PMC free article] [PubMed] [Google Scholar]
  314. Steiger A, Farfan J, Fisher N, Heller HC, Fernandez FX, Ruby NF, 2022. Reversible Suppression of Fear Memory Recall by Transient Circadian Arrhythmia. Front Integr Neurosci 16, 900620. [DOI] [PMC free article] [PubMed] [Google Scholar]
  315. Stephan FK, Kovacevic NS, 1978. Multiple retention deficit in passive avoidance in rats is eliminated by suprachiasmatic lesions. Behavioral Biology 22, 456–462. [DOI] [PubMed] [Google Scholar]
  316. Stephens G, McGaugh JL, Alpern HP, 1967. Periodicity and memory in mice. Psychonomic Science 8, 201–202. [Google Scholar]
  317. Takada M, Nishida K, Gondo Y, Kikuchi-Hayakawa H, Ishikawa H, Suda K, Kawai M, Hoshi R, Kuwano Y, Miyazaki K, Rokutan K, 2017. Beneficial effects of Lactobacillus casei strain Shirota on academic stress-induced sleep disturbance in healthy adults: a double-blind, randomised, placebo-controlled trial. Benef Microbes 8, 153–162. [DOI] [PubMed] [Google Scholar]
  318. Takahashi JS, 2017. Transcriptional architecture of the mammalian circadian clock. Nature Reviews Genetics 18, 164–179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  319. Takahashi Y, Sawa K, Okada T, 2013. The diurnal variation of performance of the novel location recognition task in male rats. Behav Brain Res 256, 488–493. [DOI] [PubMed] [Google Scholar]
  320. Takano A, Uchiyama M, Kajimura N, Mishima K, Inoue Y, Kamei Y, Kitajima T, Shibui K, Katoh M, Watanabe T, Hashimotodani Y, Nakajima T, Ozeki Y, Hori T, Yamada N, Toyoshima R, Ozaki N, Okawa M, Nagai K, Takahashi K, Isojima Y, Yamauchi T, Ebisawa T, 2004. A Missense Variation in Human Casein Kinase I Epsilon Gene that Induces Functional Alteration and Shows an Inverse Association with Circadian Rhythm Sleep Disorders. Neuropsychopharmacology 29, 1901–1909. [DOI] [PubMed] [Google Scholar]
  321. Takase M, Taira M, Sasaki H, 1998. Sleep-wake rhythm of autistic children. Psychiatry Clin Neurosci 52, 181–182. [DOI] [PubMed] [Google Scholar]
  322. Tam SK, Hasan S, Choi HM, Brown LA, Jagannath A, Hughes S, Hankins MW, Foster RG, Vyazovskiy VV, Bannerman DM, Peirson SN, 2017. Constant Light Desynchronizes Olfactory versus Object and Visuospatial Recognition Memory Performance. J Neurosci 37, 3555–3567. [DOI] [PMC free article] [PubMed] [Google Scholar]
  323. Tam SKE, Brown LA, Wilson TS, Tir S, Fisk AS, Pothecary CA, van der Vinne V, Foster RG, Vyazovskiy VV, Bannerman DM, Harrington ME, Peirson SN, 2021. Dim light in the evening causes coordinated realignment of circadian rhythms, sleep, and short-term memory. Proc Natl Acad Sci U S A 118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  324. Tamimi Z, Abusamak M, Al-Waeli H, Al-Tamimi M, Al Habashneh R, Ghanim M, Al-Nusair M, Gao Q, Nicolau B, Tamimi F, 2022. NSAID chronotherapy after impacted third molar extraction: a randomized controlled trial. Oral Maxillofac Surg 26, 663–672. [DOI] [PubMed] [Google Scholar]
  325. Tapp WN, Holloway FA, 1981. Phase Shifting Circadian Rhythms Produces Retrograde Amnesia. Science 211, 1056–1058. [DOI] [PubMed] [Google Scholar]
  326. Temple JL, Bernard C, Lipshultz SE, Czachor JD, Westphal JA, Mestre MA, 2017. The Safety of Ingested Caffeine: A Comprehensive Review. Frontiers in Psychiatry 8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  327. Tischkau SA, Gallman EA, Buchanan GF, Gillette MU, 2000. Differential cAMP gating of glutamatergic signaling regulates long-term state changes in the suprachiasmatic circadian clock. J Neurosci 20, 7830–7837. [DOI] [PMC free article] [PubMed] [Google Scholar]
  328. Tischkau SA, Weber ET, Abbott SM, Mitchell JW, Gillette MU, 2003. Circadian Clock-Controlled Regulation of cGMP-Protein Kinase G in the Nocturnal Domain. The Journal of Neuroscience 23, 7543–7550. [DOI] [PMC free article] [PubMed] [Google Scholar]
  329. Todd WD, Machado NL, 2019. A time to fight: Circadian control of aggression and associated autonomic support. Autonomic Neuroscience 217, 35–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
  330. Toh KL, Jones CR, He Y, Eide EJ, Hinz WA, Virshup DM, Ptáček LJ, Fu Y-H, 2001. An hPer2 Phosphorylation Site Mutation in Familial Advanced Sleep Phase Syndrome. Science 291, 1040–1043. [DOI] [PubMed] [Google Scholar]
  331. Tordjman S, Anderson GM, Pichard N, Charbuy H, Touitou Y, 2005. Nocturnal excretion of 6-sulphatoxymelatonin in children and adolescents with autistic disorder. Biol Psychiatry 57, 134–138. [DOI] [PubMed] [Google Scholar]
  332. Torres R, Fisher M, Birznieks G, Polymeropoulos C, Kay GG, Xiao C, Polymeropoulos MH, 2022. Simulated driving performance in healthy adults after night-time administration of 20 mg tasimelteon. Journal of Sleep Research 31, e 13430. [DOI] [PMC free article] [PubMed] [Google Scholar]
  333. Tsao CH, Flint J, Huang GJ, 2022. Influence of diurnal phase on behavioral tests of sensorimotor performance, anxiety, learning and memory in mice. Sci Rep 12, 432. [DOI] [PMC free article] [PubMed] [Google Scholar]
  334. Tso CF, Simon T, Greenlaw AC, Puri T, Mieda M, Herzog ED, 2017. Astrocytes Regulate Daily Rhythms in the Suprachiasmatic Nucleus and Behavior. Curr Biol 27, 1055–1061. [DOI] [PMC free article] [PubMed] [Google Scholar]
  335. Uchida K, Okamoto N, Ohara K, Morita Y, 1996. Daily rhythm of serum melatonin in patients with dementia of the degenerate type. Brain Research 717, 154–159. [DOI] [PubMed] [Google Scholar]
  336. Urban MW, Lo C, Bodinayake KK, Brunswick CA, Murakami S, Heimann AC, Kwapis JF, 2021. The circadian clock gene Per1 modulates context fear memory formation within the retrosplenial cortex in a sex-specific manner. Neurobiol Feam Mem 185, 107535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  337. Uz T, Ahmed R, Akhisaroglu M, Kurtuncu M, Imbesi M, Arslan AD, Manev H, 2005. Effect of fluoxetine and cocaine on the expression of clock genes in the mouse hippocampus and striatum. Neuroscience 134, 1309–1316. [DOI] [PubMed] [Google Scholar]
  338. Vadnie CA, McClung CA, 2017. Circadian Rhythm Disturbances in Mood Disorders: Insights into the Role of the Suprachiasmatic Nucleus. Neural Plast 2017, 1504507. [DOI] [PMC free article] [PubMed] [Google Scholar]
  339. Valentinuzzi VS, Kolker DE, Vitatema MH, Ferrari EAM, Takahashi JS, Turek FW, 2001. Effect of circadian phase on context and cued fear conditioning in C57BF/6J mice. Animal Learning & Behavior 29, 133–142. [Google Scholar]
  340. Valentinuzzi VS, Menna-Barreto L, Xavier GF, 2004a. Effect of circadian phase on performance of rats in the Morris water maze task. J Biol Rhythms 19, 312–324. [DOI] [PubMed] [Google Scholar]
  341. Valentinuzzi VS, Menna-Barreto L, Xavier GF, 2004b. Effect of circadian phase on performance of rats in the Morris water maze task. Journal of biological rhythms 19, 312–324. [DOI] [PubMed] [Google Scholar]
  342. Van der Zee EA, Havekes R, Barf RP, Hut RA, Nijholt IM, Jacobs EH, Gerkema MP, 2008. Circadian Time-Place Learning in Mice Depends on Cry Genes. Current Biology 18, 844–848. [DOI] [PubMed] [Google Scholar]
  343. van Ooijen G, O’Neill JS, 2016. Intracellular magnesium and the rhythms of life. Cell Cycle 15, 2997–2998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  344. Vanselow JT, Kramer A, 2010. Posttranslational Regulation of Circadian Clocks, in: Albrecht U (Ed.), The Circadian Clock. Springer; New York, New York, NY, pp. 79–104. [Google Scholar]
  345. Verwey M, Amir S, 2012. Variable restricted feeding disrupts the daily oscillations of Period2 expression in the limbic forebrain and dorsal striatum in rats. J Mol Neurosci 46, 258–264. [DOI] [PubMed] [Google Scholar]
  346. Videnovic A, Golombek D, 2017. Circadian dysregulation in Parkinson’s disease. Neurobiology of sleep and circadian rhythms 2, 53–58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  347. Videnovic A, Noble C, Reid KJ, Peng J, Turek FW, Marconi A, Rademaker AW, Simuni T, Zadikoff C, Zee PC, 2014. Circadian melatonin rhythm and excessive daytime sleepiness in Parkinson disease. JAMA Neurol 71, 463–469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  348. Volicer F, Harper DG, Manning BC, Goldstein R, Satlin A, 2001. Sundowning and circadian rhythms in Alzheimer’s disease. Am J Psychiatry 158, 704–711. [DOI] [PubMed] [Google Scholar]
  349. Wake H, Moorhouse AJ, Jinno S, Kohsaka S, Nabekura J, 2009. Resting microglia directly monitor the functional state of synapses in vivo and determine the fate of ischemic terminals. J Neurosci 29, 3974–3980. [DOI] [PMC free article] [PubMed] [Google Scholar]
  350. Walker WH, Walton JC, DeVries AC, Nelson RJ, 2020. Circadian rhythm disruption and mental health. Translational Psychiatry 10, 28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  351. Wallace A, Chinn D, Rubin G, 2003. Taking simvastatin in the morning compared with in the evening: randomised controlled trial. Bmj 327, 788. [DOI] [PMC free article] [PubMed] [Google Scholar]
  352. Wang LM, Dragich JM, Kudo T, Odom IH, Welsh DK, O’Dell TJ, Colwell CS, 2009a. Expression of the circadian clock gene Period2 in the hippocampus: possible implications for synaptic plasticity and learned behaviour. ASN neuro 1, AN20090020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  353. Wang LM, Dragich JM, Kudo T, Odom IH, Welsh DK, O’Dell TJ, Colwell CS, 2009b. Expression of the circadian clock gene Period2 in the hippocampus: possible implications for synaptic plasticity and learned behaviour. ASN Neuro 1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  354. Wang Y, Zhang Y, Wang W, Liu X, Chi Y, Lei J, Zhang B, Zhang T, 2020. Effects of circadian rhythm disorder on the hippocampus of SHR and WKY rats. Neurobiol Leam Mem 168, 107141. [DOI] [PubMed] [Google Scholar]
  355. Wansley RA, Holloway FA, 1975. Multiple retention deficits following one-trial appetitive training. Behavioral Biology 14, 135–149. [DOI] [PubMed] [Google Scholar]
  356. Wansley RA, Holloway FA, 1976. Oscillations in retention performance after passive avoidance training. Learning and Motivation 7, 296–302. [Google Scholar]
  357. Webb IC., 2017. Circadian rhythms and substance abuse: chronobiological considerations for the treatment of addiction. Current psychiatry reports 19, 1–8. [DOI] [PubMed] [Google Scholar]
  358. Weedon MN, Jones SE, Lane JM, Lee J, Ollila HM, Dawes A, Tyrrell J, Beaumont RN, Partonen T, Merikanto I, Rich SS, Rotter JI, Frayling TM, Rutter MK, Redline S, Sofer T, Saxena R, Wood AR, 2022. The impact of Mendelian sleep and circadian genetic variants in a population setting. PLoS Genet 18, e1010356. [DOI] [PMC free article] [PubMed] [Google Scholar]
  359. Weinert D, Schöttner K, Müller L, Wienke A, 2016. Intensive voluntary wheel running may restore circadian activity rhythms and improves the impaired cognitive performance of arrhythmic Djungarian hamsters. Chronobiol Int 33, 1161–1170. [DOI] [PubMed] [Google Scholar]
  360. Weissová K, Škrabalová J, Skálová K, Červená K, Bendová Z, Miletínová E, Kopřivová J, Šonka K, Dudysová D, Bartoš A, Bušková J, 2018. Circadian rhythms of melatonin and peripheral clock gene expression in idiopathic REM sleep behavior disorder. Sleep Med 52, 1–6. [DOI] [PubMed] [Google Scholar]
  361. Winocur G, Hasher L, 1999. Aging and time-of-day effects on cognition in rats. Behav Neurosci 113, 991–997. [DOI] [PubMed] [Google Scholar]
  362. Winocur G, Hasher L, 2004. Age and time-of-day effects on learning and memory in a non-matching-to-sample test. Neurobiol Aging 25, 1107–1115. [DOI] [PubMed] [Google Scholar]
  363. Winsky-Sommerer R, de Oliveira P, Loomis S, Wafford K, Dijk D-J, Gilmour G, 2019. Disturbances of sleep quality, timing and structure and their relationship with other neuropsychiatric symptoms in Alzheimer’s disease and schizophrenia: insights from studies in patient populations and animal models. Neuroscience & Biobehavioral Reviews 97, 112–137. [DOI] [PubMed] [Google Scholar]
  364. Woodruff ER, Chun LE, Hinds LR, Varra NM, Tirado D, Morton SJ, McClung CA, Spencer RL, 2018. Coordination between Prefrontal Cortex Clock Gene Expression and Corticosterone Contributes to Enhanced Conditioned Fear Extinction Recall. eNeuro 5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  365. Woodruff ER, Greenwood BN, Chun LE, Fardi S, Hinds LR, Spencer RL, 2015a. Adrenal-dependent diurnal modulation of conditioned fear extinction learning. Behavioural Brain Research 286, 249–255. [DOI] [PMC free article] [PubMed] [Google Scholar]
  366. Woodruff ER, Greenwood BN, Chun LE, Fardi S, Hinds LR, Spencer RL, 2015b. Adrenal-dependent diurnal modulation of conditioned fear extinction learning. Behav Brain Res 286, 249–255. [DOI] [PMC free article] [PubMed] [Google Scholar]
  367. Wu Z.-y., Huang S.-d., Zou J.-j., Wang Q.-x., Naveed M, Bao H.-n., Wang W, Fukunaga K, Han F, 2020. Autism spectrum disorder (ASD): Disturbance of the melatonin system and its implications. Biomedicine & Pharmacotherapy 130, 110496. [DOI] [PubMed] [Google Scholar]
  368. Xing L, Shi G, Mostovoy Y, Gentry NW, Fan Z, McMahon TB, Kwok PY, Jones CR, Ptáček LJ, Fu YH, 2019. Mutant neuropeptide S receptor reduces sleep duration with preserved memory consolidation. Sci Transl Med 11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  369. Xu Y, Padiath QS, Shapiro RE, Jones CR, Wu SC, Saigoh N, Saigoh K, Ptácek LJ, Fu YH, 2005. Functional consequences of a CKIdelta mutation causing familial advanced sleep phase syndrome. Nature 434, 640–644. [DOI] [PubMed] [Google Scholar]
  370. Yamaguchi Y, 2018. Arginine vasopressin signaling in the suprachiasmatic nucleus on the resilience of circadian clock to jet lag. Neuroscience Research 129, 57–61. [DOI] [PubMed] [Google Scholar]
  371. Yang G, Wang H, Zhang E, 2015. Editorial: Therapeutic implications of circadian rhythms. Frontiers in Pharmacology 6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  372. Yang JJ, Wang YT, Cheng PC, Kuo YJ, Huang RC, 2010. Cholinergic modulation of neuronal excitability in the rat suprachiasmatic nucleus. J Neurophysiol 103, 1397–1409. [DOI] [PubMed] [Google Scholar]
  373. Yao Y, Taub A.B.n., LeSauter J, Silver R, 2021. Identification of the suprachiasmatic nucleus venous portal system in the mammalian brain. Nature Communications 12, 5643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  374. Yegnanarayan R, Mahesh SD, Sangle S, 2006. Chronotherapeutic dose schedule of phenytoin and carbamazepine in epileptic patients. Chronobiol Int 23, 1035–1046. [DOI] [PubMed] [Google Scholar]
  375. Yokota S, Yamamoto M, Moriya T, Akiyama M, Fukunaga K, Miyamoto E, Shibata S, 2001. Involvement of calcium-calmodulin protein kinase but not mitogen-activated protein kinase in light-induced phase delays and Per gene expression in the suprachiasmatic nucleus of the hamster. J Neurochem 77, 618–627. [DOI] [PubMed] [Google Scholar]
  376. Yoon HS, Kim SH, Kim JK, Ko SH, Ko JE, Park SJ, Park MG, Lee JH, Hyon MS, 2011. Comparison of effects of morning versus evening administration of ezetimibe/simvastatin on serum cholesterol in patients with primary hypercholesterolemia. Ann Pharmacother 45, 841–849. [DOI] [PubMed] [Google Scholar]
  377. Yoshitane H, Honma S, Imamura K, Nakajima H, Nishide S.-y., Ono D, Kiyota H, Shinozaki N, Matsuki H, Wada N, Doi H, Hamada T, Honma K.-i., Fukada Y, 2012. JNK regulates the photic response of the mammalian circadian clock. EMBO reports 13, 455–461. [DOI] [PMC free article] [PubMed] [Google Scholar]
  378. Zeitzer JM, Dijk DJ, Kronauer R, Brown E, Czeisler C, 2000. Sensitivity of the human circadian pacemaker to nocturnal light: melatonin phase resetting and suppression. J Physiol 526 Pt 3, 695–702. [DOI] [PMC free article] [PubMed] [Google Scholar]
  379. Zhang L, Hirano A, Hsu PK, Jones CR, Sakai N, Okuro M, McMahon T, Yamazaki M, Xu Y, Saigoh N, Saigoh K, Lin ST, Kaasik K, Nishino S, Ptáček LJ, Fu YH, 2016. A PERIOD3 variant causes a circadian phenotype and is associated with a seasonal mood trait. Proc Natl Acad Sci U S A 113, E1536–1544. [DOI] [PMC free article] [PubMed] [Google Scholar]
  380. Zhang L, Yan R, Wu Z, 2020. Metagenomics analysis of intestinal flora modulatory effect of green tea polyphenols by a circadian rhythm dysfunction mouse model. J Food Biochem, e13430. [DOI] [PubMed] [Google Scholar]
  381. Zhang SL, Yue Z, Arnold DM, Artiushin G, Sehgal A, 2018. A Circadian Clock in the Blood-Brain Barrier Regulates Xenobiotic Efflux. Cell 173, 130–139.e110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  382. Zhang Y, Liao J, Li Q, Zhang X, Liu L, Yan J, Zhang D, Yan H, Yue W, 2021. Altered Resting-State Brain Activity in Schizophrenia and Obsessive-Compulsive Disorder Compared With Non-psychiatric Controls: Commonalities and Distinctions Across Disorders. Front Psychiatry 12, 681701. [DOI] [PMC free article] [PubMed] [Google Scholar]
  383. Zou H, Zhou H, Yan R, Yao Z, Lu Q, 2022. Chronotype, circadian rhythm, and psychiatric disorders: Recent evidence and potential mechanisms. Front Neurosci 16, 811771. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

The article includes all data used in this work.

RESOURCES