Skip to main content
F1000Research logoLink to F1000Research
. 2016 Aug 24;5:F1000 Faculty Rev-2062. [Version 1] doi: 10.12688/f1000research.9180.1

Circadian influences on dopamine circuits of the brain: regulation of striatal rhythms of clock gene expression and implications for psychopathology and disease

Michael Verwey 1, Sabine Dhir 2, Shimon Amir 1,a
PMCID: PMC5007753  PMID: 27635233

Abstract

Circadian clock proteins form an autoregulatory feedback loop that is central to the endogenous generation and transmission of daily rhythms in behavior and physiology. Increasingly, circadian rhythms in clock gene expression are being reported in diverse tissues and brain regions that lie outside of the suprachiasmatic nucleus (SCN), the master circadian clock in mammals. For many of these extra-SCN rhythms, however, the region-specific implications are still emerging. In order to gain important insights into the potential behavioral, physiological, and psychological relevance of these daily oscillations, researchers have begun to focus on describing the neurochemical, hormonal, metabolic, and epigenetic contributions to the regulation of these rhythms. This review will highlight important sites and sources of circadian control within dopaminergic and striatal circuitries of the brain and will discuss potential implications for psychopathology and disease . For example, rhythms in clock gene expression in the dorsal striatum are sensitive to changes in dopamine release, which has potential implications for Parkinson’s disease and drug addiction. Rhythms in the ventral striatum and limbic forebrain are sensitive to psychological and physical stressors, which may have implications for major depressive disorder. Collectively, a rich circadian tapestry has emerged that forces us to expand traditional views and to reconsider the psychopathological, behavioral, and physiological importance of these region-specific rhythms in brain areas that are not immediately linked with the regulation of circadian rhythms.

Keywords: circadian clock, circadian rhythms, clock gene, circadian control

Introduction

Circadian rhythms are observed when metabolic, psychological, and behavioral processes are modulated over the course of a day, even in the absence of environmental time cues 17. These endogenously rhythmic processes are intimately linked with the function of circadian clocks and oscillators distributed throughout the brain and body; in mammals, they are coordinated by the master circadian clock located in the suprachiasmatic nucleus (SCN) 811. The circadian autoregulatory feedback loop, which is composed of several core circadian clock genes, provides the molecular basis for both the generation and the output of these circadian clocks and oscillators 1216. Clock genes, and in particular the Period genes (e.g. Per1 and Per2), provide important circadian markers that help to highlight putative sites of circadian control as well as identify the environmental stimuli and internal signals that influence these region- and tissue-specific rhythms 1725. This omnipresent circadian system responds to the health and behavior of an individual, reacts dynamically to environmental pressures, and could interact with the progression and severity of several diseases 2628. Therefore, we must increasingly consider biological processes, lifestyle, health, and disease in the context of this ever-broadening understanding of the circadian system.

In the current review, we begin by highlighting several sites and sources of circadian regulation. In doing so, we aim to expand the “canonical” understanding of the circadian system and highlight region-specific interactions that have implications for the regulation of metabolism, stress, and epigenetics. In the first section, we provide several examples of how these daily rhythms in clock gene expression can influence daily rhythms in behavior and physiology. Moreover, in addition to their canonical roles within the circadian clock, several clock genes are also pleiotropic transcriptional regulators. Therefore, the expression of such clock genes may also interact with a range of other “non-circadian” cellular and metabolic processes. Through these examples, we hope to lend support to the potential importance of striatal rhythms, where the links with behavior and psychopathology still remain largely speculative. In the second section, we go on to describe several examples where the circadian system influences dopamine systems of the brain and explore some recent examples in which daily rhythms in dopamine release and signaling are able to feedback to influence the daily rhythms of clock gene expression in the dorsal striatum. Finally, in the third section of this review, we build on this work and discuss the potential circadian influences as they relate to psychopathology, and the broader dopamine circuitry, with an emphasis on the nucleus accumbens (NAcc). A major goal of this review is to help researchers appreciate that daily rhythms in circadian clock gene expression can be important regulators of local brain function and encourage us all to strive towards a better understanding of these circadian and pleiotropic influences of region-specific rhythms in circadian clock gene expression.

Sites and sources of circadian control

Light is one major environmental cue that is able to produce phase adjustments in the SCN, the master circadian clock in mammals. Specifically, a subset of intrinsically photosensitive retinal ganglion cells (ipRGCs) makes projections from the retina to the SCN and functions to entrain the daily rhythms of clock gene expression in this structure with environmental light–dark cycles 2932. One of the most important aspects in this pathway is the time of day when light stimulation occurs, rather than the simple absence, presence, or amount of light 4. For instance, light during the day will typically produce little or no change in the phase of the SCN. But light during the night will produce strong phase advances or phase delays within the SCN-based circadian clock, depending on when the nighttime light is delivered, and produce related shifts in downstream behavioral and physiological rhythms 33. Sometimes these light pulses can have adaptive effects that push the SCN back into synchrony with environmental cycles. But when artificial sources of light are used late at night, this can also push the SCN out of synchrony with the environment and could serve to exacerbate the same circadian disruptions that may have disrupted sleep in the first place. A growing literature has been supporting this hypothesis that light at night, acting through either the circadian clock or other light-responsive circuitry, can have negative consequences on sleep, health, mood, and disease 3436.

Daily rhythms in circadian clock gene expression have also been documented in many other tissues and nuclei of the brain and body, including the amygdala, hippocampus, oval nucleus of the bed nucleus of the stria terminalis, and olfactory bulbs 1720, 23, 3739. However, extra-SCN oscillations of clock gene expression are not typically entrained by light directly. Instead they utilize neural, hormonal, metabolic, and neurochemical rhythms to remain entrained with the environment and maintain an appropriate internal organization with each other 4045. The demands, reactivity, and function of each brain area and tissue are highly region specific. Likewise, the daily oscillations of circadian clock gene expression are also region specific, and each region exhibits a unique peak, trough, and phase in local oscillations of clock gene expression 24. As a result, we must also seek to better understand the influences on, and co-ordination between, the many extra-SCN circadian oscillators contained within the brain.

Daily oscillations in clock gene expression have been linked with specific functions in only a few cases. For instance, in the olfactory bulb, it took several years after circadian rhythms in clock gene expression were first observed 18 for a link to be made between these rhythms in clock gene expression and the control of olfactory responsivity 46, 47. Likewise, daily rhythms in circadian clock gene expression in the adrenal gland have also been shown to be an important gating mechanism in the control of daily rhythms in glucocorticoid release 14. In many other cases, however, the specific sites of circadian control remain unclear. For instance, changes in the light–dark cycle that simulate jet lag induce cognitive deficits that affect fear-conditioning 48. While this effect likely involves the amygdala and/or hippocampus, the direct link between changes in fear-conditioning and circadian clock gene expression in either of these structures remains tentative. Moreover, global mutations in individual clock genes also produce important changes in behavior, neurochemistry, and mood 49, 50, but, again, the site(s) of circadian control that are relevant to these effects are still being identified. Therefore, these links between site-specific clock gene expression and functional consequences in behavior and physiology continue to be another ongoing and major challenge for future research.

Metabolic processes are another major influence on, and consequence of, the circadian system 5156. In rodents, restricting food-availability to a single mealtime each day engages a network of food-entrainable oscillators in the brain 57. Restricted feeding is a compelling area of study because it is a rare example of an explicitly circadian phenomenon that does not rely on the SCN 58. However, the critical site(s) for circadian food-entrainment have remained elusive 5961 and the degree to which food-entrainment relies on circadian clock gene expression also remains unclear 60, 62, 63. Even so, the daily patterns of circadian clock gene expression are responsive to restricted feeding schedules, and this response is also region specific 57. For example, restricted feeding schedules that provide daytime or nighttime meals will adjust circadian oscillations in some overlapping brain areas, but some of these effects will also depend on the meal time rather than the restriction, per se 42, 43. The importance of the circadian time when food is consumed is further emphasized when rodents are fed a high-fat diet during either their active or their inactive period 64. Such simple changes in the time of food availability/consumption can exacerbate or mitigate weight gain, even if the number of calories consumed does not change. The description of these complex interactions between the metabolic and circadian systems continues to be another major interest in the field. A better understanding of this interaction will highlight a secondary pathway, which, in addition to light, could be used to suggest appropriate mealtimes that produce adaptive adjustments within the circadian system and help to “normalize” rhythm disruptions linked to jet lag, shift-work, or psychopathology.

Extra-SCN rhythms in clock gene expression integrate state variables and, depending on the tissue, can respond to a wide variety of environmental and hormonal cues. For example, the hypothalamic-pituitary-adrenal (HPA) axis provides an interesting example of interaction between the systems that regulate circadian rhythms and stress. There is a strong daily rhythm in the release of glucocorticoids from the adrenal that depends on the SCN 8. Subsequently, these rhythms in glucocorticoid release are also crucial to the entrainment and maintenance of circadian rhythms in clock gene expression in parts of the limbic forebrain and amygdala 41, 65, 66. Acute stressors, which produce a short-term peak in glucocorticoid release, also produce acute changes in the expression of certain clock genes in many of the same regions 67, 68. Therefore, in diseases like depression, where baseline and stress-reactive changes in glucocorticoid release occur, it is becoming increasingly important to also consider the potential downstream effects on the rhythms of clock gene expression in the brain.

An emerging relationship also exists between circadian rhythms and the field of epigenetics. Epigenetics broadly refers to the study of modifications to the chromatin structure of DNA, which help to regulate gene expression without altering the underlying genetic code. Chromatin is composed of DNA wrapped around an octamer of histone proteins that form the nucleosome. Post-translational modifications to the N-terminal tails of histones are involved in diverse biological processes such as transcriptional activation and inactivation, chromosome packaging, and DNA damage and repair 69, 70. The histone modifications that produce these diverse effects include acetylation, methylation, phosphorylation, and ubiquitination. In addition to acting within the circadian autoregulatory feedback loop, clock proteins such as CLOCK, PERIOD (PER), and CRYPTOCHROME (CRY) also facilitate some of these epigenetic processes 71. Specifically, CLOCK acts within the canonical circadian feedback loop, but it is also a histone acetyltransferase 72. Acetylation of the histone tail neutralizes the positive charge and renders the chromatin more “open”. This conformational state is associated with active transcription and gene expression 73 and is facilitated by histone acetyltransferases such as CLOCK. Moreover, PER and CRY provide negative feedback within the circadian loop and, as it turns out, also recruit SIN3A/SIN3B and go on to associate with histone deacetylases that remove acetyl groups from the histone tails and limit gene transcription 74, 75. The time course of these epigenetic effects remains unclear but is possibly relevant for hour-to-hour changes that are observed across the 24-hour cycle. Moreover, in conditions where clock genes are mutated or where single-nucleotide polymorphisms create lasting changes in circadian clock protein function, the epigenetic consequences could have additional life-long or even trans-generational consequences.

Epigenetic effects also lie at the interface between metabolic and circadian rhythms 71. In particular, nicotinamide adenine dinucleotide (NAD +) is involved in reduction–oxidation reactions, which are key to metabolism at the cellular level and are rhythmic over 24 hours 76. As a result of these daily rhythms, NAD + is ideally positioned to respond to restricted feeding schedules that limit food-availability to a single meal each day. Downstream of this response, NAD + is also a cofactor for SIRTUIN (SIRT) proteins, which are also histone deacetylases 7779. This epigenetic pathway is a compelling possibility for how restricted feeding schedules feedback onto and modulate diverse clock-controlled processes. These metabolic pathways may also be differentially important depending on which brain area or tissue is examined. Consistent with this idea, the shifts that are observed in the daily oscillations of clock gene expression produced by restricted feeding schedules are typically region specific 42, 43, 57. As a result, such entrainment mechanisms would appear to interact and “summate” with the other hormonal and neurochemical signals, which also influence clock gene expression in a region-specific manner outside of the SCN 45, 66. Collectively, these findings point to a distributed response, which influences several brain areas, interacts with many biological systems, and helps to highlight several brain areas that are sensitive to the effects of feeding.

Dopamine, drugs, and disease

The dopamine systems of the brain have been implicated in many aspects of behavior, from fundamental motor control and endocrine release to higher-order processing like prediction error and attention 80. Dopamine systems also interact with the expression of circadian clock genes. For instance, in the retina, dopamine synthesis and receptors are adjusted in response to light, and dopamine signaling goes on to influence the expression of clock genes 8183. There is also a growing interest in the potential interactions between dopamine systems and circadian rhythms in the brain 84, 85. Drugs of abuse such as cocaine and amphetamine act, in part, through the dopamine system to increase locomotor activity in rodents. When PER1- and PER2-mutant mice are tested, the acute effects of these drugs can be quite similar between mutants and wild-type controls, but locomotor sensitization is fundamentally changed in the mutant mice 49. Studies have also shown that drugs of abuse can also induce the expression of several circadian clock genes in the dorsal striatum 86, which demonstrates another reciprocal interaction between these two systems. While these initial findings supported the general hypothesis that certain clock-related processes can influence dopamine plasticity and function, they did not determine the sites of action or the directionality of these effects. Therefore, a major goal in this area has been to uncover the brain areas and the molecular processes underlying these interactions.

One area of overlap between these two systems is found in the demonstration that the expression of enzymes integral to the production and metabolism of dopamine is influenced by circadian clock genes. In particular, an enzyme that is involved in dopamine metabolism, monoamine oxidase A (MAOa), is clock controlled and influences mood-related behaviors 87. Likewise, the rate-limiting enzyme in the production of dopamine, tyrosine hydroxylase (TH), is also regulated by the circadian clock 50, 88, 89. Beyond the synthesis and metabolism, dopamine release also goes on to provide a robust circadian drive that makes important contributions to rhythms in behavior and physiology. For instance, melatonin release from the pineal gland is highly rhythmic and often used as a circadian marker in human studies. The release of melatonin also relies on the heteromerization of adrenergic receptors with dopamine D4-receptors, which positions dopamine as an important regulator of pineal function 90. Likewise, daily rhythms in dopamine have also been observed in other hypothalamic neuroendocrine neurons, which control the precisely timed release of reproductive hormones such as prolactin. These dopamine neuroendocrine neurons also express circadian clock genes, which provide some of the circadian regulation within this system 91, 92. So, depending on where we look in the brain, dopamine could be an important input to the daily rhythms of circadian clock gene expression, or dopamine could also function as an important daily cue that goes on to provide a diurnal/circadian signal to other brain areas or neural systems.

Dopamine cell bodies in the ventral tegmental area (VTA) and substantia nigra (SN) of the ventral midbrain hold tremendous importance for motivation and emotion. Dopamine cell bodies in the VTA make several distinct projections and provide major dopamine input to the ventral striatum, cortex, and limbic system. In contrast, dopamine cell bodies in the SN are best known for their roles in motor functions and project mainly to the dorsal striatum 93. Because of their profound behavioral relevance, there has also been a growing interest in understanding the behavioral and neurochemical implications of the circadian clock(s) related to these circuits 94. In particular, dopamine release provides a robustly rhythmic signal in both the dorsal and the ventral striatum 95. Specifically in nocturnal rats, during the night when these rodents are most active, extracellular dopamine levels in the striatum are elevated 45, 96, 97. Daily changes in dopamine production, transport, reuptake, and metabolism could also help to increase the amplitude of daily rhythms within this system 95, 98. Daily rhythms in dopamine release within this circuitry also produce circadian rhythms in several electrophysiological parameters such as local field potential, firing frequency, and coherence 99. Therefore, daily rhythms in dopamine release from VTA and SN projections to the dorsal and ventral striatum could provide an important daily drive that goes on to influence motivation, emotion, and ultimately behavior 94.

Robust daily rhythms in PER2 expression are also observed in both the dorsal and the ventral striatum 24, 37, 100102. Therefore, we and others have become interested in describing the importance of dopamine for the maintenance, entrainment, and generation of rhythms in clock gene expression in the striatum 45. To this end, we have used unilateral injections of 6-hydroxydopamine (6-OHDA) in the medial forebrain bundle to lesion dopamine projections on one side of the brain and study the effects on daily rhythms of PER2 expression in the dorsal striatum. Compared to the devastating behavioral effects of bilateral lesions, unilateral lesions allow the rodents to remain healthy. These rodents continue to exhibit relatively normal rhythmicity in many aspects of behavior and physiology 103, but dopamine denervation causes a significant decrease in the amplitude of the rhythm of PER2 expression on the lesioned side 45. After these initial observations, we went on to use systemic administration of pharmacological dopamine receptor antagonists or agonists to describe the receptor specificity of PER2 expression rhythms in the dorsal striatum. Under ad libitum feeding conditions, we found that dopamine appeared to be acting through D2-receptors to synchronize PER2 expression rhythms in the medium spiny neurons of the dorsal striatum 45. However, another study recently used D1-receptor knockout mice under restricted feeding conditions and found blunted PER2 expression rhythms in the dorsal striatum of mutant mice 104. These findings suggest that the D1 and D2 specificity may not be as clear as the pharmacology initially indicated or that these effects may also depend on feeding conditions. A D2-receptor-specific effect would be consistent with other observations in the striatum and retina 82, 83, 105, 106 and could point to a more general pathway that links dopamine signaling with effects on circadian clock gene expression. However, because robust dopamine receptor antibodies are largely unavailable, moving forward, we must rely on genetic approaches that can reliably differentiate D1- and D2-receptor-containing neurons in the striatum.

The amplitude of the daily rhythm in dopamine release from VTA and SN terminal regions may also contribute to the amplitude of rhythms in clock gene expression 45, 96, 97. We have already shown that dopamine release and clock gene expression is intimately linked in the dorsal striatum 45. The amplitude of this rhythm, however, could be a major factor driving clock gene expression within the ventral striatum as well. In line with this hypothesis, large amplitude rhythms in dopamine release are observed in the NAcc core, while smaller amplitude rhythms are observed in the shell 95. Likewise, large amplitude rhythms in PER2 expression are observed in the NAcc core, while smaller amplitude rhythms are observed in the shell 24. One could therefore speculate that the rhythmic release of dopamine may also serve as an important driver of circadian clock gene expression throughout both the dorsal and the ventral striatum.

The ability of dopamine to cause changes in clock gene expression could also have a number of important implications for disease severity and treatment. For instance, in conditions such as Parkinson’s disease, dopamine projections to the dorsal striatum degenerate. Reduced dopamine input would then likely blunt daily rhythms in clock gene expression in the striatum of these patients and could be an important consideration in the etiology of circadian disruptions linked with this disorder 84. Moreover, drug treatments that produce a surge of dopamine release, such as methamphetamine or cocaine, also produce alterations in the expression of several core circadian clock genes in both the dorsal and the ventral striatum 37, 107110. This general principle could also be applied to virtually any prescription drug treatment that is given systemically and which acts as an agonist or antagonist of dopamine signaling. For example, methylphenidate is used in the treatment of attention deficit hyperactivity disorder and produces elevated dopamine release. This surge in dopamine release could go on to influence the daily rhythms of clock gene expression for hours or days, long after the drug itself has worn off. Conversely, several antipsychotics used in the treatment of schizophrenia have historically antagonized dopamine receptors, which could also blunt striatal rhythms in clock gene expression. In summary, the broadening tapestry of circadian control could be implicated in a wide range of disorders, and these circadian considerations would seem to highlight several potential benefits and/or pitfalls of many treatment options.

Diverse and reciprocal effects of dopamine on the circadian system and the circadian system on dopamine still leave the causal relationship(s) between these two systems unclear. Novel therapeutic avenues could be raised through a better understanding of these interactions, which could go on to help patients improve nighttime sleep, improve daytime alertness, help reduce symptoms, or help confine symptoms to a time of day when they are more tolerable. Such chronotherapeutic improvements will probably end up being the result of careful timing of prescription pharmaceuticals in conjunction with the appropriate management and timing of other environmental stimuli (e.g. light and food). A better understanding of the times of day when symptoms are consistently better or worse may also be able to guide patients towards an improved general understanding of their own conditions.

Psychopathology

The general links between disrupted circadian rhythms and the symptoms of psychopathology have been discussed for some time. Indeed, major depressive disorder, bipolar disorder, and seasonal affective disorder all exhibit links with certain polymorphisms in clock genes 111113. Recently, however, specific brain areas that are relevant to the circadian control of psychopathological symptoms have also started to be uncovered. Generally, stressful stimuli can be used to induce depression-like phenotypes in rodents, and this has provided several interesting models to study depression. In one such animal model of depression, unpredictable chronic mild stress was shown to change the amplitude of rhythms of clock gene expression in the NAcc 25. Importantly, the severity of this blunting of NAcc rhythms was correlated with the severity of depression-like behaviors in these same mice, and this has provided a compelling link between circadian rhythms in the NAcc, stressors, and symptom severity within this model of depression.

Circadian rhythms in the NAcc have also been shown to be affected in another model of depression, which uses learned helplessness to induce a depressive phenotype in rodents. Remarkably, this model also produces blunted circadian rhythms in the NAcc 114, which provides converging evidence for the link between symptoms of depression and rhythms within the NAcc. However, it is not yet known whether changes in circadian rhythm are a cause or consequence of stress or depression. In addition, both groups have also reported changes in the amplitude of clock gene expression within the SCN, which were also associated with depression-related behaviors 25, 115, and it is another major challenge to dissociate SCN and extra-SCN effects. However, the ability to induce site-specific disruptions of circadian clock gene expression may eventually help to highlight and differentiate SCN from extra-SCN circadian contributions to psychopathology, and we look forward to studies that disrupt the expression of circadian clock genes selectively within the dorsal or ventral striatum.

Other brain areas could also make direct and indirect contributions to activity within these dopamine circuits. For many years, we have known that the habenula can inhibit the activity of dopamine neurons in the VTA and SN 116. The habenula also exhibits daily oscillations in neuronal firing, and certain cells either fire or suppress firing in response to environmental light 117, 118. Remarkably, a more recent paper has shown that the lateral habenula also responds to learned helplessness 119, the same model of depression used by Landgraf et al. Therefore, because the habenula receives inputs from midbrain dopamine structures 120 and provides feedback to the VTA 121, we suggest that it could also provide important diurnal influences on the VTA. Another source of circadian regulation of dopamine circuitry also comes from the medial prefrontal cortex (mPFC). When the mPFC is lesioned, rhythms of cFos immunoreactivity in the NAcc are severely blunted without affecting daily rhythms in cFos immunoreactivity in the VTA 122. Inactivation experiments have shown that the mPFC can also blunt the diurnal rhythm in amphetamine reward 123, suggesting that the mPFC can modulate the circadian properties of this dopamine circuitry. Thus, in addition to endogenous rhythmicity and clock gene expression within the NAcc, there are other circadian influences from the mPFC as well as light-responsive and stress-responsive rhythms from the lateral habenula that should also be considered. As a result, we must continue to strive to consider the behavioral implications at a network level rather than focusing too much on any single node.

Conclusions

It is a major challenge to describe interactions within the broadening context of the circadian system, which contains multiple loci of control and feedback. The effects of glucocorticoids, for instance, could represent an important interface between circadian regulation and psychopathology. The effects of neurotransmitters such as dopamine on clock gene expression provide another level of analysis, which could link diseases and disorders with circadian disruptions and symptoms. Region-specific rhythms of clock gene expression are sensitive to hormonal and neurochemical controls, but this also varies from region to region. The importance of these interactions between circadian systems and health is further emphasized by the finding that frequent changes in the light–dark cycle challenge the circadian system and have been shown to shorten lifespan 124. In order to advance our understanding of the implications for health and disease, we must first encourage all researchers to consider the circadian influences within their own given areas of expertise. It is no longer enough to simply control for the time of day; instead, we should each strive to understand how these daily fluctuations are influencing research and findings in our respective areas.

Acknowledgements

The authors would like to acknowledge Andrew Chapman, Ph.D. for his helpful editing of this manuscript.

Editorial Note on the Review Process

F1000 Faculty Reviews are commissioned from members of the prestigious F1000 Faculty and are edited as a service to readers. In order to make these reviews as comprehensive and accessible as possible, the referees provide input before publication and only the final, revised version is published. The referees who approved the final version are listed with their names and affiliations but without their reports on earlier versions (any comments will already have been addressed in the published version).

The referees who approved this article are:

  • Diego Golombek, Department of Science and Technology, National University of Quilmes, Buenos Aires, Argentina

  • Christopher Colwell, Department of Psychiatry & Biobehavioural Sciences, University of California Los Angeles, Los Angeles, CA, USA

Funding Statement

This work was supported by grants from the Natural Sciences and Engineering Research Council of Canada, the Canadian Institutes of Health Research, and the Fonds de Recherche du Québec Santé.

The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

[version 1; referees: 2 approved]

References

  • 1. Aschoff J: Exogenous and endogenous components in circadian rhythms. Cold Spring Harb Symp Quant Biol. 1960;25:11–28. 10.1101/SQB.1960.025.01.004 [DOI] [PubMed] [Google Scholar]
  • 2. Pittendrigh CS: Circadian rhythms and the circadian organization of living systems. Cold Spring Harb Symp Quant Biol. 1960;25:159–84. 10.1101/SQB.1960.025.01.015 [DOI] [PubMed] [Google Scholar]
  • 3. Pittendrigh CS, Daan S: A functional analysis of circadian pacemakers in nocturnal rodents. I. The stability and lability of spontaneous frequency. J Comp Physiol. 1976;106(3):223–52. 10.1007/BF01417856 [DOI] [Google Scholar]
  • 4. Daan S, Pittendrigh CS: A Functional analysis of circadian pacemakers in nocturnal rodents. II. The variability of phase response curves. J Comp Physiol. 1976;106(3):253–66. 10.1007/BF01417857 [DOI] [Google Scholar]
  • 5. Daan S, Pittendrigh CS: A functional analysis of circadian pacemakers in nocturnal rodents. III. Heavy water and constant light: Homeostasis of frequency? J Comp Physiol. 1976;106(3):267–90. 10.1007/BF01417858 [DOI] [Google Scholar]
  • 6. Pittendrigh CS, Daan S: A functional analysis of circadian pacemakers in nocturnal rodents. IV. Entrainment: Pacemaker as clock. J Comp Physiol. 1976;106(3):291–331. 10.1007/BF01417859 [DOI] [Google Scholar]
  • 7. Pittendrigh CS, Daan S: A functional analysis of circadian pacemakers in nocturnal rodents. V. Pacemaker structure: A clock for all seasons. J Comp Physiol. 1976;106(3):333–55. 10.1007/BF01417860 [DOI] [Google Scholar]
  • 8. Moore RY, Eichler VB: Loss of a circadian adrenal corticosterone rhythm following suprachiasmatic lesions in the rat. Brain Res. 1972;42(1):201–6. 10.1016/0006-8993(72)90054-6 [DOI] [PubMed] [Google Scholar]
  • 9. Stephan FK, Zucker I: Circadian rhythms in drinking behavior and locomotor activity of rats are eliminated by hypothalamic lesions. Proc Natl Acad Sci U S A. 1972;69(6):1583–6. 10.1073/pnas.69.6.1583 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10. Rusak B, Groos G: Suprachiasmatic stimulation phase shifts rodent circadian rhythms. Science. 1982;215(4538):1407–9. 10.1126/science.7063851 [DOI] [PubMed] [Google Scholar]
  • 11. Ralph MR, Foster RG, Davis FC, et al. : Transplanted suprachiasmatic nucleus determines circadian period. Science. 1990;247(4945):975–8. 10.1126/science.2305266 [DOI] [PubMed] [Google Scholar]
  • 12. Reppert SM, Weaver DR: Coordination of circadian timing in mammals. Nature. 2002;418(6901):935–41. 10.1038/nature00965 [DOI] [PubMed] [Google Scholar]
  • 13. Sato TK, Panda S, Miraglia LJ, et al. : A functional genomics strategy reveals Rora as a component of the mammalian circadian clock. Neuron. 2004;43(4):527–37. 10.1016/j.neuron.2004.07.018 [DOI] [PubMed] [Google Scholar]
  • 14. Oster H, Damerow S, Kiessling S, et al. : The circadian rhythm of glucocorticoids is regulated by a gating mechanism residing in the adrenal cortical clock. Cell Metab. 2006;4(2):163–73. 10.1016/j.cmet.2006.07.002 [DOI] [PubMed] [Google Scholar]
  • 15. DeBruyne JP, Weaver DR, Reppert SM: CLOCK and NPAS2 have overlapping roles in the suprachiasmatic circadian clock. Nat Neurosci. 2007;10(5):543–5. 10.1038/nn1884 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16. Kornmann B, Schaad O, Bujard H, et al. : System-driven and oscillator-dependent circadian transcription in mice with a conditionally active liver clock. PLoS Biol. 2007;5(2):e34. 10.1371/journal.pbio.0050034 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17. Yamazaki S, Numano R, Abe M, et al. : Resetting central and peripheral circadian oscillators in transgenic rats. Science. 2000;288(5466):682–5. 10.1126/science.288.5466.682 [DOI] [PubMed] [Google Scholar]
  • 18. Abe M, Herzog ED, Yamazaki S, et al. : Circadian rhythms in isolated brain regions. J Neurosci. 2002;22(1):350–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19. Shieh K: Distribution of the rhythm-related genes rPERIOD1, rPERIOD2, and rCLOCK, in the rat brain. Neuroscience. 2003;118(3):831–43. 10.1016/S0306-4522(03)00004-6 [DOI] [PubMed] [Google Scholar]
  • 20. Amir S, Lamont EW, Robinson B, et al. : 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. 2004;24(4):781–90. 10.1523/JNEUROSCI.4488-03.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21. Granados-Fuentes D, Prolo LM, Abraham U, et al. : The suprachiasmatic nucleus entrains, but does not sustain, circadian rhythmicity in the olfactory bulb. J Neurosci. 2004;24(3):615–9. 10.1523/JNEUROSCI.4002-03.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22. Yoo SH, Yamazaki S, Lowrey PL, et al. : PERIOD2::LUCIFERASE real-time reporting of circadian dynamics reveals persistent circadian oscillations in mouse peripheral tissues. Proc Natl Acad Sci U S A. 2004;101(15):5339–46. 10.1073/pnas.0308709101 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23. Lamont EW, Robinson B, Stewart J, et al. : 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. 2005;102(11):4180–4. 10.1073/pnas.0500901102 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24. Harbour VL, Weigl Y, Robinson B, et al. : Comprehensive mapping of regional expression of the clock protein PERIOD2 in rat forebrain across the 24-h day. PLoS One. 2013;8(10):e76391. 10.1371/journal.pone.0076391 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25. Logan RW, Edgar N, Gillman AG, et al. : Chronic Stress Induces Brain Region-Specific Alterations of Molecular Rhythms that Correlate with Depression-like Behavior in Mice. Biol Psychiatry. 2015;78(4):249–58. 10.1016/j.biopsych.2015.01.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26. Martino TA, Tata N, Belsham DD, et al. : Disturbed diurnal rhythm alters gene expression and exacerbates cardiovascular disease with rescue by resynchronization. Hypertension. 2007;49(5):1104–13. 10.1161/HYPERTENSIONAHA.106.083568 [DOI] [PubMed] [Google Scholar]
  • 27. Martino TA, Oudit GY, Herzenberg AM, et al. : Circadian rhythm disorganization produces profound cardiovascular and renal disease in hamsters. Am J Physiol Regul Integr Comp Physiol. 2008;294(5):R1675–83. 10.1152/ajpregu.00829.2007 [DOI] [PubMed] [Google Scholar]
  • 28. Takahashi JS, Hong HK, Ko CH, et al. : The genetics of mammalian circadian order and disorder: implications for physiology and disease. Nat Rev Genet. 2008;9(10):764–75. 10.1038/nrg2430 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29. Provencio I, Rodriguez IR, Jiang G, et al. : A novel human opsin in the inner retina. J Neurosci. 2000;20(2):600–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30. Gooley JJ, Lu J, Chou TC, et al. : Melanopsin in cells of origin of the retinohypothalamic tract. Nat Neurosci. 2001;4(12):1165. 10.1038/nn768 [DOI] [PubMed] [Google Scholar]
  • 31. Hattar S, Liao HW, Takao M, et al. : Melanopsin-containing retinal ganglion cells: architecture, projections, and intrinsic photosensitivity. Science. 2002;295(5557):1065–70. 10.1126/science.1069609 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32. Panda S, Sato TK, Castrucci AM, et al. : Melanopsin ( Opn4) requirement for normal light-induced circadian phase shifting. Science. 2002;298(5601):2213–6. 10.1126/science.1076848 [DOI] [PubMed] [Google Scholar]
  • 33. Abe H, Honma S, Namihira M, et al. : Phase-dependent induction by light of rat Clock gene expression in the suprachiasmatic nucleus. Brain Res Mol Brain Res. 1999;66(1–2):104–10. 10.1016/S0169-328X(99)00031-5 [DOI] [PubMed] [Google Scholar]
  • 34. Amaral FG, Castrucci AM, Cipolla-Neto J, et al. : Environmental control of biological rhythms: effects on development, fertility and metabolism. J Neuroendocrinol. 2014;26(9):603–12. 10.1111/jne.12144 [DOI] [PubMed] [Google Scholar]
  • 35. Boivin DB, Boudreau P: Impacts of shift work on sleep and circadian rhythms. Pathol Biol (Paris). 2014;62(5):292–301. 10.1016/j.patbio.2014.08.001 [DOI] [PubMed] [Google Scholar]
  • 36. Cho Y, Ryu SH, Lee BR, et al. : Effects of artificial light at night on human health: A literature review of observational and experimental studies applied to exposure assessment. Chronobiol Int. 2015;32(9):1294–310. 10.3109/07420528.2015.1073158 [DOI] [PubMed] [Google Scholar]
  • 37. Masubuchi S, Honma S, Abe H, et al. : Clock genes outside the suprachiasmatic nucleus involved in manifestation of locomotor activity rhythm in rats. Eur J Neurosci. 2000;12(12):4206–14. 10.1111/j.1460-9568.2000.01313.x [DOI] [PubMed] [Google Scholar]
  • 38. Asai M, Yoshinobu Y, Kaneko S, et al. : Circadian profile of Per gene mRNA expression in the suprachiasmatic nucleus, paraventricular nucleus, and pineal body of aged rats. J Neurosci Res. 2001;66(6):1133–9. 10.1002/jnr.10010 [DOI] [PubMed] [Google Scholar]
  • 39. Wang LM, Dragich JM, Kudo T, et al. : Expression of the circadian clock gene Period2 in the hippocampus: possible implications for synaptic plasticity and learned behaviour. ASN Neuro. 2009;1(3): pii: e00012. 10.1042/AN20090020 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40. Inouye ST, Kawamura H: Persistence of circadian rhythmicity in a mammalian hypothalamic "island" containing the suprachiasmatic nucleus. Proc Natl Acad Sci U S A. 1979;76(11):5962–6. 10.1073/pnas.76.11.5962 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41. Segall LA, Perrin JS, Walker CD, et al. : Glucocorticoid rhythms control the rhythm of expression of the clock protein, Period2, in oval nucleus of the bed nucleus of the stria terminalis and central nucleus of the amygdala in rats. Neuroscience. 2006;140(3):753–7. 10.1016/j.neuroscience.2006.03.037 [DOI] [PubMed] [Google Scholar]
  • 42. Verwey M, Khoja Z, Stewart J, et al. : Differential regulation of the expression of Period2 protein in the limbic forebrain and dorsomedial hypothalamus by daily limited access to highly palatable food in food-deprived and free-fed rats. Neuroscience. 2007;147(2):277–85. 10.1016/j.neuroscience.2007.04.044 [DOI] [PubMed] [Google Scholar]
  • 43. Verwey M, Khoja Z, Stewart J, et al. : Region-specific modulation of PER2 expression in the limbic forebrain and hypothalamus by nighttime restricted feeding in rats. Neurosci Lett. 2008;440(1):54–8. 10.1016/j.neulet.2008.05.043 [DOI] [PubMed] [Google Scholar]
  • 44. Buhr ED, Yoo SH, Takahashi JS: Temperature as a universal resetting cue for mammalian circadian oscillators. Science. 2010;330(6002):379–85. 10.1126/science.1195262 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45. Hood S, Cassidy P, Cossette MP, et al. : Endogenous dopamine regulates the rhythm of expression of the clock protein PER2 in the rat dorsal striatum via daily activation of D 2 dopamine receptors. J Neurosci. 2010;30(42):14046–58. 10.1523/JNEUROSCI.2128-10.2010 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46. Granados-Fuentes D, Saxena MT, Prolo LM, et al. : Olfactory bulb neurons express functional, entrainable circadian rhythms. Eur J Neurosci. 2004;19(4):898–906. 10.1111/j.0953-816X.2004.03117.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47. Granados-Fuentes D, Tseng A, Herzog ED: A circadian clock in the olfactory bulb controls olfactory responsivity. J Neurosci. 2006;26(47):12219–25. 10.1523/JNEUROSCI.3445-06.2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48. Loh DH, Navarro J, Hagopian A, et al. : Rapid changes in the light/dark cycle disrupt memory of conditioned fear in mice. PLoS One. 2010;5(9): pii: e12546. 10.1371/journal.pone.0012546 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49. Abarca C, Albrecht U, Spanagel R: Cocaine sensitization and reward are under the influence of circadian genes and rhythm. Proc Natl Acad Sci U S A. 2002;99(13):9026–30. 10.1073/pnas.142039099 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50. McClung CA, Sidiropoulou K, Vitaterna M, et al. : Regulation of dopaminergic transmission and cocaine reward by the Clock gene. Proc Natl Acad Sci U S A. 2005;102(26):9377–81. 10.1073/pnas.0503584102 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51. Rutter J, Reick M, McKnight SL: Metabolism and the control of circadian rhythms. Annu Rev Biochem. 2002;71:307–31. 10.1146/annurev.biochem.71.090501.142857 [DOI] [PubMed] [Google Scholar]
  • 52. Inoue I, Shinoda Y, Ikeda M, et al. : CLOCK/BMAL1 is involved in lipid metabolism via transactivation of the peroxisome proliferator-activated receptor (PPAR) response element. J Atheroscler Thromb. 2005;12(3):169–74. 10.5551/jat.12.169 [DOI] [PubMed] [Google Scholar]
  • 53. Satoh Y, Kawai H, Kudo N, et al. : Time-restricted feeding entrains daily rhythms of energy metabolism in mice. Am J Physiol Regul Integr Comp Physiol. 2006;290(5):R1276–83. 10.1152/ajpregu.00775.2005 [DOI] [PubMed] [Google Scholar]
  • 54. Sonoda J, Mehl IR, Chong LW, et al. : PGC-1beta controls mitochondrial metabolism to modulate circadian activity, adaptive thermogenesis, and hepatic steatosis. Proc Natl Acad Sci U S A. 2007;104(12):5223–8. 10.1073/pnas.0611623104 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55. Green CB, Takahashi JS, Bass J: The meter of metabolism. Cell. 2008;134(5):728–42. 10.1016/j.cell.2008.08.022 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56. Karatsoreos IN, Bhagat S, Bloss EB, et al. : Disruption of circadian clocks has ramifications for metabolism, brain, and behavior. Proc Natl Acad Sci U S A. 2011;108(4):1657–62. 10.1073/pnas.1018375108 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57. Verwey M, Amir S: Food-entrainable circadian oscillators in the brain. Eur J Neurosci. 2009;30(9):1650–7. 10.1111/j.1460-9568.2009.06960.x [DOI] [PubMed] [Google Scholar]
  • 58. Krieger DT, Hauser H, Krey LC: Suprachiasmatic nuclear lesions do not abolish food-shifted circadian adrenal and temperature rhythmicity. Science. 1977;197(4301):398–9. 10.1126/science.877566 [DOI] [PubMed] [Google Scholar]
  • 59. Stephan FK: The "other" circadian system: food as a Zeitgeber. J Biol Rhythms. 2002;17(4):284–92. [DOI] [PubMed] [Google Scholar]
  • 60. Fuller PM, Lu J, Saper CB: Differential rescue of light- and food-entrainable circadian rhythms. Science. 2008;320(5879):1074–7. 10.1126/science.1153277 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61. Mistlberger RE, Buijs RM, Challet E, et al. : Standards of evidence in chronobiology: critical review of a report that restoration of Bmal1 expression in the dorsomedial hypothalamus is sufficient to restore circadian food anticipatory rhythms in Bmal1-/- mice. J Circadian Rhythms. 2009;7:3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62. Feillet CA, Ripperger JA, Magnone MC, et al. : Lack of food anticipation in Per2 mutant mice. Curr Biol. 2006;16(20):2016–22. 10.1016/j.cub.2006.08.053 [DOI] [PubMed] [Google Scholar]
  • 63. Storch KF, Weitz CJ: Daily rhythms of food-anticipatory behavioral activity do not require the known circadian clock. Proc Natl Acad Sci U S A. 2009;106(16):6808–13. 10.1073/pnas.0902063106 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64. Arble DM, Bass J, Laposky AD, et al. : Circadian timing of food intake contributes to weight gain. Obesity (Silver Spring). 2009;17(11):2100–2. 10.1038/oby.2009.264 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65. Segall LA, Milet A, Tronche F, et al. : Brain glucocorticoid receptors are necessary for the rhythmic expression of the clock protein, PERIOD2, in the central extended amygdala in mice. Neurosci Lett. 2009;457(1):58–60. 10.1016/j.neulet.2009.03.083 [DOI] [PubMed] [Google Scholar]
  • 66. Segall LA, Amir S: Glucocorticoid regulation of clock gene expression in the mammalian limbic forebrain. J Mol Neurosci. 2010;42(2):168–75. 10.1007/s12031-010-9341-1 [DOI] [PubMed] [Google Scholar]
  • 67. Al-Safadi S, Al-Safadi A, Branchaud M, et al. : 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. 2014;9(10):e111166. 10.1371/journal.pone.0111166 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68. Al-Safadi S, Branchaud M, Rutherford S, et al. : Glucocorticoids and Stress-Induced Changes in the Expression of PERIOD1 in the Rat Forebrain. PLoS One. 2015;10(6):e0130085. 10.1371/journal.pone.0130085 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69. Wade PA, Pruss D, Wolffe AP: Histone acetylation: Chromatin in action. Trends Biochem Sci. 1997;22(4):128–32. 10.1016/S0968-0004(97)01016-5 [DOI] [PubMed] [Google Scholar]
  • 70. Shilatifard A: Chromatin modifications by methylation and ubiquitination: implications in the regulation of gene expression. Annu Rev Biochem. 2006;75:243–69. 10.1146/annurev.biochem.75.103004.142422 [DOI] [PubMed] [Google Scholar]
  • 71. Masri S, Orozco-Solis R, Aguilar-Arnal L, et al. : Coupling circadian rhythms of metabolism and chromatin remodelling. Diabetes Obes Metab. 2015;17(Suppl 1):17–22. 10.1111/dom.12509 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72. Doi M, Hirayama J, Sassone-Corsi P: Circadian regulator CLOCK is a histone acetyltransferase. Cell. 2006;125(3):497–508. 10.1016/j.cell.2006.03.033 [DOI] [PubMed] [Google Scholar]
  • 73. Workman JL, Kingston RE: Alteration of nucleosome structure as a mechanism of transcriptional regulation. Annu Rev Biochem. 1998;67:545–79. 10.1146/annurev.biochem.67.1.545 [DOI] [PubMed] [Google Scholar]
  • 74. Naruse Y, Oh-hashi K, Iijima N, et al. : Circadian and light-induced transcription of clock gene Per1 depends on histone acetylation and deacetylation. Mol Cell Biol. 2004;24(14):6278–87. 10.1128/MCB.24.14.6278-6287.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75. Duong HA, Robles MS, Knutti D, et al. : A molecular mechanism for circadian clock negative feedback. Science. 2011;332(6036):1436–9. 10.1126/science.1196766 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76. Ramsey KM, Yoshino J, Brace CS, et al. : Circadian clock feedback cycle through NAMPT-mediated NAD + biosynthesis. Science. 2009;324(5927):651–4. 10.1126/science.1171641 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77. Asher G, Gatfield D, Stratmann M, et al. : SIRT1 regulates circadian clock gene expression through PER2 deacetylation. Cell. 2008;134(2):317–28. 10.1016/j.cell.2008.06.050 [DOI] [PubMed] [Google Scholar]
  • 78. Belden WJ, Dunlap JC: SIRT1 is a circadian deacetylase for core clock components. Cell. 2008;134(2):212–4. 10.1016/j.cell.2008.07.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79. Masri S, Rigor P, Cervantes M, et al. : Partitioning circadian transcription by SIRT6 leads to segregated control of cellular metabolism. Cell. 2014;158(3):659–72. 10.1016/j.cell.2014.06.050 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80. Wise RA: Dopamine, learning and motivation. Nat Rev Neurosci. 2004;5(6):483–94. 10.1038/nrn1406 [DOI] [PubMed] [Google Scholar]
  • 81. Iuvone PM, Galli CL, Garrison-Gund CK, et al. : Light stimulates tyrosine hydroxylase activity and dopamine synthesis in retinal amacrine neurons. Science. 1978;202(4370):901–2. 10.1126/science.30997 [DOI] [PubMed] [Google Scholar]
  • 82. Doi M, Yujnovsky I, Hirayama J, et al. : Impaired light masking in dopamine D2 receptor-null mice. Nat Neurosci. 2006;9(6):732–4. 10.1038/nn1711 [DOI] [PubMed] [Google Scholar]
  • 83. Yujnovsky I, Hirayama J, Doi M, et al. : Signaling mediated by the dopamine D2 receptor potentiates circadian regulation by CLOCK:BMAL1. Proc Natl Acad Sci U S A. 2006;103(16):6386–91. 10.1073/pnas.0510691103 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84. Videnovic A, Golombek D: Circadian and sleep disorders in Parkinson's disease. Exp Neurol. 2013;243:45–56. 10.1016/j.expneurol.2012.08.018 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85. Mendoza J, Challet E: Circadian insights into dopamine mechanisms. Neuroscience. 2014;282:230–42. 10.1016/j.neuroscience.2014.07.081 [DOI] [PubMed] [Google Scholar]
  • 86. Iijima M, Nikaido T, Akiyama M, et al. : Methamphetamine-induced, suprachiasmatic nucleus-independent circadian rhythms of activity and mPer gene expression in the striatum of the mouse. Eur J Neurosci. 2002;16(5):921–9. 10.1046/j.1460-9568.2002.02140.x [DOI] [PubMed] [Google Scholar]
  • 87. Hampp G, Ripperger JA, Houben T, et al. : Regulation of monoamine oxidase A by circadian-clock components implies clock influence on mood. Curr Biol. 2008;18(9):678–83. 10.1016/j.cub.2008.04.012 [DOI] [PubMed] [Google Scholar]
  • 88. McGeer EG, McGeer PL: Circadian rhythm in pineal tyrosine hydroxylase. Science. 1966;153(3731):73–4. 10.1126/science.153.3731.73 [DOI] [PubMed] [Google Scholar]
  • 89. Sidor MM, Spencer SM, Dzirasa K, et al. : Daytime spikes in dopaminergic activity drive rapid mood-cycling in mice. Mol Psychiatry. 2015;20(11):1406–19. 10.1038/mp.2014.167 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90. Gonzalez S, Moreno-Delgado D, Moreno E, et al. : Circadian-related heteromerization of adrenergic and dopamine D 4 receptors modulates melatonin synthesis and release in the pineal gland. PLoS Biol. 2012;10(6):e1001347. 10.1371/journal.pbio.1001347 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91. Sellix MT, Freeman ME: Circadian rhythms of neuroendocrine dopaminergic neuronal activity in ovariectomized rats. Neuroendocrinology. 2003;77(1):59–70. 10.1159/000068334 [DOI] [PubMed] [Google Scholar]
  • 92. Sellix MT, Egli M, Poletini MO, et al. : Anatomical and functional characterization of clock gene expression in neuroendocrine dopaminergic neurons. Am J Physiol Regul Integr Comp Physiol. 2006;290(5):R1309–23. 10.1152/ajpregu.00555.2005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93. Beckstead RM, Domesick VB, Nauta WJ: Efferent connections of the substantia nigra and ventral tegmental area in the rat. Brain Res. 1979;175(2):191–217. 10.1016/0006-8993(79)91001-1 [DOI] [PubMed] [Google Scholar]
  • 94. Webb IC, Lehman MN, Coolen LM: Diurnal and circadian regulation of reward-related neurophysiology and behavior. Physiol Behav. 2015;143:58–69. 10.1016/j.physbeh.2015.02.034 [DOI] [PubMed] [Google Scholar]
  • 95. Ferris MJ, Espana RA, Locke JL, et al. : Dopamine transporters govern diurnal variation in extracellular dopamine tone. Proc Natl Acad Sci U S A. 2014;111(26):E2751–9. 10.1073/pnas.1407935111 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96. Owasoyo JO, Walker CA, Whitworth UG: Diurnal variation in the dopamine level of rat brain areas: effect of sodium phenobarbital. Life Sci. 1979;25(2):119–22. 10.1016/0024-3205(79)90382-5 [DOI] [PubMed] [Google Scholar]
  • 97. Paulson PE, Robinson TE: Relationship between circadian changes in spontaneous motor activity and dorsal versus ventral striatal dopamine neurotransmission assessed with on-line microdialysis. Behav Neurosci. 1994;108(3):624–35. 10.1037/0735-7044.108.3.624 [DOI] [PubMed] [Google Scholar]
  • 98. Sleipness EP, Sorg BA, Jansen HT: Diurnal differences in dopamine transporter and tyrosine hydroxylase levels in rat brain: dependence on the suprachiasmatic nucleus. Brain Res. 2007;1129(1):34–42. 10.1016/j.brainres.2006.10.063 [DOI] [PubMed] [Google Scholar]
  • 99. Frederick A, Bourget-Murray J, Chapman CA, et al. : Diurnal influences on electrophysiological oscillations and coupling in the dorsal striatum and cerebellar cortex of the anesthetized rat. Front Syst Neurosci. 2014;8:145. 10.3389/fnsys.2014.00145 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100. Uz T, Akhisaroglu M, Ahmed R, et al. : The pineal gland is critical for circadian Period1 expression in the striatum and for circadian cocaine sensitization in mice. Neuropsychopharmacology. 2003;28(12):2117–23. 10.1038/sj.npp.1300254 [DOI] [PubMed] [Google Scholar]
  • 101. Angeles-Castellanos M, Mendoza J, Escobar C: Restricted feeding schedules phase shift daily rhythms of c-Fos and protein Per1 immunoreactivity in corticolimbic regions in rats. Neuroscience. 2007;144(1):344–55. 10.1016/j.neuroscience.2006.08.064 [DOI] [PubMed] [Google Scholar]
  • 102. Verwey M, Amir S: Variable restricted feeding disrupts the daily oscillations of Period2 expression in the limbic forebrain and dorsal striatum in rats. J Mol Neurosci. 2012;46(2):258–64. 10.1007/s12031-011-9529-z [DOI] [PubMed] [Google Scholar]
  • 103. Baier PC, Branisa P, Koch R, et al. : Circadian distribution of motor-activity in unilaterally 6-hydroxy-dopamine lesioned rats. Exp Brain Res. 2006;169(2):283–8. 10.1007/s00221-005-0343-0 [DOI] [PubMed] [Google Scholar]
  • 104. Gallardo CM, Darvas M, Oviatt M, et al. : Dopamine receptor 1 neurons in the dorsal striatum regulate food anticipatory circadian activity rhythms in mice. eLife. 2014;3:e03781. 10.7554/eLife.03781 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105. Imbesi M, Yildiz S, Dirim Arslan A, et al. : Dopamine receptor-mediated regulation of neuronal "clock" gene expression. Neuroscience. 2009;158(2):537–44. 10.1016/j.neuroscience.2008.10.044 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106. Sahar S, Zocchi L, Kinoshita C, et al. : Regulation of BMAL1 protein stability and circadian function by GSK3beta-mediated phosphorylation. PLoS One. 2010;5(1):e8561. 10.1371/journal.pone.0008561 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107. Falcon E, Ozburn A, Mukherjee S, et al. : Differential regulation of the period genes in striatal regions following cocaine exposure. PLoS One. 2013;8(6):e66438. 10.1371/journal.pone.0066438 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108. Natsubori A, Honma K, Honma S: Differential responses of circadian Per2 expression rhythms in discrete brain areas to daily injection of methamphetamine and restricted feeding in rats. Eur J Neurosci. 2013;37(2):251–8. 10.1111/ejn.12034 [DOI] [PubMed] [Google Scholar]
  • 109. Natsubori A, Honma K, Honma S: Differential responses of circadian Per2 rhythms in cultured slices of discrete brain areas from rats showing internal desynchronisation by methamphetamine. Eur J Neurosci. 2013;38(4):2566–71. 10.1111/ejn.12265 [DOI] [PubMed] [Google Scholar]
  • 110. Natsubori A, Honma K, Honma S: Dual regulation of clock gene Per2 expression in discrete brain areas by the circadian pacemaker and methamphetamine-induced oscillator in rats. Eur J Neurosci. 2014;39(2):229–40. 10.1111/ejn.12400 [DOI] [PubMed] [Google Scholar]
  • 111. McClung CA: Circadian genes, rhythms and the biology of mood disorders. Pharmacol Ther. 2007;114(2):222–32. 10.1016/j.pharmthera.2007.02.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112. Lavebratt C, Sjoholm LK, Partonen T, et al. : PER2 variantion is associated with depression vulnerability. Am J Med Genet B Neuropsychiatr Genet. 2010;153B(2):570–81. 10.1002/ajmg.b.31021 [DOI] [PubMed] [Google Scholar]
  • 113. McCarthy MJ, Nievergelt CM, Kelsoe JR, et al. : A survey of genomic studies supports association of circadian clock genes with bipolar disorder spectrum illnesses and lithium response. PLoS One. 2012;7(2):e32091. 10.1371/journal.pone.0032091 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114. Landgraf D, Long JE, Welsh DK: Depression-like behaviour in mice is associated with disrupted circadian rhythms in nucleus accumbens and periaqueductal grey. Eur J Neurosci. 2016;43(10):1309–20. 10.1111/ejn.13085 [DOI] [PubMed] [Google Scholar]
  • 115. Landgraf D, Long JE, Proulx CD, et al. : Genetic Disruption of Circadian Rhythms in the Suprachiasmatic Nucleus Causes Helplessness, Behavioral Despair, and Anxiety-like Behavior in Mice. Biol Psychiatry. 2016; pii: S0006-3223(16)31101-5. 10.1016/j.biopsych.2016.03.1050 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116. Christoph GR, Leonzio RJ, Wilcox KS: Stimulation of the lateral habenula inhibits dopamine-containing neurons in the substantia nigra and ventral tegmental area of the rat. J Neurosci. 1986;6(3):613–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117. Zhao H, Rusak B: Circadian firing-rate rhythms and light responses of rat habenular nucleus neurons in vivo and in vitro. Neuroscience. 2005;132(2):519–28. 10.1016/j.neuroscience.2005.01.012 [DOI] [PubMed] [Google Scholar]
  • 118. Sakhi K, Belle MD, Gossan N, et al. : Daily variation in the electrophysiological activity of mouse medial habenula neurones. J Physiol. 2014;592(4):587–603. 10.1113/jphysiol.2013.263319 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119. Li B, Piriz J, Mirrione M, et al. : Synaptic potentiation onto habenula neurons in the learned helplessness model of depression. Nature. 2011;470(7335):535–9. 10.1038/nature09742 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120. Shen X, Ruan X, Zhao H: Stimulation of midbrain dopaminergic structures modifies firing rates of rat lateral habenula neurons. PLoS One. 2012;7(4): e34323. 10.1371/journal.pone.0034323 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121. Faget L, Osakada F, Duan J, et al. : Afferent Inputs to Neurotransmitter-Defined Cell Types in the Ventral Tegmental Area. Cell Rep. 2016;15(12):2796–808. 10.1016/j.celrep.2016.05.057 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122. Baltazar RM, Coolen LM, Webb IC: Diurnal rhythms in neural activation in the mesolimbic reward system: critical role of the medial prefrontal cortex. Eur J Neurosci. 2013;38(2):2319–27. 10.1111/ejn.12224 [DOI] [PubMed] [Google Scholar]
  • 123. Baltazar RM, Coolen LM, Webb IC: Medial prefrontal cortex inactivation attenuates the diurnal rhythm in amphetamine reward. Neuroscience. 2014;258:204–10. 10.1016/j.neuroscience.2013.11.013 [DOI] [PubMed] [Google Scholar]
  • 124. Davidson AJ, Sellix MT, Daniel J, et al. : Chronic jet-lag increases mortality in aged mice. Curr Biol. 2006;16(21):R914–6. 10.1016/j.cub.2006.09.058 [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from F1000Research are provided here courtesy of F1000 Research Ltd

RESOURCES