• Open Access
    Review

    Impact of circadian clock dysfunction on human health

    Saptadip Samanta *
    Sk Asif Ali

    Explor Neurosci. 2022;1:4–30 DOI: https://doi.org/10.37349/en.2022.00002

    Received: March 03, 2022 Accepted: June 27, 2022 Published: September 29, 2022

    Academic Editor: Ertugrul Kilic, Istanbul Medipol University, Turkey

    This article belongs to the special issue Circadian Rhythm and Melatonin

    Abstract

    All living organisms exhibit circadian rhythms. Humans show circadian rhythm of the different physiological functions such as sleep-wake cycle, core body temperature, feeding behavior, metabolic activity, heart rate variability, hormone secretion, and others. The hypothalamic suprachiasmatic nucleus (SCN) acts as a primary circadian pacemaker. Peripheral tissues have an endogenous circadian clock; however, SCN synchronizes the circadian activity of the peripheral clocks. The retinohypothalamic tract (RHT) from retinal ganglionic cells carries the photic signal into the SCN that regulates the rhythmic expression of the core clock genes through the feedback loop. At the output level, the SCN connects with the pineal gland and the peripheral tissues with the help of neuroendocrine mediators. Disruption of circadian clock functions is detrimental to health. Shift work, night work, chronic or acute jet lag, and light-at-night have adverse effects on circadian functions. Misalignment of circadian rhythm alters the expression of core clock genes, leading to deregulation of cellular activity and metabolic functions. Circadian rhythm dysfunction causes many pathologic conditions, including sleep disorders, cardiovascular problems, metabolic dysfunction, infertility, poor physical performance, as well as cancer. The present work has reviewed the relationship between circadian clock dysfunction and impaired physiological activities.

    Keywords

    Circadian rhythm, hypothalamic suprachiasmatic nucleus, clock genes, circadian dysfunction metabolic disorder, cancer

    Introduction

    Time is an everyday phenomenon of the earth’s rotation on its axis. It can be measured but not seen or touched and weighted. Rotation of the Earth governs the natural period along with the daily light-dark (LD) cycle and other environmental phenomena like ambient temperature, tide, etc. Many biochemical and physiological functions of the living organisms rhythmically oscillate under constant environmental conditions for a period of close to 24 h. These rhythmic phenomena are called circadian rhythms [1, 2]. The circadian rhythm is present in all living organisms, from bacteria to humans. The term circadian was coined by Halberg [3]. It originated from the Latin words circa means “around” and dies means “day”. Human shows several rhythmic phenomena like a sleep-wake cycle, core body temperature, heart rate, blood pressure, feeding pattern, metabolism, hormone secretion, immune responses, and neuronal functions [49].

    The suprachiasmatic nucleus (SCN) of the anterior hypothalamus performs the functions of the central clock and regulates the oscillations of physiological rhythms. Non-visual retinal ganglionic cells transmit photic signals to the SCN through the RHT after exposure to light detected by the retina [10, 11]. The photic signal controls the functions of SCN by regulating the expression of clock genes. More specifically, photic signals lead to transient up-regulation of Period 1 (Per1) and Per2 expression in the SCN. SCN also synchronizes the circadian oscillations of peripheral tissue clocks through neuro-endocrinal factors [12, 13].

    Disruption of the circadian pattern is a common incident in our modern society. Chronic and acute jet lag, shift works, night works, and extended shifts are the causes of circadian dysfunction that potentially enhances sleep disorders, poor physical performance, metabolic syndrome, cardiovascular and inflammatory diseases, neuropsychiatric illness, and cancer [1416]. The present review has focused on the impact of circadian rhythm dysfunction on different physiological functions.

    An overview of the circadian system

    Suprachiasmatic nucleus

    SCN acts as a master clock in the mammalian brain [17, 18]. The SCN is a paired structure, bilaterally situated on either side of the third ventricle just above the optic chiasm [10, 18]. The neurons of the SCN are the smallest in size (~10 μm) and are closely packed in the core and shell region. The nuclei of SCN get photic signals through the RHT after exposure to light on the retina [10, 11]. Melanopsin (OPN4) containing non-visual retinal ganglion cells (RGCs) are most sensitive to blue light (464–484 nm wavelengths) and transmit the impulse to the SCN via RHT [19, 20]. RHT starts the signal transduction process in the SCN and ultimately regulates the expression of several clock genes. Additionally, peripheral tissues also express clock genes. RHT is a glutamatergic pathway. Glutamate and pituitary adenylate cyclase-activating polypeptide (PACAP) act as the primary neurotransmitters. Glutamate acts through α-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid (AMPA) and N-methyl-D-aspartate (NMDA) receptors. Activation of glutamatergic receptors increases intracellular calcium that activates different kinases such as calcium/calmodulin-dependent protein kinase (CaMK), mitogen-activated protein kinase (MAPK), and protein kinase A (PKA). The subsequent result is phosphorylation of cyclic adenosine monophosphate (cAMP) response element binding protein (CREB). Phosphorylated CREB (p-CREB) then translocates to the nucleus and binds with calcium/cAMP response elements (CREs) in association with co-activators for the regulation of gene expression [18, 2123]. Tischkau et al. [23] reported that the 5’-flanking region of Per1 and Per2 genes carried CRE that interacts with CREB for the expression of Per genes. The mutational study indicated the role of CRE in the expression of Per genes [17]. On the other hand, output signals from SCN reach the peripheral organs for the regulation of physiological and behavioral functions [24].

    SCN makes a complex connection with the peripheral tissue clocks through the neural and hormonal mediators for the synchronization of circadian oscillations [12, 13]. Despite SCN-mediated regulation, all cells of the body have an endogenous clock system that can function autonomously to maintain circadian oscillations. At the molecular level, the activity of the circadian clock is driven by a complex network of transcription-translation feedback loop that regulates the rhythmic expression of clock-related genes. Transcription factors circadian locomotor output cycles kaput (CLOCK) and brain and muscle aryl hydrocarbon receptor nuclear translocator-like 1 (BMAL1) induce the expression of core clock components, Per and cryptochrome (Cry). Subsequently, they regulate the activity of CLOCK and BMAL1 [25]. The output signals from the SCN neurons are associated with the expression of clock-controlled genes (CCGs). These genes contain promoter sequence (CACGTG) for binding of CLOCK-BMAL1 complex (E-box) in their upstream for the operation of the transcriptional cascade. The promoter region of arginine vasopressin (Avp), D-element binding protein (Dbp), and prokineticin 2 (Pk2) contain E-box. AVP acts as a neurotransmitter, DBP is a transcription factor, and PK2 is a regulator of different biological functions (sleep cycle, behavioral rhythm, and others) [26]. The endogenous clock system is synchronized with several environmental time cues (zeitgebers) like temperature, LD cycle, physical activity, and feeding behavior [2729]. The external zeitgebers regulate the expression of the molecular components of the endogenous clock [26].

    Events of expression of the clock genes

    Approximately 12–14 core clock genes are most important for the regulation of human circadian rhythm [6, 30, 31]. The complex transcriptional/translational feedback loops (TTFLs) control the overall expression of core clock genes (Figure 1) [32, 33]. The common core clock genes are 3 Per genes (Per13), 2 Cry genes (Cry12), Bmal1, Clock, 3 retinoid-related orphan receptor (Ror, Ror α, β, and γ) genes, 2 genes of reverse-erythroblastosis (Rev-Erb; Rev-Erb α and β). A specific enzyme casein kinase 1 (CK1; CK1δ and ε) is a part of the mammalian circadian system. Furthermore, some other genes [neuronal period-aryl hydrocarbon receptor nuclear translocator single-minded protein 2 (Npas2/Mop4), Aryl hydrocarbon receptor nuclear translocator-like protein 2 (Arntl2/Mop9), and F-box/LRR repeat protein 3 (Fbxl3)] are also associated with clock functions.

    Regulation of expression of core clock genes in the SCN through the positive and negative feedback loop in the mammalian circadian system. P: PER; C: CRY; (-): repression; (+): induction; CL: CLOCK; BM: BMAL1; AC: acetyl group; Pi: phosphate; RORE: Ror response element; FBXL3: F-box protein, E3 ligase; Ccg: clock-controlled genes

    Note. Adapted from “Melatonin: an endogenous miraculous indolamine, fights against cancer progression,” by Samanta S. J Cancer Res Clin Oncol. 2020;146:1893–922 (https://doi.org/10.1007/s00432-020-03292-w). © Springer-Verlag GmbH Germany, part of Springer Nature 2020.

    The expression of Per, Cry, Ror α, and Rev-Erb α is mediated by the CLOCK-BMAL1 complex (Figure 1). CLOCK-BMAL1 heterodimer initially influences the rhythmic expression of the Per and Cry genes after binding at the E-box. Translation of PER and CRY in the cytoplasm promotes their accumulation and forms a heterodimer. Heterodimeric PER and CRY move into the nucleus and inhibit the transactivating effects of the CLOCK-BMAL1 complex, leading to the termination of transcription of Per, Cry, Ror, and Rev-Erb genes (Figure 1). This inhibition completes the negative feedback effects (Figure 1) [34, 35]. CRY and PER concentrations decrease at night, resulting in the discontinuation of their repressive activity. This event starts a new cycle of CLOCK-BMAL1 complex-mediated transcriptional activation [6, 36]. Cytoplasmic CKs [CK1δ and CK1ε] and F-box protein (FbxL3) are involved in the degradation of PER and CRY. CK1δ and CK1ε cleave the PER proteins after phosphorylation and ubiquitination [37, 38]. The ubiquitin E3 ligase (F-box protein), called FbxL3 degrades CRYs [39].

    The positive feedback loop starts when CLOCK and BMAL1 induce the expression of other clock genes. CLOCK-BMAL1 complex recognizes E-box, which bears a specific nucleotide sequence CACGTG at the upstream of the core clock genes [40, 41]. This interaction exerts inductive effects on the expression of core clock genes. The orphan nuclear receptors Ror α and Rev-Erb α regulate the positive feedback loop. They translocate to the nucleus from the cytoplasm and recognize the RORE that contains a specific sequence (AAAGTAGGTCA) upstream of the Bmal1 promoter. Ror α induces the expression of Bmal1, while Rev-Erb α represses Bmal1 transcription. This antagonistic effect controls the rhythmic expression of Bmal1 [28, 34, 41]. The mutation of the Clock gene significantly downregulates the expression of the Per1–3 and Cry1–2 levels in SCN [42, 43]. In addition to CLOCK, NPAS2 (MOP4) can form a heterodimer with BMAL1 in the SCN to regulate the transcriptional activity. The expression of NPAS2 commonly occurs in peripheral vascularized tissues. The peripheral cells remain rhythmic in the absence of CLOCK; however, Npas2 knockout cells exhibit arrhythmic behavior [44].

    Peripheral clocks

    The circadian oscillators also exist in individual organs and cells. The SCN is the mediator of synchronization for the other oscillators, including different parts of the brain and peripheral organs. This synchronization shows a period that is closer to 24 h. However, there is a phase delay relative to SCN in peripheral tissues. Large phase delay between SCN and peripheral tissues has been observed in nocturnal animals compared to diurnal species. Mure et al. [45] compared the expression of core clock components in SCN and different tissues of baboons and mice. They observed that there was ~12 h phase delay in the peak expression of clock components in the peripheral tissues of mice and baboons. The peak expression of Bmal1 and Per1 in the baboons occurred in the evening and morning, respectively, whereas, mice showed a reverse expression pattern (peak phase in morning and evening, respectively). However, Cry1 showed ~7 h phase delay compared to Bmal1 and Per1. The phase delay of other clock components in baboons and mice showed a similar pattern of Bmal1 and Per1. On the other hand, the expression pattern of clock components in SCN was not strictly in the opposite fashion. The peak phase of Bmal1 occurred at the same time (evening) in mice and baboons. However, phase delay was observed in the expression of Per1 and Cry2. Thus, there was a large phase delay in the peak expression of clock components in peripheral tissues between nocturnal and diurnal species.

    The process of synchronization between SCN and peripheral tissues is driven by the autonomic nervous system and hormones. However, circadian modulation of body temperature and feeding behavior have a less direct effect [18]. Generally, Per genes express in the brain and other peripheral tissues such as the liver and skeletal muscle [46]. The oscillation of each Per gene in the peripheral tissue of mammals synchronizes with SCN. The phase relationship between peripheral organs and SCN is linked with LD cycles [47]. Yamazaki et al. [48] reported that the absence of SCN in transgenic rats showed gradual resetting of rhythms in different tissues. A study on genetically ablated SCN indicated that the LD cycle controls circadian rhythmicity in peripheral tissues [49]. Thus, peripheral clocks can independently oscillate with the LD cycle without the influence of SCN. The SCN and endogenous zeitgeber (melatonin) regulate the autonomic nervous system, neuroendocrine secretion, body temperature, and sleep-wake cycle. However, food acts as a potent zeitgeber and effectively affects peripheral oscillation [50]. The activity of the liver synchronizes with signals derived from feeding [5153]. The metabolites of carbohydrate, lipid, amino acids, nucleotides, and xenobiotics metabolism control the oscillation of the peripheral clock [50]. The metabolites act as the cellular transducer in a peripheral organ like the liver [50, 54]. The rhythmic expressions of different genes in hepatic tissue related to metabolic functions are driven by rhythmic food intake in mice [55, 56]. The transcriptional activity of many metabolic genes lost its rhythmicity during arrhythmic feeding, although core clock genes show rhythmic oscillation in the same condition. Arrhythmic feeding impairs the expression of several genes of metabolic enzymes related to carbohydrate and lipid metabolism (glycogen synthesis, cholesterol biosynthesis, and fatty acid chain elongation), as well as the expression of glucose transporter-2 (GLUT2) and GLUT4 in hepatic and adipose tissue respectively. The GLUT2 increases the intracellular glucose level in hepatocytes that is used for lipogenesis instead of glycogen synthesis [56]. They also reported that signaling pathways like mechanistic target of rapamycin (mTOR) and extracellular signal-regulated kinase 1/2 (ERK1/2) regulate the activity of metabolic transcription factors. The phasic activation of these proteins does not occur in arrhythmic feeding conditions, while the level of these proteins remains constant in hepatocytes during different feeding paradigms. Thus, mTOR and ERK1 and ERK2 mediated metabolic gene expression does not sustain in an arrhythmic feeding state. Moreover, it had been reported that CREB, forkhead box protein O1 (FoxO1), sterol regulatory element-binding protein (SREBP), and activating transcription factor 6 (ATF6) also regulate the transcription of several genes during the feeding rhythmicity. The expressions of CREB and FoxO1 mediated genes were repressed after feeding, while the reverse effect was observed in ATF6 and SREBP regulated genes [55]. Non-SCN tissues bear the local oscillatory pacemaker and the importance of the peripheral clock was studied by Kornmann et al. [57]. They observed that genetic manipulation in circadian dysfunctioning liver expresses genes related to the circadian rhythm, which are partly regulated by the liver clock and the signals from SCN.

    Circadian rhythm disruption and health disorders

    The disruption of circadian clock function was associated with several diseases; these include sleep disorders, metabolic dysfunction, infertility, cancer, and psychological disorders. Modern lifestyle, excessive workload, depression, shift work, exposure to excessive artificial light, and short night periods severely affect the circadian clock functions and promote the initiation of circadian mediated health disorders. According to the World Health Organization, shift workers are very prone to health risks [58].

    Sleep disorder

    Human circadian rhythm disorders are related to the activity of the internal clocks and the external LD cycle. The American Academy of Sleep Medicine has classified circadian rhythm sleep disorders (CRSDs) into several categories. These include advanced sleep phase disorder/syndrome (ASPD/ASPS), delayed sleep phase disorder/syndrome (DSPD/DSPS), free-running disorder (FRD), and irregular sleep-wake rhythm (ISWR) [26]. The features of these disorders have given in Table 1. Sleep disorders are also common phenomena in jet lag and shift work, but they are treated as extrinsic disorders and do not relate to circadian clock dysfunction. The jet lag and shift work-mediated CRSDs occur due to voluntary or imposed shifts in the sleeping time. Several non-communicable disorders like stroke, mania, depression, intracranial infection, head injury, and central nervous system (CNS) stimulant or depressant drugs are associated with CRSDs [26].

    Four common sleep disorders and their features [26]

    NameFeaturesPrevalence
    ASPD/ASPS, ~2%The sleep episode occurs in advance time than the desired clock time, leading to early sleep onset and early awakening than the desired timeASPD is more prevalent in elderly people
    This syndrome is rare and has a genetic basis. It is conferred by a single autosomal dominant gene that resides on the short arm of chromosome 2 which carries Per2 genes. Mutation results from serine-to-glycine in PER2 protein are the responsible party for this syndrome
    Mutation in CK1ε lowers the rate of degradation of PER2 protein. Misregulation of PER2 protein promotes ASPS
    DSPD/DSPS, 83%The onset of sleep and final awakening are delayed concerning the desired clock time. The body temperature, melatonin, and sleep rhythm of DSPS subjects are very similar to normal subjects, but the phase is delayed. They are less productive in their profession and social lifeDSPS may appear in early childhood but commonly occurs during adolescence and in middle age people. Social and occupational impairment can promote DSPS
    FRD, ~2%It is known as a non-24-h-sleep-wake syndrome and appears as a chronic disorder. In this pattern, progressive changes of 1-h to 2-h delays in sleep onset and wake times are the fundamental featuresFRD is common in those who are blind and it may be due to failure of the entrainment to the LD cycle
    The variable period length indicates the interaction between the environment and the sleep-wake cycle
    ISWR, 12%The pattern is temporally disorganized. The variable episodes of sleeping and waking behavior are very common. A short sleep period (a few hours) is distributed randomly throughout the day and nightISWR is very common in night workers, shift workers, students of colleges, universities, technical institutes, medical schools, law schools, and research institutes
    It has also been observed mainly in the mentally retarded and demented individuals
    Display full size

    Neurodegenerative diseases

    Neurodegenerative diseases are associated with progressive loss of vulnerable neurons in the brain, leading to decreased brain functions. Progressive loss of memory and cognitive impairments, and difficulty in moving and speaking are the common symptoms of neurodegenerative diseases [59, 60]. Neurodegenerative diseases such as Alzheimer’s disease (AD), Parkinson’s disease (PD), Huntington’s disease (HD), Pick’s disease, and progressive supranuclear palsy (PSP) are bidirectionally associated with circadian disruption [6167]. Accumulation of insoluble misfolded or mutant protein aggregates like α-synuclein in PD, β-amyloid plaque and tau in AD, and huntingtin in HD, occur in these diseases [68]. Proteotoxic stress, impaired activity of ubiquitin–proteasomal and autophagosomal/lysosomal systems, mitochondrial dysfunction, glutamate toxicity, neuroinflammation, aging, and circadian dysfunction accelerate neurodegenerative disorders [63, 66, 6971].

    The circadian clock regulates the rhythmic expression of numerous genes and is also involved in protein processing and transmission of protein aggregates [68, 72, 73]. Circadian dysfunction promotes the formation of misfolded-protein aggregates and their transport. A proper sleep pattern can help to remove the misfolded protein. However, circadian deregulation also causes sleep disorders that accelerate the pathology of neurodegenerative diseases. At the experimental level, a study on postmortem tissue of patients with neurodegenerative disorders shows disruption in rhythmic expression of circadian genes [74]. This finding is also similar to the aging brain [75].

    Disruption of circadian clock functions enhances oxidative stress (OS) that induces neurodegeneration [76]. In the brain, BMAL1 regulates the transcription of redox-related genes like nicotinamide adenine dinucleotide phosphate (NADPH) dependent quinone dehydrogenase (QR2), and aldehyde dehydrogenase 2 (ALDH2). The expression of these genes essentially regulates OS and neuronal damage [77, 78]. The clock components also interact with silent information regulator transcript 1 [SIRT1, nicotinamide adenine dinucleotide (NAD)-dependent deacetylase sirtuin 1] for maintaining the cellular redox homeostasis [79] and the transcription of both CLOCK and BMAL1 in SCN [80]. Dysfunction of circadian rhythm and antioxidant activity influences PD and AD [66, 81].

    Circadian dysfunction also decreases melatonin secretion. Melatonin is considered the hormone of darkness and shows its peak at midnight. This indolamine is involved in several regulatory functions and acts as a potent endogenous antioxidant. Melatonin and its derivatives reduce OS and protect the cells against OS-induced damage, maintain the proliferation of normal cells and prevent apoptosis-mediated cell degeneration [82, 83]. Shift work and night work hamper melatonin secretion and decrease its protective role against stress-induced various diseases, including neurodegenerative disorders. Moreover, circadian dysfunction alters the pattern of rhythmic secretion of different hormones from the pituitary. The output signal from the SCN regulates the activity of the hypothalamic-pituitary axis and secretion of tropic hormones. Misalignment of circadian rhythm increases the secretion of adrenocorticotropic hormone (ACTH), followed by the release of cortisol. Excess cortisol disrupts the sleep-wake cycle that promotes neurodegeneration [84].

    Formation of neurofibrillary tangles and deposition of extracellular amyloid beta (Aβ) have been observed in AD patients. Missense mutations, single nucleotide polymorphisms (SNPs), and deregulated alternative splicing promote malformed tau protein synthesis [85, 86], which is very unstable and fails to bind with microtubule [87]. This protein accumulates as phosphorylated tau that contributes to neurofibrillary tangles formation. β-secretase and γ-secretase cleave the amyloid precursor protein to generate Aβ which forms β-amyloid plaques [88, 89]. Ma et al. [90] reported that circadian dysfunction increased Aβ synthesis by increasing β-secretase activity. Wang et al. [91] reported that Aβ altered Per1 and Per2 expression in SCN, which are associated with AD. Experiments on drosophila indicated that sufficient expression of PER exerted neuroprotection against oxidative damage [92]. PER also maintains the sleep-wake cycle and working memory [93]. Ablation of Per decreases antioxidant activity and increases OS that collectively influences neurodegeneration [92, 94]. Altered light exposure-mediated circadian desynchrony promotes AD in mice due to alteration in the expression of SIRT1 and activation of OS pathways in the hippocampus [95]. Aβ promotes the degradation of BMAL1, followed by circadian deregulation and advancement of AD [96]. Kress et al. [97] reported that Bmal1 knockout increases Aβ plaque formation. Deregulated expression of BMAL1 also disrupts PER and CRY expression and redox homeostasis that contributes to AD [98]. Li et al. [99] reported that CK1δ promotes phosphorylated tau deposition and neurofibrillary tangle formation. Administration of CK1 inhibitor protects the prefrontal cortex and hippocampus from Aβ plaque formation [100]. Thus, CK is also associated with AD development.

    Circadian dysfunction and neuroinflammation are the common risk factor for PD. PD patients show deposition of α-synuclein as Lewy bodies. It has been suggested that α-synuclein is involved in the transport of synaptic vesicles [101]. However, aggregation of misfolded α-synuclein causes PD. A study on C57BL/6 mice in abnormal LD cycles (20:4 LD) exhibits neuroinflammation and poor motor skills. Kudo et al. [102] reported that circadian dysfunction overexpresses α-synuclein in a mouse model of PD. SNPs study in PD patients indicated a greater number of SNPs in Bmal1 and Per1 gene [103]. Although circadian dysfunction and neurodegenerative diseases are bidirectionally associated, the mutation in the Parkin gene (PRKN) increases the expression of CLOCK, CRY1, and CRY2 and decreases PER2 output. Additionally, mutant PRKN decreases mitochondrial oxygen consumption and energy production [104]. Breen et al. [105] observed the arrhythmic expression of BMAL1, PER2, and Rev-Erb α occurs in the sample taken from PD patients. Liu et al. [106] reported that deregulation of BMAL1 activity occurred in 1-methyl-4-phenyl-1, 2,3,6-tetrahydropyridine (MPTP)-treated mice, resulting in the reduction of dopaminergic neurons, decreased activity of tyrosine hydroxylase, and low levels of dopamine. Thus, improper BMAL1 activity may potentiate the pathogenesis of PD. The formation of mutant huntingtin (mHTT) protein initiates the pathologies of HD. Deregulated PER1 and PER2 expression had been observed in mice with HD [107, 108]. Smarr et al. [109] observed that SCN neurons decreased daytime firing in HD models. PSP patients exhibit poor sleep quality, sleep apnea, restless leg syndrome, and difficulty in falling asleep and staying asleep [110]. Circadian dysfunction influences PSP. CK1δ is also involved in neurofibrillary tangle formation in PSP and Pick disease [111].

    Psychiatric disorders

    Circadian disruption is associated with shift work, night work, jet lag, and exposure to outdoor light at night. A significant night-time light pollution occurs due to artificial light, which is a serious problem in different metro cities throughout the world. Light-at-night desynchronizes the circadian rhythm that promotes psycho-behavioral disorders along with several negative consequences for health. Seasonal affective disorder (SAD) may occur due to alteration of the LD cycle in the natural system during seasonal variation. Short day lengths in winter and prolonged day-light during long summer days have been observed in boreal and austral countries. People of this region can suffer from dysthymia during the short day lengths and euthymia during the long summer days.

    Circadian disruption hampers the regular sleep-wake cycle and also suppresses melatonin secretion and night-time melatonin peak. The nocturnal melatonin peak maintains different physiological activities, including sleep cycle, hormone secretion, neurological activity, and psychological behavior. Impaired melatonin secretion advances different psychological disorders such as SAD, bipolar disorder (BD), unipolar depression, schizophrenia, autism spectrum disorder (ASD), and attention deficit hyperactivity disorder (ADHD) [112114]. Treatment with melatonin in pharmacological doses improves psychiatric problems [115, 116]. Circadian rhythm dysfunction deregulates the sleep-wake cycle that influences psychological disorders. ASD is very common in children and is associated with a desynchronized sleep-wake cycle. Lai et al. [117] also reported that sleep disorders are a common problem in patients with autism. Daytime sleepiness in children increases the risk of ADHD [118].

    Several authors discussed the relationship between circadian rhythm dysfunction and psychiatric disorders, specifically depression and mood problems [119123]. Jet lag is a common problem for international flight crews. Sack [124] suggested that mood changes, particularly dysphoric mood, are associated with jet lag. Experimental data indicated that an eastward journey over a month promoted sleep deprivation and increased the risk of anxiety and depression [125]. Major depressive disorder (MDD) is a disease of alteration in mood and is associated with increased sadness and irritability. Alteration in sleep, sexual desire, or appetite, inability to experience pleasure, slowing of speech or actions, crying, and suicidal thoughts are the symptomatic changes in the patients [126]. Several reports suggested that shift work and night work-dependent circadian rhythm dysfunction increases the outcome of MDD. Jet lag or social jet lag also influences MDD [122]. Decreased latency to rapid eye movement (REM) sleep, duration of slow-wave sleep and increased duration of REM sleep, disruption of melatonin and cortisol rhythms, and elevated nocturnal body temperature are associated with MDD [127]. Studies with postmortem brains of patients with MDD revealed that the core clock genes had expressed in reduced amplitude, shifted peaks, and phases [128]. Additionally, bilaterally ablation of SCN decreases immobility during forced swim tests on rats [129]. Applications of bright light therapy, wake therapy, social rhythm therapy, and antidepressants can improve MDD [122]. SCN-specific Bmal1-knockdown in mice and exposure to T22 LD cycle-induced desynchrony of the SCN in rats accelerates depressive-like behavior [130, 131].

    Circadian rhythm disruption can increase the risk of anxiety. A study reported that non-shift workers had started to suffer from anxiety if they were facing rotating shift work schedules [132]. Similarly, nurses also showed higher anxiety scores on the Hospital Anxiety and Depression Scale [133]. Improper expression of core clock components influences anxiety-like disorders. Delta19 (Δ19) mutations in the Clock gene of mice decreased the symptoms of anxiety [134]. Per1 and Per2 deficient mice showed greater anxiety-related problems; however, lack of individual expression of either Per1 or Per2 had no effects on anxiety [135].

    BD is characterized by cyclical changes in mood between mania and depression separated by an interval of normal affect. The episodes of mood in BD depend on the behavior of the person, sleep cycle, physical activity, and energy expenditure. A meta-analysis study has established the relationship between circadian rhythm disruption and BD [136]. Jet lag can influence the BD. Traveling in different time zones, particularly east to west occurs depression, whereas traveling from west to east develops mania [137]. Le-Niculescu et al. [138] reported that core clock genes were associated with BD. Polymorphisms in circadian clock-associated genes increase the risk of development of BD [139]. Lithium slows down the fast-running circadian clock and normalizes the circadian rhythm, which is an effective treatment for BD. Moreover, bright light therapy can improve BD [122, 139].

    Schizophrenia is a typical mental disorder. The symptoms are delusion, hallucination, thought disorders, movement disorders, anhedonia, alogia, cognitive failure, and attention deficits. There are several reasons for schizophrenia. These include genetic factors (familiar), environmental (poverty, stress, exposure to viruses, and nutritional problems before birth), imbalances in neurotransmitters, and abnormalities in brain functions. Moreover, circadian rhythm disruption and sleep disorders are also associated with schizophrenia [122, 140, 141]. Studies on postmortem brain tissues of schizophrenia revealed the arrhythmic expression of many genes in the prefrontal cortex and rhythmic expression of mitochondrial genes, compared to healthy individuals [142]. Schizophrenia individuals show hyperactivity of the hypothalamic-pituitary-adrenal axis that stimulates stress-induced activities and motor functions [143, 144]. Circadian rhythm dysfunction alters the melatonin secretion in individuals, which promotes schizophrenia [145]. SNPs studies indicated that mutation in Ror β, Per2, Per3, and neuronal PAS domain protein 2 (Npas2) were associated with schizophrenia [146]. Peripheral tissues of schizophrenia tissues also showed arrhythmic expression of Clock, Per2, and Cry1 genes [147].

    Cardiovascular disorder

    Disruption of circadian rhythm is associated with cardiovascular diseases (CVD). Diurnal rhythms have been observed in heart rate, blood pressure, blood catecholamine secretion, cortisol secretion, platelet aggregation, and others. Changes in circadian rhythm modulate the sympathetic activity, secretion of catecholamines, and aldosterone biosynthesis, as the sympathetic activity controls the adrenal medullary functions as well as the renin-angiotensin system. The synergistic effect is impaired heart rate and blood pressure. Alteration in the circadian clock system also changes the fuel metabolism in cardiomyocytes. Normally, nicotinamide phosphoribosyltransferase (NAMPT) regulates NAD biosynthesis. Dysregulation of NAD synthesis alters the utilization of fatty acids, the primary fuel source. This activity is mediated by the peroxisome proliferators activated receptor γ (PPARγ) signaling system. Circadian dysfunction in endothelial cells alters the activity of Akt and nitric oxide (NO) signaling. NO is essential for vascular relaxation. Dysregulation of NO activity initiates hypertension (Figure 2) [148150].

    Effects of circadian clock dysfunction on different tissues and progression of various disorders. NMDAR: NMDA receptor; AMPAR: AMPA receptor; Cyp7A1: cholesterol 7α-hydroxylase; HMG-CoA: 3-hydroxy-3-methylglutaryl coenzyme A; PEPCK: phosphoenolpyruvate carboxykinase; NAFLD: non-alcoholic fatty liver disease; HCC: hepatocellular carcinoma; GI: gastrointestinal; E: epinephrine; NE: norepinephrine; CHK 1: checkpoint kinase 1; WEE1: a specific G2 check point kinase; p53: tumor suppressor protein 53; ↑: increase; ↓: decrease

    Metabolic disorder

    In the recent era, metabolic disorders are a serious problem in our society. These include obesity, hyperglycemia, type 2 diabetes, dyslipidemia, and hypertension [149, 151, 152]. Metabolic disorders alter blood triglyceride, high-density lipoprotein, and fasting blood glucose levels, as well as inflammatory markers [C-reactive protein, tumor necrosis factor-α (TNF-α), interleukin-6, and plasminogen activator inhibitor type 1] and anti-inflammatory components [149]. Circadian clock disorder is a vital contributor to cancer, diabetes, obesity, and cardiometabolic disorders (Figure 2) [153]. Various nuclear hormone receptors like PPARα, PPARγ, Rev-Erb α, Ror α, hepatocyte nuclear factor 4α (HNF4α), thyroid hormone receptor α (TRα), and nuclear receptor related 1 (NURR1) directly regulate the metabolic pathway. The circadian clock has a link with these nuclear hormone receptors. Imbalance in the circadian clock affects PPARγ-mediated lipogenesis and lipid storage, PPARα-dependent ketogenesis, and hepatic fatty acid oxidation [154]. The occurrence of type 2 diabetes mellitus and obesity are much higher in night shift workers than the day shift workers [153, 155]. An animal model study in an alternative LD cycle revealed that the profile of blood glucose, insulin, and triglyceride levels differ according to the alternative circadian system [156]. Polymorphic variations indicate that clock genes like Clock, Per2, Per3, Bmal2, Cry1, and Cry2 are responsible for metabolic disorders [157]. Mutant animals exhibit chronodisruption and disorder in lipid metabolism by changing the pattern of cholesterol metabolism, expression of lipase enzyme, and production of glycerol and free fatty acids during the circadian period; all of which are associated with adipocyte hypertrophy and dyslipidemia [158]. Gabriel and Zierath [159] reported that skeletal muscle has a strong circadian profile. Shift workers are prone to insulin resistance and type 2 diabetes. Timed exercise is most beneficial for improving health and metabolic diseases. Different organs are directly or indirectly affected by metabolic disorders (Figure 2).

    Hepatic dyslipidemia and steatosis

    The liver maintains the metabolic homeostasis in the body. The hepatic clock genes modulate several metabolic activities like bile acid synthesis, apolipoprotein synthesis, and gluconeogenesis. Clock component Rev-Erb α represses the expression of small heterodimer partner (SHP) and E4 promoter binding protein 4 (E4BP4). These two proteins suppress the expression of bile acid synthesizing enzyme Cyp7A1, which regulates cholesterol and bile acid homeostasis. Circadian clock disruption has an impact on the activity of hepatic Cyp7A1, HMG-CoA reductase, PEPCK. Thus, circadian dysfunction- mediated enzymatic deregulation increases the susceptibility to dyslipidemia, steatosis, and hyperglycemia (Figure 2) [149, 160, 161]. The reduced form of redox cofactors NAD(H) and NADP(H) have an impact on the activity of CLOCK/BMAL1 and NPAS2/BMAL1 [162]. The OS-induced reduction of NADPH levels alters the CLOCK/BMAL1-induced transcriptional activity. Circadian rhythm oscillates NAD levels in the liver. The expression of the Nampt gene is upregulated by the CLOCK/BMAL1 complex. The RNA levels of NAMPT and concentration of NAD decrease in the liver of ClockΔ19/Δ19 and Bmal1−/− mice; alternatively, CRY1 and CRY2 mutation in mice show a higher level of these two factors [149]. A high-fat diet (HFD) can alter the expression of core clock genes such as Clock, Bmal1, Per1, Per2, Cry1, and Cry2 in the SCN, hypothalamus, adipose tissue, and liver. A HFD modulates the activity of acetylated H3K27, Pol II binding, nuclear factor kappa B (NF-κB)-mediated transcriptional activity, PPARγ inducing functions, and metabolic activity of gut microbiota. The collective effects are metabolic disorder, inflammatory response, progression of type 2 diabetes, and CVD [152, 163]. Jacobi et al. [164] reported that BMAL1 regulates mitochondrial biogenesis, as well as mitochondrial fission and mitophagy in hepatic cells of mice. This activity correlates between the circadian clock and metabolic rhythm. Inactivation of the BMAL1 function hampers metabolic activities, starts mitochondrial swelling, and increases OS.

    Adipose tissue and obesity

    Adipose tissue influences adipogenesis in the subcutaneous visceral area. Clock genes regulate the activation of SIRT1, which increases insulin sensitivity and reduces the inflammatory response in adipocytes [149, 165]. Circadian dysregulation changes the expression patterns of enzymes involved in lipid metabolism, insulin sensitivity, obesity, as well as inflammatory response (Figure 2). Moreover, a HFD alters the functions of clock control genes and fuel metabolism. These effects disturb fatty acid mobility and contribute to lipotoxicity [149, 166]. The obesity-induced inflammatory response also initiates steatosis, NAFLD, HCC, and CVD. HFD has a great impact on adipogenesis, food consumption, and obesity. HFD and impaired expression of clock genes in adipose tissue change the pattern of fuel metabolism [166]. Deletion of Bmal1 in mouse adipocytes modulates the adipose functions and the activity of lateral hypothalamic neurons of the feeding center to increase food consumption [167]. Previously, Xu et al. [168] reported that disruption of clock genes in fat cells of Drosophila increases starvation sensation and food consumption. Another report indicated that clock genes knockout mice have a prevalence of obesity [153].

    Normally, leptin regulates feeding rhythm. Rhythmic expression of leptin has been observed in mice, which oscillates along with the LD cycle [169]. Alternatively, 24 h leptin and adiponectin profiles do not maintain in obese humans [170]. Leptin and adiponectin exert their signaling via PKA in muscle and liver [171]. Adipose tissue secretes adipokines, which influence proinflammatory activity. Obesity induces the secretion of interleukin-6, TNF-α and also increases the activity of the endoplasmic reticulum (ER) in the liver, adipocytes, and pancreas. ER stress starts over-secretion of adipokine from adipose tissue and enhances the accumulation of reactive oxygen species [172]. Thus, impaired circadian clock function alters adipose tissue functions, particularly leptin, adiponectin, proinflammatory cytokine secretion, and OS response. The synergistic effect of these factors promotes metabolic disorders. In humans, the secretion of adipokines occurs in a circadian fashion, which is involved in the regulation of metabolic homeostasis [149, 173]. Kettner et al. [174] reported that the study of chronic jet lag in the experimental animals showed impaired leptin activity and obesity, which indicates the relationship between circadian misalignment and metabolic disorder.

    Alteration in insulin secretion, hyperglycemia, and type 2 diabetes

    Blood glucose level and insulin action rhythmically change across the day. The central clock and in association with peripheral clocks control the feeding pattern, energy output, and insulin sensitivity in the body [175]. Misalignment in the circadian clock changes the glucose tolerance capacity and promotes type 2 diabetes. The clock gene mutant mice showed hypoinsulinemic hyperglycemia (Figure 2) [153, 176, 177]. Elevated fasting glucose was observed in polymorphic CRY2 [178, 179]. Melatonin has a role in insulin secretion and inhibits glucose-induced insulin release. Impaired circadian rhythm changes the pattern of melatonin secretion and reduces the glucose-induced insulin release from pancreatic β-cells, leading to hyperglycemia [180]. CLOCK-BMAL1 complex interacts with the circadian rhythm-regulating distal regions in pancreatic β-cells and directly influences insulin synthesis and release [181].

    CLOCK-BMAL1 complex also regulates the expression of NAMPT and SIRT1. Both are essential for the biosynthesis of NAD, gluconeogenesis in the liver, glucose-induced insulin secretion from the pancreas, and glucose metabolism [149]. Impaired glucose tolerance had observed in Nampt-deficient (Nampt−/−) mice. Intra-peritoneal injection of nicotinamide mononucleotide (NMN) improves the condition [182]. However, overexpression of SIRT1 in pancreatic β-cells exhibited better glucose tolerance and insulin secretion [183]. Stenvers et al. [175] reported that chronodisruption was a factor in the development of insulin resistance, followed by the progression of type 2 diabetes. Core clock gene mutation, shift work or night work, exposure to artificial light-at-night, and sleep disturbance impact glucose metabolism and insulin sensitivity. Thus, dysfunction of the biological clock decreases insulin release, insulin activity, and glucose metabolism, leading to the advancement of the conditions of type 2 diabetes.

    Circadian dysregulation and cancer progression

    SCN-mediated dysfunction of circadian events advances cancers in the different tissues. Clock components regulate the expression of factors related to the cell cycle, DNA repair, metabolism, redox homeostasis, and others. They control the secretion of cytokines, hormones, neurotransmitters, tumor-specific metabolites, and the activity of oncogenic proteins. Collectively, disruption of circadian function alters cellular functions, leading to the progression of cancer development (Figure 2) [184]. In our modern society, people are involved in shift work and night work. They have been exposed to bright light-at-night in industrial sectors, the medical profession, transport services, and others, resulting in misalignment of circadian rhythm that leads to disruption of the sleep-wake cycle, metabolic disorder, and cancer progression. The epidemiological and experimental studies revealed that circadian dysfunction was associated with cancer. International Agency for Research on Cancer (IARC) has categorized the shift-work as “carcinogenic to humans (Group 2A)” [185]. A recent report indicates that night shift work enhances the risks of breast, prostate, and colorectal cancer [186188].

    The study of SNP indicates the role of clock genes in disease progression, including cancer [6, 28, 157, 189]. Any alteration or improper expression of clock genes such as Clock, Bmal1, Ck1ε, Cry1–2, Npas2, and Per1–3 promotes the formation of different cancers like breast, colorectal, prostate, pancreatic, and non-Hodgkin lymphoma [6, 28, 190]. Deregulation of expression of the core clock genes Per1, Per2, Per3, Cry1, Cry2, Bmal1, and Clock are associated with the development of cancer in humans [30, 191]. Impaired expression of Per1, Per2 and/or Per3 had been observed in several cancers, including breast [191, 192], prostate [193, 194], oral [195], pancreatic [196, 197], hepatocellular [198], colorectal [199], non-small cell lung cancer [200], renal [201], bladder [202], myeloid leukemia [203, 204], glioma [205], and squamous cell carcinoma [206]. The cell cycle suppressor protein WEE1 is also regulated by PER and CRY. Alteration in Cry expression deregulates the synthesis of WEE1 (a specific G2 check point kinase, cell cycle suppressor protein) and cyclin D1; the result is impaired cell proliferation and tumor growth [207, 208]. Mutation in the Cry2 gene increases the risk of lymphoma [209]. Valenzuela et al. [157] reported that polymorphisms of Clock, Per2, Per3, Bmal1, Cry1, Cry2, Tim, and Ror β genes are associated with breast colorectal and prostate cancer. Variants of Cry genes increase the risk of non-Hodgkin’s lymphoma [209, 210]. Circadian disruption causes dyslipidemia as well as NAFLD [211]. SNP of Per3 and Cry1 genes promotes HCC [212]. Shift work changes the lifestyle and pattern of metabolism that cause NAFLD and then non-alcoholic steatohepatitis (NASH); the result is hepatic fibrosis followed by cirrhosis and finally HCC [213]. Bile acid can influence the progression of HCC [211]. Circadian disruption affects cholesterol, bile acid, and xenobiotic metabolism. Bile acids promote non-genotoxic HCC by inducing NAFLD, and NASH. Chronodisruption increases the risk of ovarian cancers in females who are involved in night shift work. This might be associated with inappropriate expression of circadian genes [214]. Jim et al. [215] reported that variation in Arntl, Cry2, CK1ε, NpasS2, Per3, and Rev1 genes influences ovarian cancer.

    Infertility disorder

    Human infertility disorder is multifactorial, and one-third of the cases are male-associated. The activity of the reproductive system depends on the expression of the steroidogenic acute regulatory protein (STAR), rate of steroidogenesis, and luteinizing hormone (LH) surge before ovulation in females; while maintenance of sperm attributes in males are also crucial. All these activities are directly or indirectly associated with clock gene expression, as the SNP of Clock and Bmal1 is correlated with infertility [216218].

    Circadian disruption, physical performance, and accident rates

    In our modern life, several activities such as medical services, uninterrupted power supply, operation of nuclear power plants, transport services, monitoring of satellite movements and the space shuttle, continuous operation of industrial work, and military activities are going on round the clock. In these sectors, people are working on different shifts. Sometimes the shifts are extended due to a shortage of manpower, and demand for service. The shifting duties, as well as extended shift, disrupts the circadian clock functions and various physiological rhythms (sleep cycle, melatonin secretion, core body temperature regulation, ACTH/cortisol secretion, plasma triacylglycerol, glucose homeostasis, and metabolic activities) of the body [14, 35]. Disruption of circadian rhythm is detrimental to health, particularly for CVD [219]. Circadian dysfunction also decreases alertness, cognitive functions, mood, social activities, and performance of physical work of an individual [220, 221]. Physical performance of the body depends on hormone secretion, core body temperature, and metabolic activity. The healthy individuals with normal LD cycles (12:12 LD) exhibit the highest physical performance and work efficiency in the daytime. The work capacity tends to decrease from the evening and reaches the minimum level within the period 2.00 A.M. to 4.00 A.M. when the core body temperature remains at the lowest value [222]. During this period, the body feels fatigued, which increases the chance of an accident.

    The continuous phase changes in shift work and night work (about 20% in the industrialized nations) cause potential health hazards, and safety problems [219]. Many accidents (road, industrial, power plant) happen due to ill adaptation of circadian clock functions, feeling of fatigue, poor physical performance, decreased alertness, and mental disturbance [223225]. A survey-based study related to the rate of vehicle accidents revealed that the number of accidents/collisions maximally happened after 2.00 A.M. Despite many reasons for the accident, the dysfunction of circadian events could not be ignored [226]. The major disastrous events like the Exxon Valdez supertanker accident in Alaska, the Chernobyl disaster (26th April 1986 at 1.23 A.M.), and the U.S. commercial nuclear power plant accident at 3 Mile Island, Pennsylvania (28th March 1979 at 4.00 A.M.) happened at the late night and circadian dysfunction mediated poor physical performance might be associated in these mishaps [226228].

    Circadian rhythm and jet lag

    Jet lag is a transient disruption of the circadian system during transmeridian flights. It is the result of a temporary loss of synchrony between biological clock functions and an alternate LD cycle in the new time zone. The result is an abrupt change in the sleep-wake cycle according to the local time. Jet lag disrupts the feeding/fasting cycle, rest/activity cycle, and sleep-wake cycle [35]. The common symptoms of jet lag are fatigue and sleepiness during the day in the new time zone, inability to satisfactory sleep at night, loss of appetite, indigestion, cognitive distress, inability to co-ordinate with activity, loss of motivation, irritability, headache, and obesity [229231]. The exogenous and endogenous components of the circadian system are mismatched until the internal clock is gradually re-entrained. The profiles of the body temperature, sleep-wake cycle, and endocrine functions have readjusted to local time, and the discomforts of jet lag disappear slowly. Recently, the use of melatonin is more popular to adjust the circadian rhythm rapidly. Light treatment may also give beneficial effects. However, light treatment and melatonin treatment counteract each other as melatonin production is reduced by light [232, 233]. Chronic jet lag is a common phenomenon for shift workers, flight attendants, and flight crews. Different studies indicated that chronic jet lag influences the elevated level of cortisol, poor cognitive functions, low working memory, learning problems, and higher rates of breast and prostate cancer [234, 235].

    Conclusions

    The circadian system acts as a biological timekeeper and regulates physiological functions throughout the day in relation to environmental time cues. The primary activity of the circadian system occurs within the SCN, which makes synchronization with the endogenous peripheral clocks. SCN receives a light-associated impulse from the retina that governs the expression of the core clock genes. Misalignment in the circadian rhythm adversely affects human health. There are many anthropogenic causes of circadian clock dysfunction; these include rotating shift work, night work, exposure to bright light-at-night, chronic jet lag, and modern lifestyle. Circadian dysfunction influences the development of sleep problems, neurodegenerative diseases, psychological illness, metabolic syndrome (diabetes, dyslipidemias, obesity, etc.), cardiovascular disorder, and cancer. In recent times, these disorders are the major public health issues and people are facing serious problems in this concern.

    Proper maintenance of circadian rhythm is essential for health benefits as modern society is tremendously affected by artificial light, workload, jet jag, shift work, etc. Understanding the molecular mechanism of clock functions provides an opportunity for the development of chrono-pharmacology that can reset the biological clock to a new time zone. This approach will be helpful for the treatment of metabolic disorders, jet lag problems, sleep disorders, and some neuropsychiatric illnesses. Nocturnal lifestyles increase the chance of cancer progression. An approach to regulate circadian functions would be a preventive measure against cancer development. Managing external factors like the exposure to artificial light (timing, amount, and spectral composition), the timing of meal or feeding patterns, changing work schedules, and proper lighting at the workplace and home will be the preventive measure against the adverse effects of chronodisruption. Integrative research on epidemiological data, molecular aspects of circadian functions, and circadian dysfunction-related health problems will open new avenues for the study of several health issues and the development of the treatment strategies of chrono-medicine. Finally, it can say that increased public awareness about circadian dysfunction mediated disorder and an approach to maintain the normal alignment of the circadian clock will improve physiological functions and give health benefits in the near future.

    Abbreviation

    Aβ:

    amyloid beta

    AD:

    Alzheimer’s disease

    ASPD:

    advanced sleep phase disorder

    BD:

    bipolar disorder

    BMAL1:

    brain and muscle aryl hydrocarbon receptor nuclear translocator-like 1

    CK1:

    casein kinase 1

    CLOCK:

    circadian locomotor output cycles kaput

    CREB:

    cyclic adenosine monophosphate response element binding protein

    CREs:

    calcium/cyclic adenosine monophosphate response elements

    CRSDs:

    circadian rhythm sleep disorders

    Cry:

    cryptochrome

    CVD:

    cardiovascular diseases

    Cyp7A1:

    cholesterol 7α-hydroxylase

    ERK1/2:

    extracellular signal-regulated kinase 1/2

    Fbxl3:

    F-box/LRR repeat protein 3

    FRD:

    free-running disorder

    GLUT2:

    glucose transporter-2

    HCC:

    hepatocellular carcinoma

    HD:

    Huntington’s disease

    HFD:

    high-fat diet

    HMG-CoA:

    3-hydroxy-3-methylglutaryl coenzyme A

    ISWR:

    irregular sleep-wake rhythm

    LD:

    light-dark

    MDD:

    major depressive disorder

    NAD:

    nicotinamide adenine dinucleotide

    NADPH:

    nicotinamide adenine dinucleotide phosphate

    NAFLD:

    non-alcoholic fatty liver disease

    NAMPT:

    nicotinamide phosphoribosyltransferase

    NO:

    nitric oxide

    Npas2:

    neuronal period-aryl hydrocarbon receptor nuclear translocator single-minded protein 2

    OS:

    oxidative stress

    PD:

    Parkinson’s disease

    PEPCK:

    phosphoenolpyruvate carboxykinase

    Per1:

    Period 1

    PPARγ:

    peroxisome proliferators activated receptor γ

    PSP:

    progressive supranuclear palsy

    Rev-Erb:

    reverse-erythroblastosis

    RHT:

    retinohypothalamic tract

    Ror:

    retinoid-related orphan receptor

    RORE:

    retinoid-related orphan receptor response element

    SCN:

    suprachiasmatic nucleus

    SIRT1:

    silent information regulator transcript 1

    SNPs:

    single nucleotide polymorphisms

    Declarations

    Acknowledgments

    The author is grateful to Midnapore College, Midnapore, West Bengal, India, for providing all kinds of facilities to prepare this manuscript. The author expresses his sincere thanks to Dr. Amal Kanti Chakraborty, Ex-Associate Professor of English, Midnapore College Midnapore, for his convenient help during the linguistic correction.

    Author contributions

    SAA initially started literature searching and preparation of the manuscript. SS guided the overall work. He involved in the writing and correction of the manuscript and gave the final shape of this review article.

    Conflicts of interest

    The authors declare that there is no conflict of interest.

    Ethical approval

    Not applicable.

    Consent to participate

    Not applicable.

    Consent to publication

    Not applicable.

    Availability of data and materials

    Not applicable.

    Funding

    Not applicable.

    Copyright

    © The Author(s) 2022.

    References

    Baker FC, Driver HS. Circadian rhythms, sleep, and the menstrual cycle. Sleep Med. 2007;8:61322. [DOI] [PubMed]
    Bass J, Lazar MA. Circadian time signatures of fitness and disease. Science. 2016;354:9949. [DOI] [PubMed]
    Halberg F. Chronobiology. Annu Rev Physiol. 1969;31:675726. [DOI] [PubMed]
    Bass J. Circadian topology of metabolism. Nature. 2012;491:34856. [DOI] [PubMed]
    Masri S, Sassone-Corsi P. The circadian clock: a framework linking metabolism, epigenetics and neuronal function. Nat Rev Neurosci. 2013;14:6975. [DOI] [PubMed] [PMC]
    Benna C, Helfrich-Förster C, Rajendran S, Monticelli H, Pilati P, Nitti D, et al. Genetic variation of clock genes and cancer risk: a field synopsis and meta-analysis. Oncotarget. 2017;8:2397895. [DOI] [PubMed] [PMC]
    Farhud D, Aryan Z. Circadian rhythm, lifestyle and health: a narrative review. Iran J Public Health. 2018;47:106876. [PubMed] [PMC]
    Tsuchiya Y, Umemura Y, Yagita K. Circadian clock and cancer: from a viewpoint of cellular differentiation. Int J Urol. 2020;27:51824. [DOI] [PubMed]
    Pariollaud M, Lamia KA. Cancer in the fourth dimension: what is the impact of circadian disruption? Cancer Discov. 2020;10:145564. [DOI] [PubMed] [PMC]
    Klein DC, Moore RY, Reppert S. The suprachiasmatic nucleus: the mind’s clock. New York: Oxford university press; 1991. p. 467.
    Hannibal J, Hindersson P, Ostergaard J, Georg B, Heegaard S, Larsen PJ, et al. Melanopsin is expressed in PACAP-containing retinal ganglion cells of the human retinohypothalamic tract. Invest Ophthalmol Vis Sci. 2004;45:42029. [DOI] [PubMed]
    Schibler U, Gotic I, Saini C, Gos P, Curie T, Emmenegger Y, et al. Clock-talk: interactions between central and peripheral circadian oscillators in mammals. Cold Spring Harb Symp Quant Biol. 2015;80:22332. [DOI] [PubMed]
    Hastings MH, Maywood ES, Brancaccio M. Generation of circadian rhythms in the suprachiasmatic nucleus. Nat Rev Neurosci. 2018;19:45369. [DOI] [PubMed]
    Waterhouse JM, DeCoursey PJ. The relevance of circadian rhythms for human welfare. In: Dunlap JC, Loros JJ, DeCoursey PJ, editor. Chronobiology: biological timekeeping. Sunderland, MA: Sinauer associates international publisher; 2004. pp. 32553.
    Dibner C. The importance of being rhythmic: living in harmony with your body clocks. Acta Physiol (Oxf). 2020;228:e13281. [DOI] [PubMed]
    Ruan W, Yuan X, Eltzschig HK. Circadian rhythm as a therapeutic target. Nat Rev Drug Discov. 2021;20:287307. [DOI] [PubMed] [PMC]
    Reppert SM, Weaver DR. Molecular analysis of mammalian circadian rhythms. Annu Rev Physiol. 2001;63:64776. [DOI] [PubMed]
    Welsh DK, Takahashi JS, Kay SA. Suprachiasmatic nucleus: cell autonomy and network properties. Annu Rev Physiol. 2010;72:55177. [DOI] [PubMed] [PMC]
    Hattar S, Lucas RJ, Mrosovsky N, Thompson S, Douglas RH, Hankins MW, et al. Melanopsin and rod-cone photoreceptive systems account for all major accessory visual functions in mice. Nature. 2003;424:7681. [DOI] [PubMed] [PMC]
    Do MT, Yau KW. Intrinsically photosensitive retinal ganglion cells. Physiol Rev. 2010;90:154781. [DOI] [PubMed] [PMC]
    Ding JM, Buchanan GF, Tischkau SA, Chen D, Kuriashkina L, Faiman LE, et al. A neuronal ryanodine receptor mediates lightinduced phase delays of the circadian clock. Nature. 1998;394:3814. [DOI] [PubMed]
    Carlezon WA Jr, Duman RS, Nestler EJ. The many faces of CREB. Trends Neurosci. 2005;28:43645. [DOI] [PubMed]
    Tischkau SA, Mitchell JW, Tyan SH, Buchanan GF, Gillette MU. Ca2+/cAMP response element-binding protein (CREB)-dependent activation of Per1 is required for light-induced signaling in the suprachiasmatic nucleus circadian clock. J Biol Chem. 2003;278:71823. [DOI] [PubMed]
    Welsh DK, Yoo SH, Liu AC, Takahashi JS, Kay SA. Bioluminescence imaging of individual fibroblasts reveals persistent, independently phased circadian rhythms of clock gene expression. Curr Biol. 2004;14:228995. [DOI] [PubMed] [PMC]
    Patel SA, Kondratov RV. Clock at the core of cancer development. Biology (Basel). 2021;10:150. [DOI] [PubMed] [PMC]
    Toh KL. Basic science review on circadian rhythm biology and circadian sleep disorders. Ann Acad Med Singap. 2008;37:6628. [PubMed]
    Foster RG, Kreitzman L. The rhythms of life: what your body clock means to you! Exp Physiol. 2014;99:599606. [DOI] [PubMed]
    Masri S, Sassone-Corsi P. The emerging link between cancer, metabolism, and circadian rhythms. Nat Med. 2018;24:1795803. [DOI] [PubMed] [PMC]
    Morales-Santana S, Morell S, Leon J, Carazo-Gallego A, Jimenez-Lopez JC, Morell M. An overview of the polymorphisms of circadian genes associated with endocrine cancer. Front Endocrinol (Lausanne). 2019;10:104. [DOI] [PubMed] [PMC]
    Ye Y, Xiang Y, Ozguc FM, Kim Y, Liu CJ, Park PK, et al. The genomic landscape and pharmacogenomic interactions of clock genes in cancer chronotherapy. Cell Syst. 2018;6:31428.e2. [DOI] [PubMed] [PMC]
    Zhang Z, Zeng P, Gao W, Zhou Q, Feng T, Tian X. Circadian clock: a regulator of the immunity in cancer. Cell Commun Signal. 2021;19:37. [DOI] [DOI] [PubMed] [PMC]
    Partch CL, Green CB, Takahashi JS. Molecular architecture of the mammalian circadian clock. Trends Cell Biol. 2014;24:909. [DOI] [PubMed] [PMC]
    Pett JP, Korenčič A, Wesener F, Kramer A, Herzel H. Feedback loops of the mammalian circadian clock constitute repressilator. PLoS Comput Biol. 2016;12:e1005266. [DOI] [PubMed] [PMC]
    Gallego M, Virshup DM. Post-translational modifications regulate the ticking of the circadian clock. Nat Rev Mol Cell Biol. 2007;8:13948. [DOI] [PubMed]
    Potter GD, Skene DJ, Arendt J, Cade JE, Grant PJ, Hardie LJ. Circadian rhythm and sleep disruption: causes, metabolic consequences, and countermeasures. Endocr Rev. 2016;37:584608. [DOI] [PubMed] [PMC]
    Angelousi A, Kassi E, Nasiri-Ansari N, Weickert MO, Randeva H, Kaltsas G. Clock genes alterations and endocrine disorders. Eur J Clin Invest. 2018;48:e12927. [DOI] [PubMed]
    Akashi M, Tsuchiya Y, Yoshino T, Nishida E. Control of intracellular dynamics of mammalian period proteins by casein kinase I epsilon (CKIepsilon) and CKIdelta in cultured cells. Mol Cell Biol. 2002;22:1693703. [DOI] [PubMed] [PMC]
    Yang Y, Xu T, Zhang Y, Qin X. Molecular basis for the regulation of the circadian clock kinases CK1δ and CK1ε. Cell Signal. 2017;31:5865. [DOI] [PubMed]
    Busino L, Bassermann F, Maiolica A, Lee C, Nolan PM, Godinho SI, et al. SCFFbxl3 controls the oscillation of the circadian clock by directing the degradation of cryptochrome proteins. Science. 2007;316:9004. [DOI] [PubMed]
    Preitner N, Damiola F, Lopez-Molina L, Zakany J, Duboule D, Albrecht U, et al. The orphan nuclear receptor REV-ERBalpha controls circadian transcription within the positive limb of the mammalian circadian oscillator. Cell. 2002;110:25160. [DOI] [PubMed]
    Kwon I, Choe HK, Son GH, Kim K. Mammalian molecular clocks. Exp Neurobiol. 2011;20:1828. [DOI] [PubMed] [PMC]
    Jin X, Shearman LP, Weaver DR, Zylka MJ, de Vries GJ, Reppert SM. A molecular mechanism regulating rhythmic output from the suprachiasmatic circadian clock. Cell. 1999;96:5768. [DOI] [PubMed]
    Gekakis N, Staknis D, Nguyen HB, Davis FC, Wilsbacher LD, King DP, et al. Role of the CLOCK protein in the mammalian circadian mechanism. Science. 1998;280:15649. [DOI] [PubMed]
    Landgraf D, Wang LL, Diemer T, Welsh DK. NPAS2 compensates for loss of CLOCK in peripheral circadian oscillators. PLoS Genet. 2016;12:e1005882. [DOI] [PubMed] [PMC]
    Mure LS, Le HD, Benegiamo G, Chang MW, Rios L, Jillani N, et al. Diurnal transcriptome atlas of a primate across major neural and peripheral tissues. Science. 2018;359:eaao0318. [DOI] [PubMed] [PMC]
    Zylka MJ, Shearman LP, Weaver DR, Reppert SM. Three period homologs in mammals: differential light responses in the suprachiasmatic circadian clock and oscillating transcripts outside of brain. Neuron. 1998;20:110310. [DOI] [PubMed]
    Izumo M, Pejchal M, Schook AC, Lange RP, Walisser JA, Sato TR, et al. Differential effects of light and feeding on circadian organization of peripheral clocks in a forebrain Bmal1 mutant. Elife. 2014;3:e04617. [DOI] [PubMed] [PMC]
    Yamazaki S, Numano R, Abe M, Hida A, Takahashi R, Ueda M, et al. Resetting central and peripheral circadian oscillators in transgenic rats. Science. 2000;288:6825. [DOI] [PubMed]
    Husse J, Leliavski A, Tsang AH, Oster H, Eichele G. The light-dark cycle controls peripheral rhythmicity in mice with a genetically ablated suprachiasmatic nucleus clock. FASEB J. 2014;28:495060. [DOI] [PubMed]
    Eckel-Mahan KL, Patel VR, Mohney RP, Vignola KS, Baldi P, Sassone-Corsi P. Coordination of the transcriptome and metabolome by the circadian clock. Proc Natl Acad Sci U S A. 2012;109:55416. [DOI] [PubMed] [PMC]
    Stokkan KA, Yamazaki S, Tei H, Sakaki Y, Menaker M. Entrainment of the circadian clock in the liver by feeding. Science. 2001;291:4903. [DOI] [PubMed]
    Damiola F, Le Minh N, Preitner N, Kornmann B, Fleury-Olela F, Schibler U. Restricted feeding uncouples circadian oscillators in peripheral tissues from the central pacemaker in the suprachiasmatic nucleus. Genes Dev. 2000;14:295061. [DOI] [PubMed] [PMC]
    Oosterman JE, Kalsbeek A, la Fleur SE, Belsham DD. Impact of nutrients on circadian rhythmicity. Am J Physiol Regul Integr Comp Physiol. 2015;308:R33750. [DOI] [PubMed] [PMC]
    Dallmann R, Viola AU, Tarokh L, Cajochen C, Brown SA. The human circadian metabolome. Proc Natl Acad Sci U S A. 2012;109:26259. [DOI] [PubMed] [PMC]
    Vollmers C, Gill S, DiTacchio L, Pulivarthy SR, Le HD, Panda S. Time of feeding and the intrinsic circadian clock drive rhythms in hepatic gene expression. Proc Natl Acad Sci U S A. 2009;106:214538. [DOI] [PubMed] [PMC]
    Greenwell BJ, Trott AJ, Beytebiere JR, Pao S, Bosley A, Beach E, et al. Rhythmic food intake drives rhythmic gene expression more potently than the hepatic circadian clock in mice. Cell Rep. 2019;27:64957.e5. [DOI] [PubMed]
    Kornmann B, Schaad O, Bujard H, Takahashi JS, Schibler U. System-driven and oscillator-dependent circadian transcription in mice with a conditionally active liver clock. PLoS Biol. 2007;5:e34. [DOI] [PubMed] [PMC]
    Gender, work and health [Internet]. World Health Organization; c2011 [cited 2019 Nov 16]. Available from: http://apps.who.int/iris/bitstream/10665/97940/1/9789241501729_eng.pdf?ua=1
    Wyss-Coray T. Ageing, neurodegeneration and brain rejuvenation. Nature. 2016;539:1806. [DOI] [PubMed] [PMC]
    Gitler AD, Dhillon P, Shorter J. Neurodegenerative disease: models, mechanisms, and a new hope. Dis Model Mech. 2017;10:499502. [DOI] [PubMed] [PMC]
    Hood S, Amir S. Neurodegeneration and the circadian clock. Front Aging Neurosci. 2017;9:170. [DOI] [PubMed] [PMC]
    Leng Y, Musiek ES, Hu K, Cappuccio FP, Yaffe K. Association between circadian rhythms and neurodegenerative diseases. Lancet Neurol. 2019;18:30718. [DOI] [PubMed] [PMC]
    Logan RW, McClung CA. Rhythms of life: circadian disruption and brain disorders across the lifespan. Nat Rev Neurosci. 2019;20:4965. [DOI] [PubMed] [PMC]
    Werdann M, Zhang Y. Circadian rhythm and neurodegenerative disorders. Brain Sci Adv. 2020;6:7180. [DOI]
    Fifel K, Videnovic A. Circadian alterations in patients with neurodegenerative diseases: neuropathological basis of underlying network mechanisms. Neurobiol Dis. 2020;144:105029. [DOI] [PubMed]
    Carter B, Justin HS, Gulick D, Gamsby JJ. The molecular clock and neurodegenerative disease: a stressful time. Front Mol Biosci. 2021;8:644747. [DOI] [PubMed] [PMC]
    Nassan M, Videnovic A. Circadian rhythms in neurodegenerative disorders. Nat Rev Neurol. 2022;18:724. [DOI] [PubMed]
    Colwell CS. Defining circadian disruption in neurodegenerative disorders. J Clin Invest. 2021;131:e148288. [DOI] [PubMed]
    Dugger BN, Dickson DW. Pathology of neurodegenerative diseases. Cold Spring Harb Perspect Biol. 2017;9:a028035. [DOI] [PubMed] [PMC]
    Gan L, Cookson MR, Petrucelli L, La Spada AR. Converging pathways in neurodegeneration, from genetics to mechanisms. Nat Neurosci. 2018;21:13009. [DOI] [PubMed] [PMC]
    Muddapu VR, Dharshini SAP, Chakravarthy VS, Gromiha MM. Neurodegenerative diseases - is metabolic deficiency the root cause? Front Neurosci. 2020;14:213. [DOI] [PubMed] [PMC]
    Noya SB, Colameo D, Brüning F, Spinnler A, Mircsof D, Opitz L, et al. The forebrain synaptic transcriptome is organized by clocks but its proteome is driven by sleep. Science. 2019;366:eaav2642. [DOI] [PubMed]
    Wen S, Ma D, Zhao M, Xie L, Wu Q, Gou L, et al. Spatiotemporal single-cell analysis of gene expression in the mouse suprachiasmatic nucleus. Nat Neurosci. 2020;23:45667. [DOI] [PubMed]
    Cermakian N, Lamont EW, Boudreau P, Boivin DB. Circadian clock gene expression in brain regions of Alzheimer’s disease patients and control subjects. J Biol Rhythms. 2011;26:16070. [DOI] [PubMed]
    Chen CY, Logan RW, Ma T, Lewis DA, Tseng GC, Sibille E, et al. Effects of aging on circadian patterns of gene expression in the human prefrontal cortex. Proc Natl Acad Sci U S A. 2016;113:20611. [DOI] [PubMed] [PMC]
    Lananna BV, Musiek ES. The wrinkling of time: aging, inflammation, oxidative stress, and the circadian clock in neurodegeneration. Neurobiol Dis. 2020;139:104832. [DOI] [PubMed] [PMC]
    Musiek ES, Lim MM, Yang G, Bauer AQ, Qi L, Lee Y, et al. Circadian clock proteins regulate neuronal redox homeostasis and neurodegeneration. J Clin Invest. 2013;123:5389400. [DOI] [PubMed] [PMC]
    Musiek ES, Holtzman DM. Mechanisms linking circadian clocks, sleep, and neurodegeneration. Science. 2016;354:10048. [DOI] [PubMed] [PMC]
    Asher G, Gatfield D, Stratmann M, Reinke H, Dibner C, Kreppel F, et al. SIRT1 regulates circadian clock gene expression through PER2 deacetylation. Cell. 2008;134:31728. [DOI] [PubMed]
    Chang HC, Guarente L. SIRT1 mediates central circadian control in the SCN by a mechanism that decays with aging. Cell. 2013;153:144860. [DOI] [PubMed] [PMC]
    Singh P, Hanson PS, Morris CM. SIRT1 ameliorates oxidative stress induced neural cell death and is down-regulated in Parkinson’s disease. BMC Neurosci. 2017;18:46. [DOI] [PubMed] [PMC]
    Samanta S. Physiological and pharmacological perspectives of melatonin. Arch Physiol Biochem. 2020;[Epub ahead of print]. [DOI] [PubMed]
    Samanta S. Melatonin: an endogenous miraculous indolamine, fights against cancer progression. J Cancer Res Clin Oncol. 2020;146:1893922. [DOI] [PubMed]
    Han KS, Kim L, Shim I. Stress and sleep disorder. Exp Neurobiol. 2012;21:14150. [DOI] [PubMed] [PMC]
    Caillet-Boudin ML, Buée L, Sergeant N, Lefebvre B. Regulation of human MAPT gene expression. Mol Neurodegener. 2015;10:28. [DOI] [PubMed] [PMC]
    Irwin DJ. Tauopathies as clinicopathological entities. Parkinsonism Relat Disord. 2016;22:S2933. [DOI] [PubMed] [PMC]
    Strang KH, Golde TE, Giasson BI. MAPT mutations, tauopathy, and mechanisms of neurodegeneration. Lab Invest. 2019;99:91228. [DOI] [PubMed] [PMC]
    Steiner H, Fukumori A, Tagami S, Okochi M. Making the final cut: pathogenic amyloid-β peptide generation by γ-secretase. Cell Stress. 2018;2:292310. [DOI] [PubMed] [PMC]
    Yuksel M, Tacal O. Trafficking and proteolytic processing of amyloid precursor protein and secretases in Alzheimer’s disease development: an up-to-date review. Eur J Pharmacol. 2019;856:172415. [DOI] [PubMed]
    Ma Z, Jiang W, Zhang EE. Orexin signaling regulates both the hippocampal clock and the circadian oscillation of Alzheimer’s disease-risk genes. Sci Rep. 2016;6:36035. [DOI] [PubMed] [PMC]
    Wang X, Wang L, Yu Q, Xu Y, Zhang L, Zhao X, et al. Alterations in the expression of Per1 and Per2 induced by Aβ31-35 in the suprachiasmatic nucleus, hippocampus, and heart of C57BL/6 mouse. Brain Res. 2016;1642:518. [DOI] [PubMed]
    Krishnan N, Kretzschmar D, Rakshit K, Chow E, Giebultowicz JM. The circadian clock gene period extends healthspan in aging Drosophila melanogaster. Aging (Albany NY). 2009;1:93748. [DOI] [PubMed] [PMC]
    Dissel S, Melnattur K, Shaw PJ. Sleep, performance, and memory in flies. Curr Sleep Med Rep. 2015;1:4754. [DOI] [PubMed] [PMC]
    Krishnan N, Rakshit K, Chow ES, Wentzell JS, Kretzschmar D, Giebultowicz JM. Loss of circadian clock accelerates aging in neurodegeneration-prone mutants. Neurobiol Dis. 2012;45:112935. [DOI] [PubMed] [PMC]
    Bellanti F, Iannelli G, Blonda M, Tamborra R, Villani R, Romano A, et al. Alterations of clock gene RNA expression in brain regions of a triple transgenic model of Alzheimer’s disease. J Alzheimers Dis. 2017;59:61531. [DOI] [PubMed] [PMC]
    Song H, Moon M, Choe HK, Han DH, Jang C, Kim A, et al. Aβ-induced degradation of BMAL1 and CBP leads to circadian rhythm disruption in Alzheimer’s disease. Mol Neurodegener. 2015;10:13. [DOI] [PubMed] [PMC]
    Kress GJ, Liao F, Dimitry J, Cedeno MR, FitzGerald GA, Holtzman DM, et al. Regulation of amyloid-β dynamics and pathology by the circadian clock. J Exp Med. 2018;215:105968. [DOI] [PubMed] [PMC]
    Kondratova AA, Dubrovsky YV, Antoch MP, Kondratov RV. Circadian clock proteins control adaptation to novel environment and memory formation. Aging (Albany NY). 2010;2:28597. [DOI] [PubMed] [PMC]
    Li G, Yin H, Kuret J. Casein kinase 1 delta phosphorylates tau and disrupts its binding to microtubules. J Biol Chem. 2004;279:1593845. [DOI] [PubMed]
    Chen C, Gu J, Basurto-Islas G, Jin N, Wu F, Gong CX, et al. Up-regulation of casein kinase 1ε is involved in tau pathogenesis in Alzheimer’s disease. Sci Rep. 2017;7:13478. [DOI] [PubMed] [PMC]
    Scott D, Roy S. α-Synuclein inhibits intersynaptic vesicle mobility and maintains recycling-pool homeostasis. J Neurosci. 2012;32:1012935. [DOI] [PubMed] [PMC]
    Kudo T, Loh DH, Truong D, Wu Y, Colwell CS. Circadian dysfunction in a mouse model of Parkinson’s disease. Exp Neurol. 2011;232:6675. [DOI] [PubMed]
    Gu Z, Wang B, Zhang YB, Ding H, Zhang Y, Yu J, et al. Association of ARNTL and PER1 genes with Parkinson’s disease: a case-control study of Han Chinese. Sci Rep. 2015;5:15891. [DOI] [PubMed] [PMC]
    Pacelli C, Rotundo G, Lecce L, Menga M, Bidollari E, Scrima R, et al. Parkin mutation affects clock gene-dependent energy metabolism. Int J Mol Sci. 2019;20:2772. [DOI] [PubMed] [PMC]
    Breen DP, Vuono R, Nawarathna U, Fisher K, Shneerson JM, Reddy AB, et al. Sleep and circadian rhythm regulation in early Parkinson disease. JAMA Neurol. 2014;71:58995. [DOI] [PubMed] [PMC]
    Liu WW, Wei SZ, Huang GD, Liu LB, Gu C, Shen Y, et al. BMAL1 regulation of microglia-mediated neuroinflammation in MPTP-induced Parkinson’s disease mouse model. FASEB J. 2020;34:657081. [DOI] [PubMed]
    Morton AJ, Wood NI, Hastings MH, Hurelbrink C, Barker RA, Maywood ES. Disintegration of the sleep-wake cycle and circadian timing in Huntington’s disease. J Neurosci. 2005;25:15763. Erratum in: J Neurosci. 2005;25:3994. [DOI] [PubMed] [PMC]
    Maywood ES, Fraenkel E, McAllister CJ, Wood N, Reddy AB, Hastings MH, et al. Disruption of peripheral circadian timekeeping in a mouse model of Huntington’s disease and its restoration by temporally scheduled feeding. J Neurosci. 2010;30:10199204. [DOI] [PubMed] [PMC]
    Smarr B, Cutler T, Loh DH, Kudo T, Kuljis D, Kriegsfeld L, et al. Circadian dysfunction in the Q175 model of Huntington’s disease: network analysis. J Neurosci Res. 2019;97:160623. [DOI] [PubMed] [PMC]
    Walsh CM, Ruoff L, Walker K, Emery A, Varbel J, Karageorgiou E, et al. Sleepless night and day, the plight of progressive supranuclear palsy. Sleep. 2017;40:zsx154. [DOI] [PubMed] [PMC]
    Schwab C, DeMaggio AJ, Ghoshal N, Binder LI, Kuret J, McGeer PL. Casein kinase 1 delta is associated with pathological accumulation of tau in several neurodegenerative diseases. Neurobiol Aging. 2000;21:50310. [DOI] [PubMed]
    Pacchierotti C, Iapichino S, Bossini L, Pieraccini F, Castrogiovanni P. Melatonin in psychiatric disorders: a review on the melatonin involvement in psychiatry. Front Neuroendocrinol. 2001;22:1832. [DOI] [PubMed]
    Morera-Fumero AL, Abreu-Gonzalez P. Role of melatonin in schizophrenia. Int J Mol Sci. 2013;14:903750. [DOI] [PubMed] [PMC]
    Tordjman S, Chokron S, Delorme R, Charrier A, Bellissant E, Jaafari N, et al. Melatonin: pharmacology, functions and therapeutic benefits. Curr Neuropharmacol. 2017;15:43443. [DOI] [PubMed] [PMC]
    Sun X, Wang Y, Jiang N, Du Z. The potential role of melatonin on mental disorders: insights from physiology and pharmacology. Bipolar Disord. 2016;2. [DOI]
    Geoffroy PA, Micoulaud Franchi JA, Lopez R, Schroder CM; membres du consensus Mélatonine SFRMS. The use of melatonin in adult psychiatric disorders: expert recommendations by the French institute of medical research on sleep (SFRMS). Encephale. 2019;45:41323. [DOI] [PubMed]
    Lai MC, Lombardo MV, Baron-Cohen S. Autism. Lancet. 2014;383:896910. [DOI]
    Cortese S, Lecendreux M, Mouren MC, Konofal E. ADHD and insomnia. J Am Acad Child Adolesc Psychiatry. 2006;45:3845. [DOI] [PubMed]
    Karatsoreos IN. Links between circadian rhythms and psychiatric disease. Front Behav Neurosci. 2014;8:162. [DOI] [PubMed] [PMC]
    Charrier A, Olliac B, Roubertoux P, Tordjman S. Clock genes and altered sleep-wake rhythms: their role in the development of psychiatric disorders. Int J Mol Sci. 2017;18:938. [DOI] [PubMed] [PMC]
    Schuch JB, Genro JP, Bastos CR, Ghisleni G, Tovo-Rodrigues L. The role of CLOCK gene in psychiatric disorders: evidence from human and animal research. Am J Med Genet B Neuropsychiatr Genet. 2018;177:18198. [DOI] [PubMed]
    Walker WH 2nd, Walton JC, DeVries AC, Nelson RJ. Circadian rhythm disruption and mental health. Transl Psychiatry. 2020;10:28. [DOI] [PubMed] [PMC]
    von Schantz M, Leocadio-Miguel MA, McCarthy MJ, Papiol S, Landgraf D. Genomic perspectives on the circadian clock hypothesis of psychiatric disorders. Adv Genet. 2021;107:15391. [DOI] [PubMed]
    Sack RL. Clinical practice. Jet lag. N Engl J Med. 2010;362:4407. [DOI] [PubMed]
    Fève-Montange M, Van Cauter E, Refetoff S, Désir D, Tourniaire J, Copinschi G. Effects of “jet lag” on hormonal patterns. II. Adaptation of melatonin circadian periodicity. J Clin Endocrinol Metab. 1981;52:6429. [DOI] [PubMed]
    Belmaker RH, Agam G. Major depressive disorder. N Engl J Med. 2008;358:5568. [DOI] [PubMed]
    Vadnie CA, McClung CA. Circadian rhythm disturbances in mood disorders: insights into the role of the suprachiasmatic nucleus. Neural Plast. 2017;2017:1504507. [DOI] [PubMed] [PMC]
    Li JZ, Bunney BG, Meng F, Hagenauer MH, Walsh DM, Vawter MP, et al. Circadian patterns of gene expression in the human brain and disruption in major depressive disorder. Proc Natl Acad Sci U S A. 2013;110:99505. [DOI] [PubMed] [PMC]
    Tataroğlu O, Aksoy A, Yilmaz A, Canbeyli R. Effect of lesioning the suprachiasmatic nuclei on behavioral despair in rats. Brain Res. 2004;1001:11824. [DOI] [PubMed]
    Ben-Hamo M, Larson TA, Duge LS, Sikkema C, Wilkinson CW, de la Iglesia HO, et al. Circadian forced desynchrony of the master clock leads to phenotypic manifestation of depression in rats. eNeuro. 2017;3:ENEURO.023716.2016. [DOI] [PubMed] [PMC]
    Landgraf D, Long JE, Proulx CD, Barandas R, Malinow R, Welsh DK. Genetic disruption of circadian rhythms in the suprachiasmatic nucleus causes helplessness, behavioral despair, and anxiety-like behavior in mice. Biol Psychiatry. 2016;80:82735. [DOI] [PubMed] [PMC]
    Kalmbach DA, Pillai V, Cheng P, Arnedt JT, Drake CL. Shift work disorder, depression, and anxiety in the transition to rotating shifts: the role of sleep reactivity. Sleep Med. 2015;16:15328. [DOI] [PubMed] [PMC]
    Flo E, Pallesen S, Magerøy N, Moen BE, Grønli J, Hilde Nordhus I, et al. Shift work disorder in nurses--assessment, prevalence and related health problems. PLoS One. 2012;7:e33981. [DOI] [PubMed] [PMC]
    Roybal K, Theobold D, Graham A, DiNieri JA, Russo SJ, Krishnan V, et al. Mania-like behavior induced by disruption of CLOCK. Proc Natl Acad Sci U S A. 2007;104:640611. [DOI] [PubMed] [PMC]
    Spencer S, Falcon E, Kumar J, Krishnan V, Mukherjee S, Birnbaum SG, et al. Circadian genes Period 1 and Period 2 in the nucleus accumbens regulate anxiety-related behavior. Eur J Neurosci. 2013;37:24250. [DOI] [PubMed] [PMC]
    Melo MCA, Abreu RLC, Linhares Neto VB, de Bruin PFC, de Bruin VMS. Chronotype and circadian rhythm in bipolar disorder: a systematic review. Sleep Med Rev. 2017;34:4658. [DOI] [PubMed]
    Katz G, Knobler HY, Laibel Z, Strauss Z, Durst R. Time zone change and major psychiatric morbidity: the results of a 6-year study in Jerusalem. Compr Psychiatry. 2002;43:3740. [DOI] [PubMed]
    Le-Niculescu H, Patel SD, Bhat M, Kuczenski R, Faraone SV, Tsuang MT, et al. Convergent functional genomics of genome-wide association data for bipolar disorder: comprehensive identification of candidate genes, pathways and mechanisms. Am J Med Genet B Neuropsychiatr Genet. 2009;150B:15581. [DOI] [PubMed]
    Bellivier F, Geoffroy PA, Etain B, Scott J. Sleep- and circadian rhythm-associated pathways as therapeutic targets in bipolar disorder. Expert Opin Ther Targets. 2015;19:74763. [DOI] [PubMed]
    Bromundt V, Köster M, Georgiev-Kill A, Opwis K, Wirz-Justice A, Stoppe G, et al. Sleep-wake cycles and cognitive functioning in schizophrenia. Br J Psychiatry. 2011;198:26976. [DOI] [PubMed]
    Benson KL. Sleep in schizophrenia: pathology and treatment. Sleep Med Clin. 2015;10:4955. [DOI] [PubMed]
    Seney ML, Cahill K, Enwright JF 3rd, Logan RW, Huo Z, Zong W, et al. Diurnal rhythms in gene expression in the prefrontal cortex in schizophrenia. Nat Commun. 2019;10:3355. [DOI] [PubMed] [PMC]
    Coulon N, Brailly-Tabard S, Walter M, Tordjman S. Altered circadian patterns of salivary cortisol in individuals with schizophrenia: a critical literature review. J Physiol Paris. 2016;110:43947. [DOI] [PubMed]
    Ho RT, Fong TC, Wan AH, Au-Yeung FS, Chen EY, Spiegel D. Associations between diurnal cortisol patterns and lifestyle factors, psychotic symptoms, and neurological deficits: a longitudinal study on patients with chronic schizophrenia. J Psychiatr Res. 2016;81:1622. [DOI] [PubMed]
    Monteleone P, Maj M, Fusco M, Kemali D, Reiter RJ. Depressed nocturnal plasma melatonin levels in drug-free paranoid schizophrenics. Schizophr Res. 1992;7:7784. [DOI] [PubMed]
    Mansour HA, Talkowski ME, Wood J, Chowdari KV, McClain L, Prasad K, et al. Association study of 21 circadian genes with bipolar I disorder, schizoaffective disorder, and schizophrenia. Bipolar Disord. 2009;11:70110. [DOI] [PubMed] [PMC]
    Johansson AS, Owe-Larsson B, Hetta J, Lundkvist GB. Altered circadian clock gene expression in patients with schizophrenia. Schizophr Res. 2016;174:1723. [DOI] [PubMed]
    Hsu CP, Oka S, Shao D, Hariharan N, Sadoshima J. Nicotinamide phosphoribosyltransferase regulates cell survival through NAD+ synthesis in cardiac myocytes. Circ Res. 2009;105:48191. [DOI] [PubMed] [PMC]
    Maury E, Ramsey KM, Bass J. Circadian rhythms and metabolic syndrome: from experimental genetics to human disease. Circ Res. 2010;106:44762. [DOI] [PubMed] [PMC]
    Dyar KA, Ciciliot S, Wright LE, Biensø RS, Tagliazucchi GM, Patel VR, et al. Muscle insulin sensitivity and glucose metabolism are controlled by the intrinsic muscle clock. Mol Metab. 2013;3:2941. Erratum in: Mol Metab. 2014;3:857. [DOI] [PubMed] [PMC]
    Huang PL. A comprehensive definition for metabolic syndrome. Dis Model Mech. 2009;2:2317. [DOI] [PubMed] [PMC]
    Maury E. Off the clock: from circadian disruption to metabolic disease. Int J Mol Sci. 2019;20:1597. [DOI] [PubMed] [PMC]
    Karatsoreos IN. The relationship between circadian disruption and the development of metabolic syndrome and type 2 diabetes. Chronol Physiol Ther. 2014;4:137. [DOI]
    Serin Y, Acar Tek N. Effect of circadian rhythm on metabolic processes and the regulation of energy balance. Ann Nutr Metab. 2019;74:32230. [DOI] [PubMed]
    Gan Y, Yang C, Tong X, Sun H, Cong Y, Yin X, et al. Shift work and diabetes mellitus: a meta-analysis of observational studies. Occup Environ Med. 2015;72:728. [DOI] [PubMed]
    Shamsi NA, Salkeld MD, Rattanatray L, Voultsios A, Varcoe TJ, Boden MJ, et al. Metabolic consequences of timed feeding in mice. Physiol Behav. 2014;128:188201. [DOI] [PubMed]
    Valenzuela FJ, Vera J, Venegas C, Muñoz S, Oyarce S, Muñoz K, et al. Evidences of polymorphism associated with circadian system and risk of pathologies: a review of the literature. Int J Endocrinol. 2016;2016:2746909. [DOI] [PubMed] [PMC]
    Shostak A, Meyer-Kovac J, Oster H. Circadian regulation of lipid mobilization in white adipose tissues. Diabetes. 2013;62:2195203. [DOI] [PubMed] [PMC]
    Gabriel BM, Zierath JR. Circadian rhythms and exercise - re-setting the clock in metabolic disease. Nat Rev Endocrinol. 2019;15:197206. [DOI] [PubMed]
    Gälman C, Angelin B, Rudling M. Bile acid synthesis in humans has a rapid diurnal variation that is asynchronous with cholesterol synthesis. Gastroenterology. 2005;129:144553. [DOI] [PubMed]
    Duez H, van der Veen JN, Duhem C, Pourcet B, Touvier T, Fontaine C, et al. Regulation of bile acid synthesis by the nuclear receptor Rev-erbα. Gastroenterology. 2008;135:68998. [DOI] [PubMed]
    Rutter J, Reick M, Wu LC, McKnight SL. Regulation of clock and NPAS2 DNA binding by the redox state of NAD cofactors. Science. 2001;293:5104. [DOI] [PubMed]
    Yang SC, Shieh KR. Implications of circadian rhythms on metabolic disorders. Tzu Chi Med J. 2009;21:2858. [DOI]
    Jacobi D, Liu S, Burkewitz K, Kory N, Knudsen NH, Alexander RK, et al. Hepatic Bmal1 regulates rhythmic mitochondrial dynamics and promotes metabolic fitness. Cell Metab. 2015;22:70920. [DOI] [PubMed] [PMC]
    Yoshizaki T, Milne JC, Imamura T, Schenk S, Sonoda N, Babendure JL, et al. SIRT1 exerts anti-inflammatory effects and improves insulin sensitivity in adipocytes. Mol Cell Biol. 2009;29:136374. [DOI] [PubMed] [PMC]
    Kohsaka A, Laposky AD, Ramsey KM, Estrada C, Joshu C, Kobayashi Y, et al. High-fat diet disrupts behavioral and molecular circadian rhythms in mice. Cell Metab. 2007;6:41421. [DOI] [PubMed]
    Paschos GK, Ibrahim S, Song WL, Kunieda T, Grant G, Reyes TM, et al. Obesity in mice with adipocyte-specific deletion of clock component Arntl. Nat Med. 2012;18:176877. [DOI] [PubMed] [PMC]
    Xu K, Zheng X, Sehgal A. Regulation of feeding and metabolism by neuronal and peripheral clocks in Drosophila. Cell Metab. 2008;8:289300. [DOI] [PubMed] [PMC]
    Ahima RS, Prabakaran D, Flier JS. Postnatal leptin surge and regulation of circadian rhythm of leptin by feeding. Implications for energy homeostasis and neuroendocrine function. J Clin Invest. 1998;101:10207. [DOI] [PubMed] [PMC]
    Mingrone G, Manco M, Granato L, Calvani M, Scarfone A, Mora EV, et al. Leptin pulsatility in formerly obese women. FASEB J. 2005;19:13802. [DOI] [PubMed]
    Kahn BB, Alquier T, Carling D, Hardie DG. AMP-activated protein kinase: ancient energy gauge provides clues to modern understanding of metabolism. Cell Metab. 2005;1:1525. [DOI] [PubMed]
    Hotamisligil GS. Inflammation and endoplasmic reticulum stress in obesity and diabetes. Int J Obes (Lond). 2008;32 Suppl 7:S524. [DOI] [PubMed] [PMC]
    Loboda A, Kraft WK, Fine B, Joseph J, Nebozhyn M, Zhang C, et al. Diurnal variation of the human adipose transcriptome and the link to metabolic disease. BMC Med Genomics. 2009;2:7. [DOI] [PubMed] [PMC]
    Kettner NM, Mayo SA, Hua J, Lee C, Moore DD, Fu L. Circadian dysfunction induces leptin resistance in mice. Cell Metab. 2015;22:44859. [DOI] [PubMed] [PMC]
    Stenvers DJ, Scheer FAJL, Schrauwen P, la Fleur SE, Kalsbeek A. Circadian clocks and insulin resistance. Nat Rev Endocrinol. 2019;15:7589. [DOI] [PubMed]
    Turek FW, Joshu C, Kohsaka A, Lin E, Ivanova G, McDearmon E, et al. Obesity and metabolic syndrome in circadian Clock mutant mice. Science. 2005;308:10435. [DOI] [PubMed] [PMC]
    Marcheva B, Ramsey KM, Buhr ED, Kobayashi Y, Su H, Ko CH, et al. Disruption of the clock components CLOCK and BMAL1 leads to hypoinsulinaemia and diabetes. Nature. 2010;466:62731. [DOI] [PubMed] [PMC]
    Dupuis J, Langenberg C, Prokopenko I, Saxena R, Soranzo N, Jackson AU, et al.; Anders Hamsten on behalf of Procardis Consortium; MAGIC investigators; Bergman RN, Buchanan TA, Collins FS, Mohlke KL, Tuomilehto J, Valle TT, et al. New genetic loci implicated in fasting glucose homeostasis and their impact on type 2 diabetes risk. Nat Genet. 2010;42:10516. Erratum in: Nat Genet. 2010;42:464. [DOI] [PubMed] [PMC]
    Barker A, Sharp SJ, Timpson NJ, Bouatia-Naji N, Warrington NM, Kanoni S, et al. Association of genetic Loci with glucose levels in childhood and adolescence: a meta-analysis of over 6,000 children. Diabetes. 2011;60:180512. [DOI] [PubMed] [PMC]
    Perelis M, Marcheva B, Ramsey KM, Schipma MJ, Hutchison AL, Taguchi A, et al. Pancreatic β cell enhancers regulate rhythmic transcription of genes controlling insulin secretion. Science. 2015;350:aac4250. [DOI] [PubMed] [PMC]
    Harada N, Inagaki N. Role of clock genes in insulin secretion. J Diabetes Investig. 2016;7:8223. [DOI] [PubMed] [PMC]
    Revollo JR, Körner A, Mills KF, Satoh A, Wang T, Garten A, et al. Nampt/PBEF/visfatin regulates insulin secretion in beta cells as a systemic NAD biosynthetic enzyme. Cell Metab. 2007;6:36375. [DOI] [PubMed] [PMC]
    Moynihan KA, Imai S. Sirt1 as a key regulator orchestrating the response to caloric restriction. Drug Discov Today Dis Mech. 2006;3:117. [DOI]
    Alamoudi AA. Why do cancer cells break from host circadian rhythm? Insights from unicellular organisms. Bioessays. 2021;43:e2000205. [DOI] [PubMed]
    IARC Monographs Vol 124 group. Carcinogenicity of night shift work. Lancet Oncol. 2019;20:10589. [DOI] [PubMed]
    Ijaz S, Verbeek J, Seidler A, Lindbohm ML, Ojajärvi A, Orsini N, et al. Night-shift work and breast cancer--a systematic review and meta-analysis. Scand J Work Environ Health. 2013;39:43147. [DOI] [PubMed]
    Schernhammer ES, Laden F, Speizer FE, Willett WC, Hunter DJ, Kawachi I, et al. Night-shift work and risk of colorectal cancer in the nurses’ health study. J Natl Cancer Inst. 2003;95:8258. [DOI] [PubMed]
    Wegrzyn LR, Tamimi RM, Rosner BA, Brown SB, Stevens RG, Eliassen AH, et al. Rotating night-shift work and the risk of breast cancer in the nurses’ health studies. Am J Epidemiol. 2017;186:53240. [DOI] [PubMed] [PMC]
    Sulli G, Lam MTY, Panda S. Interplay between circadian clock and cancer: new frontiers for cancer treatment. Trends Cancer. 2019;5:47594. [DOI] [PubMed] [PMC]
    Fu L, Kettner NM. The circadian clock in cancer development and therapy. Prog Mol Biol Transl Sci. 2013;119:22182. [DOI] [PubMed] [PMC]
    Chen ST, Choo KB, Hou MF, Yeh KT, Kuo SJ, Chang JG. Deregulated expression of the PER1, PER2 and PER3 genes in breast cancers. Carcinogenesis. 2005;26:12416. [DOI] [PubMed]
    Winter SL, Bosnoyan-Collins L, Pinnaduwage D, Andrulis IL. Expression of the circadian clock genes Per1 and Per2 in sporadic and familial breast tumors. Neoplasia. 2007;9:797800. [DOI] [PubMed] [PMC]
    Jung-Hynes B, Huang W, Reiter RJ, Ahmad N. Melatonin resynchronizes dysregulated circadian rhythm circuitry in human prostate cancer cells. J Pineal Res. 2010;49:608. [DOI] [PubMed] [PMC]
    Cao Q, Gery S, Dashti A, Yin D, Zhou Y, Gu J, et al. A role for the clock gene per1 in prostate cancer. Cancer Res. 2009;69:761925. [DOI] [PubMed] [PMC]
    Xiong H, Yang Y, Yang K, Zhao D, Tang H, Ran X. Loss of the clock gene PER2 is associated with cancer development and altered expression of important tumor-related genes in oral cancer. Int J Oncol. 2018;52:27987. [DOI]
    Pogue-Geile KL, Lyons-Weiler J, Whitcomb DC. Molecular overlap of fly circadian rhythms and human pancreatic cancer. Cancer Lett. 2006;243:557. [DOI] [PubMed]
    Relles D, Sendecki J, Chipitsyna G, Hyslop T, Yeo CJ, Arafat HA. Circadian gene expression and clinicopathologic correlates in pancreatic cancer. J Gastrointest Surg. 2013;17:44350. [DOI] [PubMed]
    Lin YM, Chang JH, Yeh KT, Yang MY, Liu TC, Lin SF, et al. Disturbance of circadian gene expression in hepatocellular carcinoma. Mol Carcinog. 2008;47:92533. [DOI] [PubMed]
    Mostafaie N, Kállay E, Sauerzapf E, Bonner E, Kriwanek S, Cross HS, et al. Correlated downregulation of estrogen receptor beta and the circadian clock gene Per1 in human colorectal cancer. Mol Carcinog. 2009;48:6427. [DOI] [PubMed]
    Gery S, Komatsu N, Kawamata N, Miller CW, Desmond J, Virk RK, et al. Epigenetic silencing of the candidate tumor suppressor gene Per1 in non-small cell lung cancer. Clin Cancer Res. 2007;13:1399404. [DOI] [PubMed]
    Mazzoccoli G, Piepoli A, Carella M, Panza A, Pazienza V, Benegiamo G, et al. Altered expression of the clock gene machinery in kidney cancer patients. Biomed Pharmacother. 2012;66:1759. [DOI] [PubMed]
    Litlekalsoy J, Rostad K, Kalland KH, Hostmark JG, Laerum OD. Expression of circadian clock genes and proteins in urothelial cancer is related to cancer-associated genes. BMC Cancer. 2016;16:549. [DOI] [PubMed] [PMC]
    Yang MY, Chang JG, Lin PM, Tang KP, Chen YH, Lin HY, et al. Downregulation of circadian clock genes in chronic myeloid leukemia: alternative methylation pattern of hPER3. Cancer Sci. 2006;97:1298307. [DOI] [PubMed]
    Yang MY, Yang WC, Lin PM, Hsu JF, Hsiao HH, Liu YC, et al. Altered expression of circadian clock genes in human chronic myeloid leukemia. J Biol Rhythms. 2011;26:13648. [DOI] [PubMed]
    Xia HC, Niu ZF, Ma H, Cao SZ, Hao SC, Liu ZT, et al. Deregulated expression of the Per1 and Per2 in human gliomas. Can J Neurol Sci. 2010;37:36570. [DOI] [PubMed]
    Hsu CM, Lin SF, Lu CT, Lin PM, Yang MY. Altered expression of circadian clock genes in head and neck squamous cell carcinoma. Tumour Biol. 2012;33:14955. [DOI] [PubMed]
    Gery S, Komatsu N, Baldjyan L, Yu A, Koo D, Koeffler HP. The circadian gene per1 plays an important role in cell growth and DNA damage control in human cancer cells. Mol Cell. 2006;22:37582. [DOI] [PubMed]
    Blakeman V, Williams JL, Meng QJ, Streuli CH. Circadian clocks and breast cancer. Breast Cancer Res. 2016;18:89. [DOI] [PubMed] [PMC]
    Hoffman AE, Zheng T, Stevens RG, Ba Y, Zhang Y, Leaderer D, et al. Clock-cancer connection in non-Hodgkin’s lymphoma: a genetic association study and pathway analysis of the circadian gene cryptochrome 2. Cancer Res. 2009;69:360513. [DOI] [PubMed] [PMC]
    Zhu Y, Leaderer D, Guss C, Brown HN, Zhang Y, Boyle P, et al. Ala394Thr polymorphism in the clock gene NPAS2: a circadian modifier for the risk of non-Hodgkin’s lymphoma. Int J Cancer. 2007;120:4325. [DOI] [PubMed] [PMC]
    Kettner NM, Voicu H, Finegold MJ, Coarfa C, Sreekumar A, Putluri N, et al. Circadian homeostasis of liver metabolism suppresses hepatocarcinogenesis. Cancer Cell. 2016;30:90924. [DOI] [PubMed] [PMC]
    Zhang Z, Ma F, Zhou F, Chen Y, Wang X, Zhang H, et al. Functional polymorphisms of circadian negative feedback regulation genes are associated with clinical outcome in hepatocellular carcinoma patients receiving radical resection. Med Oncol. 2014;31:179. [DOI] [PubMed]
    Michelotti GA, Machado MV, Diehl AM. NAFLD, NASH and liver cancer. Nat Rev Gastroenterol Hepatol. 2013;10:65665. [DOI] [PubMed]
    Bhatti P, Cushing-Haugen KL, Wicklund KG, Doherty JA, Rossing MA. Nightshift work and risk of ovarian cancer. Occup Environ Med. 2013;70:2317. [DOI] [PubMed] [PMC]
    Jim HS, Lin HY, Tyrer JP, Lawrenson K, Dennis J, Chornokur G, et al.; Georgia Chenevix-Trench on behalf of the AOCS management group 95,96; Schildkraut JM, Kelemen LE, Ramus SJ, Monteiro AN, Goode EL, Narod SA, et al. Common genetic variation in circadian rhythm genes and risk of epithelial ovarian cancer (EOC). J Genet Genome Res. 2015;2:017. [DOI] [PubMed] [PMC]
    Alvarez JD, Hansen A, Ord T, Bebas P, Chappell PE, Giebultowicz JM, et al. The circadian clock protein BMAL1 is necessary for fertility and proper testosterone production in mice. J Biol Rhythms. 2008;23:2636. [DOI] [PubMed] [PMC]
    Hodžić A, Ristanović M, Zorn B, Tulić C, Maver A, Novaković I, et al. Genetic variation in circadian rhythm genes CLOCK and ARNTL as risk factor for male infertility. PLoS One. 2013;8:e59220. [DOI] [PubMed] [PMC]
    Boden MJ, Varcoe TJ, Kennaway DJ. Circadian regulation of reproduction: from gamete to offspring. Prog Biophys Mol Biol. 2013;113:38797. [DOI] [PubMed]
    Wright KP Jr, Bogan RK, Wyatt JK. Shift work and the assessment and management of shift work disorder (SWD). Sleep Med Rev. 2013;17:4154. [DOI] [PubMed]
    Rajaratnam SM, Howard ME, Grunstein RR. Sleep loss and circadian disruption in shift work: health burden and management. Med J Aust. 2013;199:S115. [DOI]
    Wilsmore BR, Grunstein RR, Fransen M, Woodward M, Norton R, Ameratunga S. Sleep habits, insomnia, and daytime sleepiness in a large and healthy community-based sample of New Zealanders. J Clin Sleep Med. 2013;9:55966. [DOI] [PubMed] [PMC]
    Johnsson A. Light, circadian and circannual rhythms. In: Bjertness E, editor. The Norwegian academy of science and lett. Solar radiation and human health; 2008. pp. 5775.
    Knutsson A. Health disorders of shift workers. Occup Med (Lond). 2003;53:1038. [DOI] [PubMed]
    Folkard S, Tucker P. Shift work, safety and productivity. Occup Med (Lond). 2003;53:95101. [DOI] [PubMed]
    Barger LK, Cade BE, Ayas NT, Cronin JW, Rosner B, Speizer FE, et al.; Harvard Work Hours, Health, and Safety Group. Extended work shifts and the risk of motor vehicle crashes among interns. N Engl J Med. 2005;352:12534. [DOI] [PubMed]
    Rajaratnam SM, Arendt J. Health in a 24-h society. Lancet. 2001;358:9991005. [DOI] [PubMed]
    Bobko N, Karpenko A, Gerasimov A, Chernyuk V. The mental performance of shift workers in nuclear and heat power plants of Ukraine. Int J Ind Ergon. 1998;21:33340. [DOI]
    Biological rhythms: implications for the worker [Internet]. Washington, DC: U.S. Government Printing Office; c1991 [cited 2020 Jan 27]. Available from: https://www.princeton.edu/~ota/disk1/1991/9108/9108.PDF
    Arendt J, Stone B, Skene DJ. Chapter 55 - Jet lag and sleep disruption. In: Kryger M, Roth T, Dement WC, editor. Principles and practice of sleep medicine. 4th. Philadelphia: Elsevier Saunders; 2005. pp. 65972. [DOI] [PubMed]
    Wong PM, Hasler BP, Kamarck TW, Muldoon MF, Manuck SB. Social jetlag, chronotype, and cardiometabolic risk. J Clin Endocrinol Metab. 2015;100:461220. [DOI] [PubMed] [PMC]
    Hammer GP, Auvinen A, De Stavola BL, Grajewski B, Gundestrup M, Haldorsen T, et al. Mortality from cancer and other causes in commercial airline crews: a joint analysis of cohorts from 10 countries. Occup Environ Med. 2014;71:31322. [DOI] [PubMed]
    Waterhouse J, Reilly T, Atkinson G, Edwards B. Jet lag: trends and coping strategies. Lancet. 2007;369:111729. [DOI] [PubMed]
    Skene DJ. Optimization of light and melatonin to phase-shift human circadian rhythms. J Neuroendocrinol. 2003;15:43841. [DOI] [PubMed]
    Cho K, Ennaceur A, Cole JC, Suh CK. Chronic jet lag produces cognitive deficits. J Neurosci. 2000;20:RC66. [DOI] [PMC]
    Gibson EM, Wang C, Tjho S, Khattar N, Kriegsfeld LJ. Experimental ‘jet lag’ inhibits adult neurogenesis and produces long-term cognitive deficits in female hamsters. PLoS One. 2010;5:e15267. [DOI] [PubMed] [PMC]