You are on page 1of 17

Reviews

Rhythms of life: circadian disruption


and brain disorders across the lifespan
Ryan W. Logan and Colleen A. McClung*
Abstract | Many processes in the human body — including brain function — are regulated over
the 24-hour cycle, and there are strong associations between disrupted circadian rhythms (for
example, sleep–wake cycles) and disorders of the CNS. Brain disorders such as autism, depression
and Parkinson disease typically develop at certain stages of life, and circadian rhythms are
important during each stage of life for the regulation of processes that may influence the
development of these disorders. Here, we describe circadian disruptions observed in various
brain disorders throughout the human lifespan and highlight emerging evidence suggesting
these disruptions affect the brain. Currently , much of the evidence linking brain disorders and
circadian dysfunction is correlational, and so whether and what kind of causal relationships might
exist are unclear. We therefore identify remaining questions that may direct future research
towards a better understanding of the links between circadian disruption and CNS disorders.

Circadian rhythms are near-24-hour oscillations found adolescence, and this shift in timing may be conserved
in essentially every physiological process in the human among rodents and non-​human primates (NHPs)10.
brain and body1. The suprachiasmatic nucleus (SCN) In older adults, rhythms often return to being substan-
in the hypothalamus serves as the master pacemaker tially earlier, and this shift may be accompanied by a
that sets the timing of rhythms by regulating neuronal weakening of circadian rhythms11,12.
activity, body temperature and hormonal signals2 (Fig. 1). Circadian dysfunction is observed in several brain
In individual cells, molecular rhythms are generated disorders, typically emerging at different stages of life.
by a transcriptional–translational feedback loop involv- Disruption of circadian rhythms is associated with
ing core transcriptional activators — circadian locomo- higher risk of brain disorders. For example, chronic
tor output cycles kaput (CLOCK), the closely related shift-​workers (workers with rotating schedules or con-
neuronal PAS domain protein 2 (NPAS2) and brain and sistent night shifts) are vulnerable to various diseases13,14,
muscle ARNT-​like protein 1 (BMAL1) — that regulate including psychiatric disorders such as depression15–19.
the expression of many genes, including those encod- Much of the currently available evidence linking brain
ing period (PER) and cryptochrome (CRY)1, which, disorders to circadian dysfunction is correlational and
once translated, inhibit their own transcription. Many may not be surprising, given the importance of circa-
other proteins, including various kinases, phosphatases dian rhythms for brain function20. Identifying causal
and other transcriptional cofactors, regulate this core relationships between rhythm dysfunction and psychi-
molecular clock (Fig. 1). atric disorders may have important implications for the
Circadian rhythms are set by both genetic and envi- treatment of disorders.
ronmental factors. Most people have sleep–wake and In psychiatry, neurology and behavioural medicine,
activity rhythms that, in the absence of environmental novel circadian-​based interventions are rare and often
cues, are slightly longer than 24 hours3,4, but the length overlooked. Even in cases in which circadian disruption
of this period can be affected by circadian gene variants5. is not a primary cause of a disorder, stabilizing sleep–
Environmental factors such as light exposure, social wake patterns using behavioural, environmental or
cues, meal times and work schedules also influence the pharmacological tools might, for example, ease symp-
University of Pittsburgh period, phase and amplitude of these rhythms6,7. toms. Understanding the extent to which circadian
School of Medicine, Circadian rhythms emerge during early infancy rhythms are involved in normal functioning and patho-
Department of Psychiatry,
Pittsburgh, PA, USA.
but undergo various changes through the lifespan and logical processes requires systems-​biology approaches.
with ageing (Fig. 2). In general, the timing of sleep onset Such approaches have proved valuable for investigating
*e-​mail: mcclungca@
upmc.edu and waking and other biological rhythms (for example, the complex relationships between rhythms and meta-
https://doi.org/10.1038/ fluctuations in melatonin levels) is earlier relative to bolic disorders21 and continue to provide strong trans-
s41583-018-0088-y adults during early childhood8,9 and shifts later during lational potential for treating metabolic disorders and

NAtuRe Reviews | NEuRosCiEnCE volume 20 | JANUARY 2019 | 49


Reviews

a Circadian timing system b Mammalian molecular clock


Brain targets Auxiliary loop Clock-regulated
SCN SCN ipRGCs pathways:
• ArcN • RMTg • SCN • Neurogenesis
• DmH • VTA • IGL • Cortical development
Light ipRGCs • PVN • NAc • LHb HDAC3 • Neurotransmission
NCOR REV • Synaptic homeostasis
• PVT • Sptm • LH ROR
Retino- • Hormones • mPOA • DR • VLPO • Inflammation
hypothalamic • SNS and PNS • LHb • LC • MA • Neuronal metabolism
tract • Body temperature • SPZ RRE Bmal1 • Mitochondrial function
• Redox signalling
• Oxidative stress
Thermogenesis • Blood–brain barrier
Lymph
node Positive loop
Brown

NPAS2
adipose

CLOCK
BMAL1
Lung tissue • Ror
• Reverba
Heart
E-box CCG CCGs
• Per1, Per2
• Cry1, Cry2

Immunity
Feeding schedule Liver
Spleen
Negative loop Tissue-specific CCGs
Pancreas
Adrenal gland
CRY
Kidney Heart
PER

NPAS2
Intestine Metabolism Ac

CLOCK
BMAL1
SIRT1 Brain Liver
Uterus
E-box CCG Nampt
Reproduction
White
adipose
tissue Metabolic loop CCG acrophases and
Fat storage amplitudes vary
between tissues
SIRT1
Exercise Muscle
Bone marrow
Fibroblast NAD+ NAM

NAMPT
Stem cells NMN

Fig. 1 | The circadian timing system. a | The circadian timing system negative-​feedback loop. The auxiliary loop includes the nuclear retinoic
synchronizes clocks across the entire body to adapt and optimize physiology acid receptor-​related orphan receptors (RORα and RORβ) and REV-​ERBs
to changes in our environment. Light is received by specialized melanopsin- (REV-​ERBα and REV-​ERBβ), which are also transcriptionally regulated by
producing photoreceptive retinal ganglion cells (ipRGCs) in the eye. These CLOCK–BMAL1 heterodimers. REV-​ERBα (REV in the figure) and RORα
ipRGCs project through the retinohypothalamic tract to the suprachias­matic repress and activate the transcription of Bmal1, respectively , by inhibiting
nucleus (SCN), among other brain regions. The SCN relays timing information and activating the ROR or REV-​ERB response elements (RREs). CLOCK–
to other areas of the brain via direct projections (dark green boxes) and BMAL1 complexes also control the expression of nicotinamide phosphori-
indirect projections (light green boxes). Humoral signals and the peripheral bosyltransferase (NAMPT), which is the rate-​limiting enzyme of NAD+
nervous system (that is, the sympathetic nervous system (SNS) and biosynthesis from nicotinamide (NAM). NAM is modified by NAMPT to
parasympathetic nervous system (PNS)) convey information from the SCN produce nicotinamide mononucleotide (NMN), which in turn is converted
to orchestrate peripheral clocks. Feeding schedules and exercise can also to NAD+ by several adenyltransferases. Thus, NAMPT oscillations control
entrain central and peripheral clocks. Circadian rhythms are key regulators circadian fluctuations in NAD+ levels, which in turn modulate sirtuin 1 (SIRT1)
of thermogenesis, immune function, metabolism, reproduction and stem activity and signalling. High levels of NAD+ promote SIRT1 activation. SIRT1
cell development. b | The mammalian molecular clock is composed of interacts directly with CLOCK–BMAL1 to deacetylate BMAL1 and inhibit
transcriptional and translational feedback loops that oscillate with a CLOCK-​driven transcription. Between tissues and cell types, CCGs and
near-24-hour cycle. The positive loop is driven by the heterodimerization of other molecular and cellular rhythms may be expressed with different
either circadian locomotor output cycles protein kaput (CLOCK) or neuronal acrophases (phase of peak expression), amplitudes and even periodicities.
PAS domain-​containing protein 2 (NPAS2) with brain and muscle ARNT-​like 1 ArcN, arcuate nucleus; DmH, dorsomedial hypothalamus; DR , dorsal raphe;
(BMAL1) in the nucleus. The resulting heterodimers bind to enhancer IGL , intergeniculate leaflet; LC, locus coeruleus; LH, lateral hypothalamus;
boxes (E-​boxes) in gene promoters to regulate the transcription of clock-​ LHb, lateral habenula; MA , medial amygdala; mPOA , medial preoptic area;
controlled genes (CCGs), including those encoding period (PER) proteins NAc, nucleus accumbens; PVN, paraventricular nucleus of the hypothalamus;
and cryptochrome (CRY) proteins. PER and CRY proteins accumulate in PVT, paraventricular nucleus of the thalamus; RMTg, rostromedial tegmental
the cytoplasm during the circadian cycle, eventually dimerizing and nucleus; Sptm, septum; SPZ, subparaventricular zone; VLPO, ventrolateral
shuttling to the nucleus to inhibit their own transcription, thus closing the preoptic nucleus; VTA , ventral tegmental area.

50 | JANUARY 2019 | volume 20 www.nature.com/nrn


Reviews

Wake Sleep be associated with certain brain disorders that typically


Infancy
emerge during these life periods, such as neurodevelop-
Core body temperature mental, psychiatric or neurodegenerative disorders. As
Melatonin most of the research linking circadian disruption and
Amplitude

Cortisol brain disorders is correlational, we emphasize key find-


ings with a particular focus on mechanism and causality.
In addition, we discuss implications for treatments
and interventions.
12:00 18:00 00:00 06:00 12:00
Childhood Prenatal period and early childhood
Many of our insights into the development of the cir-
cadian system are from laboratory investigations of
Amplitude

rodents and NHPs. By mid-​to-late gestation, rhythms


of metabolic activity and gene expression are evident in the
SCN of rodents and NHPs23–25; however, the functional
importance of early prenatal development of the SCN
12:00 18:00 00:00 06:00 12:00 for fetal rhythms and physiology is less well understood.
Adolescence In humans, melatonin and dopamine receptors appear as
early as gestational week 18 in the fetal SCN26–28, suggest-
ing melatonin and dopamine may serve as the primary
Amplitude

communicators of circadian information for the fetus29,30.


Melatonin from the mother readily passes through the
placenta and the fetal blood–brain barrier and may
thus be the predominant relay of time-​of-day infor-
12:00 18:00 00:00 06:00 12:00
mation to the fetus, particularly during the transition
Adulthood from dusk to night, as circulating melatonin rises31–34.
Circadian rhythms of melatonin along with cortico­
steroid are detectable in the umbilical artery and vein,
Amplitude

along with fetal circulation35,36. Fetal rhythms are also


entrained by maternal core body temperature, feeding
and hormone release27 (Fig. 3).
12:00 18:00 00:00 06:00 12:00
Prenatal and maternal rhythms. Epidemiological
Later life studies report increased risks of preterm delivery, low
birthweights and miscarriage in pregnant women who
Amplitude

are chronic shift-​workers37–40. Abnormal sleep, feeding


and work schedules might disrupt maternal rhythms,
such as melatonin fluctuations, and thus desynchronize
the maternal SCN from peripheral oscillators and lead to
deleterious effects of shift-​work on the fetus. Pregnant
12:00 18:00 00:00 06:00 12:00
rats exposed to simulated shift-​work (that is, complete
Time
reversal of the light–dark cycle every 3–4 days for sev-
Fig. 2 | Rhythms across the lifespan. Schematic of circadian rhythm changes from eral weeks) showed substantially reduced weight gain
infancy , adolescence, adulthood and older age. During infancy , sleep–wake rhythms during early pregnancy and reduced fat-​pad and liver
are ultradian and consolidate during the first year of development. From childhood to weights, and also exhibited lower-​amplitude rhythms
adolescence, there is a marked shift from an early to a late chronotype, which of corticosterone, glucose, insulin and leptin levels41.
subsequently becomes earlier during adulthood, with shorter sleep durations through
The same manipulation also affected metabolic and
adulthood. Rhythms undergo a gradual loss of amplitude with ageing. Temperature
circadian gene expression rhythms in the fetal liver, but
rhythms peak during childhood, and amplitudes steadily reduce during ageing.
Melatonin rhythms are delayed during adolescence, with overall levels peaking during did not affect placental or fetal growth41. Exposing preg-
childhood and considerably decreasing during ageing. Similarly , rodent studies have nant NHPs to constant light suppresses the emergence
demonstrated that suprachiasmatic nucleus (SCN) activity rhythms gradually decline of melatonin and body temperature rhythms in their
with ageing (not shown). Cortisol rhythms peak earlier in the morning during childhood offspring following birth42–44, suggesting that maternal
and, with age, gradually widen and reduce in overall amplitude. The amplitude of rhythms are important for the early brain development
rhythmic gene expression in the brain and other tissues is reduced during ageing, of fetal circadian systems45. Moreover, the rhythms of the
affecting tissue homeostasis and function (not shown). expression of circadian genes (specifically, Bmal1 (also
known as Arntl) and Per2) and genes encoding NMDA
premature ageing through molecular and behavioural receptor subunits (including Grin1b, Grin3a and Grin3b)
therapeutics targeting the circadian system22. In this are suppressed in the hippocampus of adult offspring
Review, we describe the changes in circadian rhythms of rats exposed to constant light during pregnancy46.
that occur with human development into adulthood Alterations in the rhythms of circadian gene expres-
and with ageing and highlight evidence suggesting sion are also associated with impaired spatial memory
that circadian disruptions at different life stages may in these offspring46,47, and these memory deficits can be

NAtuRe Reviews | NEuRosCiEnCE volume 20 | JANUARY 2019 | 51


Reviews

Maternal system Eye


RHT
SCN
Hormonal and ANS Hormonal and
neural signals neural signals
Pineal gland

Temperature and Melatonin Peripheral


feeding rhythms clocks

Mother
• Melatonin
• Corticosterone
• Metabolism
enta
• Activity, e.g. exercise Plac SCN

Adrenal glands ? Pituitary gland


Fetus
• Hormones
• Metabolism
• Sleep–wake cycles • Heart
• Liver
• Respiration

Fetal circadian rhythms


• Tissue homeostasis and neurodevelopment
• Development and consolidation of feeding,
metabolic and sleep–wake rhythms

Fig. 3 | Maternal and fetal rhythms. Fetal tissues such as adrenal glands and the suprachiasmatic nucleus (SCN) may be
entrained by rhythms of maternal feeding schedules, core body temperature and melatonin. Entraining signals, including
melatonin and glucocorticoids, may cross the placental barrier to entrain, modulate or aid the development of fetal
circadian rhythms among brain and peripheral tissues. ANS, autonomic nervous system; RHT, retinohypothalamic tract.

prevented by regularly scheduled supplementation of nocturnal sleep onset is coupled with sunset for the
melatonin to the mother46. Gestational circadian disrup- first few months after birth and then to family bedtime
tion also promotes social avoidance and hyperactivity thereafter, suggesting that the circadian system is ini-
behaviours in pups, even when they are cross-​fostered tially entrained by light but is subsequently entrained
by non-​disrupted mothers48, further highlighting the by social and environmental factors. Most infants,
importance of circadian stability during pregnancy on toddlers (2–3-year-​olds) and young children (4–11-year-​
the mother and offspring. olds) display earlier sleep–wake patterns and a stronger
Peripheral oscillators
Circadian ‘clocks’ in other parts Other studies have shown that even dim light at night morning preference than do adults, according to the
of the brain and in peripheral has physiological and behavioural consequences on par- Children’s Chronotype Questionnaire — even when
organs that are driven by the ents and offspring. For example, chronically exposing considering the high intra-​individual and interindi-
suprachiasmatic nucleus and mice to dim light at night impairs adaptive immune vidual variability of sleep onset, mid-​points and wake
various other nonphotic cues
and stimuli.
function49 and increases depression-​like behaviours50 times measured by parental report questionnaires and
in their offspring. Notably, these effects are evident in actigraphy8,9. However, on non-​scheduled days, children
Actigraphy offspring from parents exposed to dim light at night tend to wake up later, substantially delaying their sleep–
The monitoring of physical before breeding, suggesting that repeated exposure to wake times, indicating that social and environmental
activity patterns (that is, rest–
dim light at night leads to epigenetic modifications that factors, such as schedules imposed by school and par-
activity cycles) over many days,
weeks or months, typically can be passed between generations. Such studies are ents, have a major influence on survey-​based measures
using non-​invasive devices, particularly important because rates of shift-​work and of chronotype9. Importantly, children who consistently
such as wrist actometers. extended work schedules have increased over the past display earlier sleep onsets on scheduled days tend to
decade, as have exposures to prolonged periods of light delay less than do children who have later sleep onsets
Chronotype
The preference for early
at night from work and social demands, and the use of and wake times9.
or late activities (translating electronic devices51. Social environment and scheduling demands by the
to morning or evening parents can interact with endogenous circadian and
chronotypes, respectively) Rhythms in early childhood. Circadian rhythms sleep rhythms, creating a state of misalignment between
during the 24-hour cycle.
gradually emerge between birth and the first several social schedule, endogenous circadian physiology and
Sleep homeostasis months of life. Temperature rhythms appear almost sleep homeostasis. Sleep loss and consistently shorter
A type of homeostasis by immediately after birth in full-​term infants, whereas sleep durations have been linked to inattentiveness, frus-
which prolonged periods of other rhythms, including rest–activity, sleep–wake tration and hyperactivity in 2–5-year-​old children53,54.
wakefulness lead to increases and hormonal cycles, typically develop between 3 and Interestingly, before the age of 3 years, poor sleep quality
in the intensity and duration
of subsequent sleep, in
6 months of age23,52. During the first year of life, sleep– and shorter sleep durations, which could be the result
compensation for a sleep wake rhythms continue to consolidate, coinciding with of inconsistent sleep–wake cycles and less-​robust circa-
deficit or deprivation. increased melatonin secretion at sunset 23. Notably, dian rhythms, are predictive of hyperactivity, cognitive

52 | JANUARY 2019 | volume 20 www.nature.com/nrn


Reviews

deficits and impulsivity at 6 years of age55,56, even in chil- to determine the extent to which different light–dark
dren showing normative sleep in the interim period57. schedules might improve long-​term health outcomes in
Inconsistent sleep–wake cycles during early develop- infants in the NICU.
ment might be contributing to the steadily rising rates
of emotional and behavioural problems in young chil- Neurodevelopmental disorders
dren58–61. Additional — ideally longitudinal — studies Sleep disturbances and circadian disruptions are associ-
from infancy and through early childhood are needed ated with several neurodevelopmental disorders, includ-
to better understand the interplay between environment ing attention-​deficit hyperactivity disorder (ADHD),
and biology in the development of the circadian system and autism spectrum disorders (ASDs), Prader–Willi syn-
the rest of the brain and their potential involvement in drome (PWS) and Smith–Magenis syndrome (SMS).
cognitive, mood and behavioural issues. In support of One of the most common neurodevelopmental disor-
this, a large twin study suggested that shared environ- ders of childhood, ADHD, is characterized by inatten-
ments (rather than genetics) are the primary influencer tiveness, impulsivity and hyperactivity and is associated
of sleep abnormalities and behavioural problems62, with high rates of co-​occurring sleep problems and cir-
although these findings do not preclude the involve- cadian alterations86. Reductions in sleep quality, delays in
ment of genetic or other biological factors that shape circadian phase and evening preference are consistently
early parenting practices. reported in children and adults with ADHD87 and may
At birth, light is the most important stimulus for be correlated with the severity of ADHD symptoms88. In
entraining circadian rhythms. Rodent studies demon- adults with ADHD, a loss of circadian gene expression
strate that melanopsin photoreceptors are present prena- rhythms in the oral mucosa accompanies delays in corti-
tally, are light-​responsive after gestation and undergo sol rhythm and reduced-​amplitude melatonin rhythms89.
functional reorganization and maturation during Advancing melatonin rhythms using bright light therapy
the first few weeks of life63–69. In rats, photic inputs to the improves ADHD symptoms of hyperactivity and impul-
SCN are mature after postnatal day 10 (ref.64), whereas in sivity in adults independent of changes in sleep time,
mice, these pathways reach maturity by postnatal day 14, sleep efficiency, wake time or waking during sleeping87,90.
when the eyes fully open67,69. A few studies have shown Abnormal melatonin rhythms (for example, with
that the eyes of infants who were born prematurely can phase delays and reduced amplitude) are also reported
respond to light as early as gestational week 30, sug- in children with ADHD87. Studies of sleep–wake cycles in
Melanopsin photoreceptors
Opsin class G protein-​coupled gesting that photic input pathways may be functional children with ADHD are often complicated by the effects
receptors expressed in a during the third trimester of development70,71. Studies in of stimulants used to treat the disorder91. Interestingly,
small percentage (~2%) of baboons reveal that metabolic activity and gene expres- modafinil, a stimulant used to treat narcolepsy and shift
photosensitive mammalian sion in the SCN are highly responsive to light inputs in work disorders, may be effective for treating ADHD92
retinal ganglion cells.
full-​term infants and can be entrained by low-​intensity and, along with targeting dopaminergic signalling, may
Bright light therapy light, similar to human infants72–74. also target therapeutic mechanisms distinct from those
A therapy for mood and sleep Regular light–dark schedules and/or low-​intensity targeted by amphetamine and methylphenidate, including
disorders that involves the use lighting, as well as particular wave lengths of light that central histamine and orexin signalling pathways, to
of bright light, typically in the
affect the circadian clock, may be beneficial for the over- improve sleep efficiency and quality93.
morning, to shift and stabilize
endogenous circadian timing. all health of premature infants75–77. In neonatal intensive Studies have also tenuously associated dysfunction
care units (NICUs) worldwide, constant lighting condi- in circadian rhythms with ASD, a neurodevelopmental
Sleep efficiency tions are often used to aid the care and nursing staff to disorder characterized by impairments in social com-
A measure of the integrity of readily respond to potential emergencies and frequent munication, restricted interests and repetitive behav-
sleep architecture and quality
of sleep, consisting of the ratio
health monitoring78. However, rearing premature infants iours. Sleep problems are very common in children with
of total sleep time to the (32 weeks gestational age or older) under regular light– ASD94–96, although revealing any underlying circadian
amount of time spent in bed. dark schedules in the NICU leads to greater and more abnormalities has proved difficult97,98. The most con-
rapid weight gain than rearing infants under constant sistent related finding in prepubertal children, puber-
Narcolepsy
bright light or dim light, effectively shortening their hos- tal adolescents and young adults with ASD is reduced
A chronic neurological disorder
characterized by excessive pital stays76,77. Premature infants kept in regular light– melatonin levels in the evening99–102. A genetic basis for
daytime sleepiness, with many dark conditions also fed orally sooner and spent fewer impaired melatonin synthesis in ASD was proposed fol-
individuals experiencing days on ventilator assistance, but had similar motor lowing the finding that unaffected parents of children
intermittent, uncontrollable coordination compared with infants under continu- with ASD have lower evening melatonin levels and
episodes of cataplexy (a
sudden loss of muscle tone
ous near darkness in the NICU (28–32 weeks of gesta- lower activity of the acetylserotonin methyltransferase
during wakefulness). tion)79–85. Moreover, compared with infants in constant (ASMT) enzyme102, which converts N-​acetylserotonin to
dim light, these infants cry less and are more active in melatonin103. Variants identified in the promoter of the
Amphetamine the daytime81. Regular light–dark schedules also support the ASMT gene were specific to individuals with ASD com-
A potent stimulant drug that
maturation of rest–activity, sleep–wake and melatonin pared with controls and were associated with reduced
can be used in the treatment of
attention-​deficit hyperactivity rhythms earlier in premature infants71,85. Although more ASMT activity and melatonin levels102. In addition to
disorder. controlled comparative studies are necessary, these find- its role in sleep and circadian rhythms, melatonin is
ings strongly suggest that regular light–dark and feeding also a potent antioxidant. Reductions in melatonin
Methylphenidate schedules may substitute, in part, for the early loss of the during early development have been hypothesized to
A stimulant drug primarily
used in the treatment of
influence of maternal rhythms and loss of the constant lead to a build-​up of oxidative stress, which is harm-
attention-​deficit hyperactivity darkness environment in utero in premature infants. ful to the developing nervous system and increases the
disorder. Larger multisite randomized trials will be necessary risk of neurodevelopmental disorders such as ASD104.

NAtuRe Reviews | NEuRosCiEnCE volume 20 | JANUARY 2019 | 53


Reviews

A double-​blind placebo-​controlled study of 125 children RAI1 is implicated in melatonin secretion, and inverted
and adolescents (2–17.5 years of age) supports the use of melatonin rhythms may be responsible for these sleep
melatonin supplementation to treat sleep disturbances disturbances132. Mice with a hetero­zygous deletion of
associated with ASD105 and may lead to other therapeutic Rai1 display a shortened locomotor activity period133,134,
approaches targeting circadian rhythms and sleep issues, although these mice are of mixed genetic background
as well as other symptoms of ASD106. containing C57BL/6, which do not produce pineal mel-
Other neurodevelopmental disorders, such as PWS atonin. Introducing the mutation into another mouse
and SMS, are linked more directly to dysfunctional background capable of generating endogenous mel-
circadian rhythm pathways107,108. PWS is characterized atonin may yield more pronounced effects on sleep
by hypotonia and failure to thrive in infancy and early and circadian rhythm phenotypes relevant for SMS.
toddlerhood followed by hyperphagia and childhood In human cells from patients with SMS and the mouse
obesity, intellectual impairment, hypogonadism and brain, RAI1 activates the transcription of CLOCK via
obsessive–compulsive-​related behaviours109, and it is direct binding to enhancer elements within intron 1 of
due to a paternal deficiency of the 15q11–13 locus110. the CLOCK gene135, demonstrating a possible mecha-
PWS is also accompanied by shorter sleep duration and nism by which loss of RAI1 may result in circadian
excessive daytime sleepiness71,111,112. The 15q11–13 locus dysfunction at the molecular and cellular level. Further
contains several clusters of genes encoding small nucle- study is required to definitively link molecular clock dis-
olar RNAs (snoRNAs), including SNORD116, which ruption observed in animal models of SMS and patients
is expressed in the brain of mice and humans113–115. In with the disorder to the reversal in melatonin secretion.
mice, paternal deletion of Snord116 increases the expres- Circadian, sleep and behavioural disturbances are the
sion of several genes specifically during the light phase, most severe and difficult to treat for individuals with
including Ube3a, which encodes ubiquitin ligase E3A SMS. Attenuating the rise of melatonin during the morn-
(UBE3A)116. UBE3A targets the circadian transcription ing via pharmacological antagonism of β1-adrenergic
factor BMAL1 for proteasomal degradation and thus is receptors by acebutolol (as pineal melatonin production
crucial for maintaining the pace of molecular clock tim- is stimulated by sympathetic nervous system activation
ing117. UBE3A belongs to a cluster of imprinted genes of β1- and α1-receptors) has been effective for reducing
also within the 15q11–13 locus. Specifically, genomic daytime naps, improving concentration and reducing the
imprinting, a rare epigenetic mechanism of gene regu- frequency of tantrums in a sample of ten patients, aged
lation and developmental imprinting118, leads to parent-​ 4–18 years136. Combining β1-adrenergic receptor anta­
of-origin-​specific effects of deletion of the 15q11–13 gonism with timed administration of exogenous mel-
locus. For example, SNORD116 is expressed exclusively atonin greatly improves sleep (that is, leading to fewer
from the paternally inherited allele113, whereas mater- sleep awakenings and better organized sleep stages) and
nally inherited alleles of UBE3A are linked to PWS119,120. also reduces disruptive behaviours during the daytime
Maternal duplications of the 15q11–13 locus lead to in children and adolescents with SMS137. Thus, therapies
increased expression of UBE3A and epigenetic silencing targeting the reversal of melatonin rhythms in patients
of the paternal antisense transcript, UBE3A-​ATS121–124. with SMS may improve sleep, as well as behaviour,
The maternal allele of UBE3A is imprinted only in neu- attention and learning137–139.
rons and is expressed by alleles in other cell types120,125,126.
In mice with maternal loss of Ube3a, period lengthen- Adolescence
ing of locomotor rhythms and molecular rhythms in Humans, rodents and some NHPs undergo changes in
the brain are dependent on core clock factors, including circadian rhythms during adolescence140. At the onset
CLOCK and BMAL1 (ref.127). Intriguingly, activation of of puberty, sleep–wake cycles and melatonin rhythms
the epigenetically silenced paternal Ube3a allele restores start to phase delay141 — a shift that is dependent on the
neuronal circadian function127, suggesting that target- presence of gonadal hormones, and that coincides with
ing the locus may be useful for treatment and involve sexual maturity142,143. Environmental factors, such as
molecular-​clock-dependent mechanisms. Thus, duplica- social pressure from peers and/or decreasing parental
tion of the 15q11–13 locus and developmental imprint- involvement in bedtime routines, can also contribute
ing of SNORD116 and UBE3A may provide a molecular to changes in adolescent sleep timing and duration140.
explanation for the circadian rhythm disruptions Sensitivity to the phase-​shifting effects of light may also
common to patients with PWS. play a role, which is especially important when consid-
SMS is a rare neurodevelopmental disorder charac- ering the increasingly pervasive use of electronic devices
terized by craniofacial and skeletal anomalies, metabolic at night and their influence on sleep timing, duration
problems and obesity, along with intellectual disability and quality144. Adolescents may be more sensitive to
and stereotypical behaviours, and is a result of a hete- the ability of light exposure to shift rhythms at specific
rozygous deletion of chromosomal region 17p11.2. times of day. For example, light exposure during the
The deleted region contains several genes, including the night (23:00 h to 24:00 h) suppresses melatonin levels
gene encoding retinoic acid induced one (RAI1), the loss in adolescents, with greater suppression occurring in
of which is thought to be responsible for many of the early adolescents (prepubertal to mid-​pubertal adoles-
disease characteristics. Symptoms of SMS also include cents aged 9–14 years) than in late adolescents (aged
sleep and circadian abnormalities, with the most 11–15 years, in late pubertal to post-​pubertal stages)145.
extreme being a complete reversal of sleep–wake pat- Early-​morning light exposure (between 03:00 h and
terns (that is, daytime sleep and night-​time wake)128–131. 04:00 h) also reduced melatonin, although no differences

54 | JANUARY 2019 | volume 20 www.nature.com/nrn


Reviews

Eveningness
were found between early and late adolescents145. Thus, use disorders (Fig. 4). Experimentation with substances
The preference for activity in the sensitivity of the circadian system to light exposure often begins during mid-​adolescence to late adolescence
the evening and later bed depends on the developmental stage and time of day, (ages ~14–18 years), and the frequency of use peaks dur-
times, related to endogenous and early adolescents may be particularly vulnerable ing early to mid-20s153. Circadian disruptions (such as
circadian phase, and akin to
to the phase-​delaying effects of light at night145. Although social jet-​lag) during adolescence increase vulnerabil-
evening chronotype.
the amount of sleep needed per night is variable, the vast ity to substance use, and conversely, chronic exposure
majority of US adolescents sleep less than the suggested to these substances can lead to long-​lasting changes in
8–10 hours of sleep per night, according to the American the circadian and sleep networks154. Weekday–weekend
Academy of Pediatrics146,147. Early school start times, differences in sleep timing and duration are associated
which are typically more than an hour earlier in US high with increased risk-​taking behaviours, substance use
schools (ages 15–18 years) than in elementary schools and depressed mood155,156. Moreover, adolescents with
(ages 9–10 years) or middle schools (ages 11–13 years), shorter sleep duration are more likely to use substances,
may contribute to these problems of shorter sleep10,148. including caffeine, nicotine, alcohol and marijuana157–159,
On weekends, adolescents in the US tend to sleep later, and to engage in other risky behaviours 157,160,161.
both reverting to their natural sleep–wake cycle and pos- Longitudinal follow-​up studies have found that week-
sibly to recover from cumulative sleep loss during the end delay and sleep-​timing variability during adoles-
school week149. Transitions from weekday to weekend cence and young adulthood positively correlate with
compound the phase delay in the endogenous clock150,151 alcohol use 2 years later162 and the incidence of alcohol
that is often referred to as ‘social jet-​lag’148,152. use disorder symptoms 3 and 5 years later163 and are
associated with an earlier onset of alcohol use disor-
Substance use disorders. Emerging evidence from stud- der164. Greater eveningness and later sleep timing in ado-
ies in humans and animal models suggests that disrup- lescents and young adults are associated with increased
tions to sleep and circadian rhythms during adolescence substance abuse and may not merely be a consequence of
have consequences on brain development and might increased opportunity to engage in substance use later at
contribute to the vulnerability to mood and substance night165–167. However, the roles of biological factors versus

Adolescence
• Social pressures
(e.g. peers, schedule and family)
• Brain development
(e.g. cognition, emotionality and impulsivity)

Cognitive control, mood regulation and


reward function
• Dopaminergic activity?
• Inhibitory control? Psychiatric disorders
Circadian genes • Substance use disorders
• Variants • Depression
• Polymorphisms • Anxiety
• Expression • Bipolar disorder
• Schizophrenia
Circadian rhythms and sleep
• Environmental light
• Melatonin rhythms
• Sleep–wake cycles
• Sleep timing
• Sleep duration

Interventions
• Timed melatonin
• Timed light exposure
• Routine schedules

Fig. 4 | Effects of social constraints, sleep and circadian disruptions on adolescent brain function. A proposed model
for the associations between sleep, circadian rhythms and psychiatric disorders during adolescence. Neural circuits
responsible for controlling cognition, mood and reward mature rapidly during adolescence and may be negatively
affected by sleep deprivation and circadian misalignment, potentially leading to poor decision making, impulsivity and
risky behaviours. Basic and clinical studies have linked circadian genes and their variants to corticostriatal dopamine and
glutamatergic signalling that are important for cognition, mood and reward. These changes contribute to the vulnerability
of developing psychiatric disorders. Several psychiatric disorders, including mood disorders, substance use disorders and
schizophrenia, are associated with alterations in rhythms and/or sleep that affect cognition, motivation and impulsivity.
Sleep and/or circadian rhythm disruptions (for example, genetic and/or environmental perturbations) are also associated
with vulnerability to and the progression of psychiatric disorders. Interventions for targeting sleep and/or circadian
rhythms may be therapeutically effective for treating disorders that emerge during vulnerable periods of adolescence.

NAtuRe Reviews | NEuRosCiEnCE volume 20 | JANUARY 2019 | 55


Reviews

social influences remain unclear: social context is also that these interventions improved rates of alcohol absti-
important, and social networks may affect sleep duration nence, suggesting that improving sleep quality alone is
and substance use in adolescents168. not sufficient to prevent relapse. Determining whether
During adolescence, many developmental changes chronobiological interventions can reduce the risk of
occur in the frontostriatal reward circuitry, which includes developing substance use disorders or treat individuals
the ventral striatum (VS; including the nucleus accum- with these disorders is an important and translationally
bens), dorsal striatum (DS) and medial prefrontal cortex relevant avenue of future research.
(mPFC)169,170. These changes include increases in syn-
aptic pruning and myelination107,171, as well as increases Mood disorders. Adolescence is a time during which
in the availability of dopamine in the limbic circuitry172,173. serious psychiatric disorders, including major depres-
The combination of heightened activity in reward-​ sion, bipolar disorder and schizophrenia, tend to emerge.
related circuits and underdeveloped cognitive control The overall prevalence of disorders with severe impair-
centres contributes to greater emotionality, impulsivity ment and/or distress during this developmental period
and reward-​seeking behaviour174,175. Circadian and sleep is 22.2%182, and the median age of onset for mood dis-
disruptions may amplify these responses by further orders is 13 years of age182. A major component of many
reducing cortical control and increasing the activity of mood, anxiety and psychotic disorders is a disrupted
reward circuitry in human adolescents. Chronic expo- sleep–wake cycle183–188. In addition, circadian disrup-
sure to drugs of abuse can lead to epigenetic changes tions can precipitate mood and psychotic episodes in
affecting the molecular clock that may then alter func- individuals who already have psychiatric disorders189–191.
tion of the reward circuitry for long periods of time, Mood symptoms can also occur in a seasonal pattern,
creating more vulnerability for future drug abuse176. as in seasonal affective disorder and in many cases of
Functional MRI studies reveal that the activation bipolar disorder192,193; symptoms may be related to the
(that is, glucose uptake) of striatal and prefrontal brain amount, duration and intensity of light varying across
regions during monetary reward tasks is greater in the the seasons (that is, short versus long days)194. However,
afternoon or evening than in the morning in late ado- these disorders are very heterogeneous, and any asso-
lescents and young adults (aged 19–24 years)177, and is ciation with disruptions in the sleep–wake cycle can be
related to morningness and eveningness chronotypes in highly variable from individual to individual. Thus, sim-
adults178. In an age-​matched cohort of late adolescents, ilar to genetic studies, identifying a common circadian
greater VS responses to reward outcome and reduced or sleep-​related mechanism that is causal for all bipolar
mPFC responses to reward anticipation are associated disorder or major depression is unlikely.
with eveningness179 and may predict alcohol depend- Similar to substance use disorders, delayed rhythms
ence167, further suggesting that certain circadian pheno- during adolescence and early adulthood are strongly
types emerging during early to late adolescence could associated with depression and the severity of mood
contribute to vulnerability to substance use. In fact, symptoms195,196. More recent work has shown that circa-
larger weekday–weekend advances of sleep onset (for dian disruptions in adolescents who are at high risk of
example, similar to the transition from early school start psychosis (that is, subthreshold symptoms of psychosis,
times to weekend schedules) are associated with lower social impairment and higher suspicion–paranoia)197
mPFC reactivity to the anticipation and receipt of mone- and individuals with mood disorders predict worse
tary reward in early adolescents (aged 11–13 years), even prognosis and symptoms in longitudinal studies198,199.
after adjusting for puberty, sex and total sleep time149. Together, these findings suggest that circadian and sleep
Lower responsivity of the PFC could lead to impaired disruptions may have a role in the vulnerability to mood
inhibition of striatal circuits and related behaviours, disorders and the precipitation of disorder symptoms.
including impulsivity and risk-​taking. Whether these Although the mechanisms underlying these asso-
relationships between neural responses to reward and ciations remain to be elucidated, one possibility is that
alcohol use are attributable to circadian misalignment disrupted sleep and circadian rhythms affect the synaptic
or other factors, some of which may affect chronotype, pruning and maturation of neural circuits during ado-
remains unclear. For example, sleep deprivation can lescence3,4,107,108,200. Many of these selective synaptic prun-
also alter neural reward circuitry: a single night of total ing and refinement processes occur during sleep, and
sleep deprivation leads to decreases in mPFC activity sleep is crucial for proper circuit maturation and long-​
and increases in VS activity in response to rewarding term memory formation in humans and animals201–204.
Frontostriatal reward
stimuli in young adults (aged 18–25 years)180. Further In addition, slow-​wave activity during sleep dramatically
circuitry investigations in humans and animals are necessary decreases specifically during adolescence; this reduc-
Neural pathways that connect to determine how circadian misalignment and/or dis- tion is suggested to result directly from synaptic refine-
frontal regions with basal ruption, along with sleep deprivation, independently ment, further demonstrating important links between
ganglia and mediate reward
and together, may affect the development of reward and sleep and synaptic changes107,108. Excessive pruning or
value and motivation.
cognitive neural systems and whether interventions that too little pruning during adolescence may be linked
Slow-​wave activity target these systems are clinically beneficial. A recent to the development of psychiatric disorders205,206; for
An electrophysiological meta-​analysis performed on nine studies investigated example, 5 days of 50% sleep restriction in rats during
measure of slow (0.5–4 Hz), the impact of insomnia treatment on alcohol use disor- early adolescence led to alterations in the fractions
synchronized oscillatory
activity primarily expressed
der181. Behavioural (but not pharmacological) interven- of projections from the motor cortex to other regions of
during non-​rapid eye tions for insomnia improved sleep quality and reduced the brain207. Therefore, insults during adolescence that
movement sleep. symptoms of depression; however, there was no evidence affect sleep and circadian stability may contribute to

56 | JANUARY 2019 | volume 20 www.nature.com/nrn


Reviews

the pathophysiology underlying the vulnerability and gastrointestinal dysfunction, sleep disorders, diabetes
progression of these diseases later in life107,108. As an and depression227. A large study of 11,450 Canadian
estimated 7% of adolescents meet clinical criteria for nurses found that the strongest relationship between
delayed sleep phase syndrome, and the majority of ado- work schedule and depression was in those who work
lescents are reported to be sleep deprived208–210, delay- rapidly rotating schedules and undefined rotating
ing school start times should be universally considered schedules (as opposed to slowly rotating sched-
as a potential intervention for curbing the effect of ules) 228. This study suggests more severe circadian
these conditions211. disruptions are associated with a greater likelihood
Treatments to amplify, phase advance or delay cir- of developing depression. Another study of ~14,000
cadian rhythms have been developed for potential use workers in electronics manufacturing in South Korea
in psychiatric disorders in adults. These treatments revealed that, compared with daytime workers, shift-​
include bright light therapy, acute sleep deprivation, workers showed higher rates of insomnia, depression
interpersonal and social rhythm therapy and therapeu- and suicidal ideation (with the strongest association
tic melatonin agonists (such as agomelatine)189,212–214. with insomnia)229.
Moreover, antidepressant and mood-​stabilizing medica- A survey of ~4,000 flight attendants as part of the
tions, such as lithium, valproic acid and selective seroto- US National Health and Nutrition Survey (NHANES)
nin reuptake inhibitors (SSRIs), all have dramatic effects found that, compared with the general population,
on gene expression rhythms that may, at least in part, flight attendants are 2–5.7-fold more likely to have sleep
underlie their therapeutic efficacy215–217. Additionally, disorders, depression, anxiety and/or fatigue230. Female
the rapid and lasting antidepressant effects of low-​dose flight attendants also have a higher prevalence of repro-
ketamine can be predicted based on measures of loco- ductive cancers, skin cancers and other conditions such
motor activity rhythms before treatment and on how as chronic bronchitis and heart disease230,231, probably as
well the treatment strengthens low-​amplitude activity a consequence of carcinogens in cabin environments,
rhythms, suggesting that the therapeutic effects are including cosmic radiation and poor aircraft ventilation,
directly related to the strengthening of the circadian but perhaps also exacerbated by night shift-​work, irreg-
system218. Thus, rhythm alignment and amplification ular schedules and frequent crossing of time zones232,233.
are important therapeutic approaches to helping prevent Taken together, many health problems during adult-
or treat psychiatric disorders. Such chronotherapy for hood, including sleep problems, depression and anxiety,
these disorders, however, is in its infancy and needs to are linked to circadian disruptions that are caused by
be evidence-​based. Chronotherapies may also need to be shift work or other environmental or genetic disruptions
personalized based on the particular needs of the indi- to the sleep–wake cycle.
vidual; for example, subjects with depression are more
Delayed sleep phase
responsive to light therapy combined with wake therapy Later life
syndrome and sleep-​time stabilization if they are evening chrono- Older people (ages ~65 years and older) generally sleep
A circadian rhythm disorder in types and have a positive diurnal variation in mood, with less and have poorer sleep efficiency, increased night-​
which sleep timing is shifted the best mood occurring in the evening219. time awakening, increased sleep latency and greater
later by 2 or more hours
levels of daytime sleepiness234–236. Several studies have
relative to normative
sleep–wake cycles and often Adulthood reported declines in melatonin levels with ageing,
accompanied by similar shifts For most people, there is a gradual shift towards although others find no difference compared with
in body-​temperature and an earlier chronotype from adolescence into adult- younger people in a healthy population237,238. When
hormone rhythms. hood141. Between 20 and 50 years of age, men continue other circadian outputs are measured, older individuals
Acute sleep deprivation
to have greater evening preference on average than show lower-​amplitude body temperature rhythms, with
A partial night or full night of women141,220,221, and these sex differences disappear the minimum of the endogenous phase occurring nearly
sleep restriction. This approach after the age of 50 years, coinciding with menopause in 2 hours earlier than in younger individuals239.
can produce a short-​term women141,220,222. However, these measures of chronotype Importantly, measures of activity and physiological
antidepressant effect.
are also affected by social and work schedules, and some rhythms in older people can be confounded by health
Interpersonal and social individuals may never shift back to a morning prefer- problems, medications and eye problems such as cata-
rhythm therapy ence after adolescence223. Mutations in clock genes such racts that can reduce light input to the central clock. In
A cognitive behavioural as CRY1, NFIL3 and RORC have been linked to extreme humans and rodent models, the amplitude of peripheral
therapy often used for bipolar phase delay (that is, delayed sleep–wake phase disorder) oscillatory rhythms dampens with age, but whether this
disorder that acts to stabilize
sleep–wake and social
in genetic studies of families224,225. The risk of depression effect is due to a loss of intrinsic core clock function,
schedules. is increased in people with this disorder and particu- to dysfunction of SCN signalling and connections240,241
larly in those with a biological circadian misalignment or to a deficit in entrainment to the environment is
Dim light melatonin onset (as measured by the dim light melatonin onset (DLMO) unclear242. The number of vasopressin-​expressing cells
(DLMO). A marker of
phase). Moreover, most people who are diagnosed in the SCN in humans243,244 and rodents245,246 reduces
endogenous circadian phase
that can be used for research with bipolar disorder have an evening chronotype that with age, possibly altering SCN output. Moreover, some
or to determine the persists after adolescence191,226. SCN cells become silent in older animals (as measured
appropriate timing of in vitro), and SCN neurons lose phase coherence with
treatments. Shift-​work disorder and chronic jet-​lag. Our 24-hour age247,248. These SCN network changes result in desyn-
Fatigue
society depends on both shift-​work and frequent travel chronization and a reduction in the amplitude of neu-
The feeling of exhaustion and across time zones. Shift-​work has been associated ronal activity rhythms248,249 and, ultimately, physiological
lack of energy. with increased risk of cancer, obesity, heart disease, and behavioural rhythms241.

NAtuRe Reviews | NEuRosCiEnCE volume 20 | JANUARY 2019 | 57


Reviews

In addition to SCN changes, gene expression rhythms and amplitude reduction in locomotor activity rhythms
in the human mPFC also decline with age; however, sur- during the active phase, and these effects progress with
prisingly, a separate set of genes becomes newly rhyth- age276. SCN neuronal firing is also reduced in these mice
mic in older people (aged 65 years and older)11. These from young adulthood and progressively worsens with
rhythmic genes might be protective, as the individuals age, suggesting a weakened circadian pacemaker276
in this post-​mortem study did not experience psychiatric and a potential loss of dopaminergic modulation of
or neurodegenerative disorders before death. There is circadian rhythms via midbrain, striatal and SCN
also accumulating evidence linking disrupted sleep and circuits277,278. MitoPark transgenic mice also show a pro-
circadian rhythms to neurodegenerative disorders in found disruption of locomotor rhythms in constant
humans and animal models235,236. conditions and are more vulnerable to changes in the
light–dark cycle279.
Neurodegenerative disorders. In older adults, less-​ Sundowning syndrome is prevalent in people with
robust circadian rhythms and more fragmented pat- dementia and is characterized by increases in aggres-
terns of activity are risk factors for the development sion, restlessness, delirium and agitation specifically
of dementia 250. Moreover, the incidence of single-​ in the late afternoon or early evening 280,281. Clinical
nucleotide polymorphisms (SNPs) in CLOCK and and preclinical data suggest that a lack of appropriate
BMAL1, and in BMAL1 and PER1, are associated with light–dark cues, disturbances in sleep (particularly,
increased risk of Alzheimer disease (AD) and reductions in REM sleep) and a deterioration of the
Parkinson disease (PD), respectively251–253. Compared SCN pacemaker and its outputs all contribute to sun-
with healthy aged people, individuals with AD or PD downing syndrome281. Phase delays and reductions in
show considerably lower melatonin rhythm amplitudes the amplitude of body temperature rhythms correlate
and excessive sleepiness, as well as other sleep–wake with the severity of sundowning symptoms, suggest-
cycle disturbances, such as later sleep onsets254–258. ing that a breakdown of the circadian system may be
In individuals with PD, these sleep–wake symptoms responsible282. Aggression, for example, may be mod-
often precede the development of motor or cognitive ulated by the circadian system. A recent study in mice
symptoms and may even be useful as a diagnostic bio- revealed that rhythms in the propensity of aggressive
marker259,260, whereas in patients with AD, sleep disrup- behaviours are modulated via a polysynaptic circuit
tions tend to begin after diagnosis261. AD brains show from VIP neurons in the SCN to GABAergic neurons
marked loss of neurons in the SCN243,244,262, and SCN of the subparaventricular zone, which in turn inner-
neuronal loss correlates with reductions in the ampli- vate the ventromedial hypothalamus, an area known
tude of motor activity rhythms in the same individuals to regulate aggression283. Further studies are needed to
before death240. In addition, although individuals with fully elucidate the biological mechanisms that underlie
AD still show cycling clock gene expression in multiple this syndrome, which affects many elderly people and
REM sleep behaviour brain regions, they lose typical phase coherence within their families.
disorder and across regions263. Several processes regulated by the circadian clock
(RBD). A disorder in which In mice, SCN levels of sirtuin 1 (SIRT1), a metabolic could contribute to neurodegeneration, including the
there is a loss of the paralysis
sensing protein and mediator of CLOCK and BMAL1 regulation of oxidative stress, inflammation, dopamine
that normally occurs during
rapid eye movement sleep, function (Fig. 1), decline with age, potentially explaining synthesis and cellular metabolism. One simple hypothe-
allowing the person to the increase in circadian rhythm disruptions in older sis is that disruptions in circadian rhythms disrupt sleep,
physically ‘act out’ their people241. Indeed, brain-​specific knockout of Sirt1 in in turn resulting in reduced clearance of misfolded and
dreams. rodents is sufficient to phenocopy the effects of ageing aggregated proteins that contribute to neurodegenerative
Restless legs syndrome
on the SCN and circadian behaviour, whereas overex- disorders from the brain. A recent study found that even
A condition marked by pression of Sirt1 is protective241. Recent studies revealed one night of sleep deprivation led to accumulation of
uncomfortable sensations in that SIRT1 binds to CLOCK at the promoter of the gene amyloid-​β (Aβ; one of the proteins that is pathologically
the legs accompanied by the encoding tyrosine hydroxylase (TH) in dopaminer- accumulated in the brain in AD) in the right hippocam-
urge to move them.
gic neurons to control rhythms in TH expression and, pus and thalamus in the human brain284. These increases
MitoPark transgenic mice ultimately, dopamine synthesis264 (Fig. 5). When iron is were negatively correlated with mood but were not asso-
A mouse model of Parkinson present in the cytosol (as may occur more frequently ciated with a particular AD-​linked genotype (that is, the
disease in which dopamine with ageing)265,266, dopamine can be oxidized into dopa- APOE genotype). In line with the clearance hypothesis,
neurons are deficient in mine quinone (DAQ), which is highly reactive and during sleep, the glymphatic system285 exchanges cere-
mitochondrial function through
inactivation of mitochondrial
toxic267,268. DAQ has been found in human induced brospinal fluid with interstitial fluid and clears Aβ286.
transcription factor A. pluripotent stem cell-​derived dopaminergic neurons Sleep deprivation increases interstitial fluid levels of Aβ
that carry mutations associated with PD269. Thus, defi- and increases plaque formation in mice and flies287,288,
REM sleep cits in SIRT1 may contribute to toxic levels of DAQ and whereas in Drosophila, genetically increasing sleep
A phase of sleep characterized
neurodegeneration. reduces Aβ deposition288. However, based on estimates
by rapid movement of the
eyes, low muscle tone and PD, in particular, is strongly associated with rapid of anatomical and fluid dynamic parameters, the plau-
dreams that can later be eye moment (REM) sleep behaviour disorder (RBD), as sibility of periarterial flow and glymphatic circulation
recollected. well as excessive daytime sleepiness, insomnia and is controversial; thus, further characterization of this
restless legs syndrome 270–273. In fact, more than 80% system is needed289.
Glymphatic system
A functional waste clearance
of people with RBD are later diagnosed with PD The build-​up and expulsion of reactive oxygen spe-
system in the brain that is or dementia 274,275. Transgenic mice overexpressing cies (ROS) also follow a circadian pattern. In diurnal
active during sleep. α-​synuclein (a model of PD) show greater fragmentation species (including humans), ROS accumulate during

58 | JANUARY 2019 | volume 20 www.nature.com/nrn


Reviews

TH promoter astrocytosis and levels of oxidative damage markers in


RRE/NBRE E-box CRE the brain. Treating Bmal1−/− mice with the antioxidant
N-​acetyl-l-​cysteine partially rescues their lifespan and

CREB
prevents age-​dependent pathology292. Interestingly,
REV NURR1

CLOCK
BMAL1
mice with a brain-​specific Bmal1 deletion also exhibit

SIRT1
increased astrocytosis, despite having mostly normal
CCK promoter locomotor circadian rhythms and sleep–wake cycling
E-box (that is, relatively normal clock function in the SCN);
thus, these neurodegenerative effects are probably due

CLOCK
BMAL1
to loss of the local control of cellular metabolism in spe-
VTA MLL1
dopaminergic cific regions of the brain that is normally mediated by
neuron BMAL1, independently or as part of the molecular clock,
Tyrosine in these cells291.
TH Whether circadian rhythm and sleep disruptions
D2R Mitochondrion are primarily causal to neurodegenerative disorders
L-DOPA
remains unclear, but environmental and biological dis-
DA ruptions to these systems do seem to worsen progres-
sion of these diseases293–295. Thus, therapies that stabilize
MAOA
circadian and sleep rhythms may at least slow the pro-
DOPAC gression of these disorders. Indeed, some studies indi-
NAc medium cate that morning light therapy may improve circadian
spiny neuron rhythms and Mini-​Mental State Examination (MMSE),
Glu DAT Cohen-​Mansfield Agitation Inventory (CMAI) and
GABA Behaviour Pathology in AD Rating Scale (BEHAVE-​
GABAR
CCK AD) scores, particularly in the early stages of dementia,
and that, even in severe cases of dementia, it improves
D1R or circadian and behavioural symptoms, most notably agi-
D3R Gi Gs
GluR D2R
tation and overall cognitive function296,297. Care centres
for patients with dementia are often dimly lit298–300, and
a randomized control trial of 189 participants found
Cellular response that increasing illumination levels in elderly care cen-
tres during daytime hours leads to a slower decline in
Fig. 5 | Circadian regulation of dopamine. The ventral tegmental area (VTA)–nucleus MMSE scores, reduced depression scores and reduced
accumbens (NAc) circuitry is regulated by a local molecular clock , which controls the functional impairment in individuals with dementia
transcription of genes involved in the dopamine (DA) signalling pathway , including those over a 1.5-year period301. Moreover, a double-​blind,
encoding: tyrosine hydroxylase (TH), the major rate-​limiting enzyme converting tyrosine multicentre study in 157 individuals with AD showed
to the DA precursor l-​dihydroxyphenylalanine (l-​DOPA); cholecystokinin (CCK), a
that evening melatonin treatment improved cognition,
neuropeptide released from presynaptic DA terminals to suppress further DA release;
and monoamine oxidase A (MAOA), a mitochondrial enzyme used to metabolize decreased nocturnal activity and increased sleep302.
monoamine neurotransmitters, including DA. For example, the circadian transcription Importantly, melatonin treatment must be carefully
factors of circadian locomotor output cycles protein kaput (CLOCK) and brain and timed to that of normal endogenous melatonin synthe-
muscle ARNT-​like 1 (BMAL1) form heterodimers and recruit the metabolic sensor and sis, as other studies that provided melatonin later in the
histone acetylase sirtuin 1 (SIRT1) to the enhancer box (E-​box) within the TH promoter to evening did not find beneficial effects302,303.
repress transcription by cAMP response element-​binding protein (CREB). CREB activates
TH transcription via the CRE site, located proximal to the transcription start site and Conclusions
the circadian E-​box. In addition, CLOCK is capable of binding the E-​box within the Cck Throughout the lifespan, sleep and circadian rhythm
promoter to promote the transcription of Cck in the VTA. Mixed lineage leukaemia disruptions are strongly linked to the pathophysiology
protein 1 (MLL1) is also recruited to CLOCK–BMAL1 heterodimers to promote the
of specific psychiatric and neurodegenerative disorders.
transcription of target genes via histone acetylation. REV-​ERBα (REV) also represses
transcription by binding the REV-​ERB response element (RRE) promoter site in antiphase Though most clinical studies are still correlational,
with the transcription factor NURR1 at the NURR1 binding motif, NBRE. Presynaptic animal model studies are beginning to examine poten-
dopamine 2 receptors (D2Rs) may also be regulated by the molecular clock. Diurnal tial causal relationships between disrupted circadian
rhythms of dopamine, glutamate and GABA levels, as well as of the expression and rhythms and multiple brain disorders and are starting
activity of their receptors, are present in the NAc, potentially temporally gating and to identify molecular mechanisms. Although the cen-
promoting neuronal responses during certain times of day. DAT, DA transporter ; tral clock in the SCN seems to be key, circadian genes
DOPAC, 3-4-dihydroxyphenylacetic acid; GABAR , GABA receptor ; Glu, glutamate; in other brain regions outside of this central pace-
GluR , glutamate receptor (ionotropic). maker also contribute locally to the control of neuronal
metabolism, neurotransmitter synthesis and activity,
the day owing to neuronal activity while awake; at night, and the disruption of these functions might contrib-
antioxidants safely remove excess ROS during sleep236,290. ute to brain disorders264,304–315. We are only beginning
Astrocytosis Bmal1−/− mice, which lack a functional molecular clock, to understand the role of peripheral oscillators in
An abnormal increase in the
number of astrocytes,
show premature ageing, neurodegeneration and a multiple regions of the brain, how they are entrained
attributable to the death of reduced lifespan291. ROS accumulation is considerably and what other functions circadian transcription fac-
nearby neurons. higher in Bmal1−/− mice and correlates with increases in tors have that might be independent of the control of

NAtuRe Reviews | NEuRosCiEnCE volume 20 | JANUARY 2019 | 59


Reviews

Mini-​Mental State
daily rhythms. These points should certainly be the addiction. Recent genome-​wide association studies with
Examination focus of future study. more than 100,000 subjects have identified novel loci
(MMSE). A standardized set of We need to better understand the bidirectional associated with sleep duration, chronotype and meta-
11 questions with a maximal relationships between circadian rhythms and environ- bolic phenotypes (for example, body mass index)320,321.
score of 30 (≤23 indicating
mental perturbations, such as stress or drugs of abuse, The role of these environmental and genetic factors
cognitive impairment) used to
assess five cognitive functional and how circadian rhythms are intimately connected underlying circadian rhythms and brain disorders can
areas: orientation, registration, to neurotransmission, metabolism, immunity and be further experimentally interrogated using cell-​based
attention and calculation, other processes. Circadian disruptions may promote assays and preclinical models. Manipulation of specific
recall and language.
vulnerability to or the progression of certain disorders, circadian genes within neural cell types and regions will
within a developmental context. With development, begin to dissect the involvement of molecular and cel-
changes occur in the sensitivity of the brain to light lular clocks in neural circuits and relevant physiological
input, in the inputs and outputs of the SCN, and in the and behavioural end points. Optogenetic and chemo­
functions of the molecular clock in specific cell types genetic tools can be used to tease apart the function
in the brain and in other tissues. In later life, certain of light-​input pathways to the brain that are dependent
disorders exhibit a defined trajectory, with sleep and on, or independent from, SCN connections in regulat-
circadian problems often preceding the development ing neural activity, as has recently been demonstrated
of other symptoms. These findings provide an opening for mood-​related behaviours322–324. With the advent of
for future, in-​depth investigation of the causal relation- new tools and resources in neuroscience, and build-
ships between sleep and/or circadian disruption and ing on a greater understanding of the relationships
neurodegeneration. between circadian rhythms, physiology and behaviour,
Recent studies have demonstrated that elements of research in humans and animals is equipped to investi-
the molecular circadian clock directly regulate cellular gate causal relationships between circadian disruption
metabolism and mitochondrial function in the liver and and various brain disorders. These findings may lead to
other tissues316–318. Further studies should determine the effective therapeutics and interventions with improved
molecular mechanisms by which circadian genes may clinical outcomes.
control the metabolism, mitochondrial function, redox Chronotherapeutic strategies of specifically timing
state, antioxidant response and activity of neurons and the delivery of certain drugs are now being used in the
glia, as dysfunction of each of these processes and states treatment of certain cancers325 and may be therapeu-
can contribute to neurodegeneration. In addition, future tically beneficial across various diseases 326. Time-​
studies should determine the circadian rhythm of the restricted feeding is also now being used to help treat
expression of drug targets. For example, drugs that and/or prevent obesity, as well as metabolic, cardiac and
target certain neurotransmitters and their receptors liver disorders327,328. Several first-​line pharmacological
may have higher therapeutic efficacy or more adverse treatments for psychiatric disorders, such as lithium
effects depending on the time of day the medication for bipolar disorder and SSRIs for major depressive
is administered and on the endogenous rhythm of disorders, modulate the circadian clock 216,217,329–331.
neurotransmission and receptor expression. Although these effects were identified later, they sug-
The evidence for associations between circadian gest that targeting rhythms or components of the clock
disruptions and brain disorders in humans is mostly using pharmacological agents might be therapeutically
epidemiological and correlational. Studies in preclinical beneficial215,216,331. Many of the most effective treat-
models have supported these findings while also offer- ments for brain disorders and other diseases modulate
ing insights into the bidirectionality, in which circadian clock function216,326,330,332,333. Importantly, treatments
disruption leads to certain disease-​related phenotypes, and interventions that target the circadian system
and the environmental factors that can alter rhythms, may have therapeutic efficacy and improve clinical
such as stress or exposure to drugs of abuse, can lead to symptoms and daily functioning for the individual.
the exacerbation or progression of certain symptoms. Such approaches, if considered, could even be added
Future research will require the continuation of exper- as adjunct therapies to first-​line medications or treat-
imental investigation of causality using translational ments to further improve therapeutic outcomes. There
approaches. For example, investigators are beginning to is much interest in the implementation of regular social,
implement laboratory manipulations in humans to mis- sleep and activity schedules in children and adolescents
align the environmental schedule with an individual’s with developmental disorders, as stable environmental
endogenous sleep–wake and melatonin rhythms to schedules seem to ameliorate daytime sleepiness and
assess the impact of circadian alignment on cognitive, daytime behavioural problems in individuals with
reward and memory functions, as well as their related ASD or ADHD. The increased use of wearable activity
neural circuits. Considering these hypotheses within and sleep-​tracking devices has also brought about new
a developmental context will be necessary, as certain possibilities in terms of diagnostic analysis and person-
developmental periods (for example, adolescence) alized treatment plans334. The field is poised to align
may be vulnerable to these disruptions. Larger-​scale, multiple approaches, including tracking of brain func-
more in-​depth assessments of chronotype utilizing tion, receptor levels and internal rhythms, that could
novel biological assays319, combined with longitudinal be valuable for treating psychiatric and neurological
follow-​ups, could further clarify the predictive asso- disorders in the future.
ciations between eveningness or delayed chronotypes
and brain disorders such as depression, bipolar disorder and Published online 20 November 2018

60 | JANUARY 2019 | volume 20 www.nature.com/nrn


Reviews

1. Reppert, S. M. & Weaver, D. R. Coordination of 29. Shearman, L. P., Zeitzer, J. & Weaver, D. R. 53. Davis, K. F., Parker, K. P. & Montgomery, G. L.
circadian timing in mammals. Nature 418, 935–941 Widespread expression of functional D1-dopamine Sleep in infants and young children: part two: common
(2002). receptors in fetal rat brain. Brain Res. Dev. Brain Res. sleep problems. J. Pediatr. Health Care 18, 130–137
2. Colwell, C. S. Linking neural activity and molecular 102, 105–115 (1997). (2004).
oscillations in the SCN. Nat. Rev. Neurosci. 12, 30. Reiter, R. J., Tan, D. X., Korkmaz, A. & Rosales-​Corral, 54. Dahl, R. E. Sleep, learning, and the developing brain:
553–569 (2011). S. A. Melatonin and stable circadian rhythms optimize early-​to-bed as a healthy and wise choice for school
3. Selemon, L. D. A role for synaptic plasticity in the maternal, placental and fetal physiology. Hum. Reprod. aged children. Sleep 28, 1498–1499 (2005).
adolescent development of executive function. Transl Update 20, 293–307 (2014). 55. Wolke, D., Meyer, R., Ohrt, B. & Riegel, K. Incidence
Psychiatry 3, e238 (2013). 31. Okatani, Y. et al. Maternal–fetal transfer of melatonin and persistence of problems at sleep onset and sleep
4. Mighdoll, M. I., Tao, R., Kleinman, J. E. & Hyde, T. M. in pregnant women near term. J. Pineal Res. 25, continuation in the preschool period: results of a
Myelin, myelin-​related disorders, and psychosis. 129–134 (1998). prospective study of a representative sample in Bavaria
Schizophr. Res. 161, 85–93 (2015). 32. Schenker, S. et al. Antioxidant transport by the human [German]. Prax Kinderpsychol. Kinderpsychiatr. 43,
5. Hsu, P. K., Ptacek, L. J. & Fu, Y. H. Genetics of human placenta. Clin. Nutr. 17, 159–167 (1998). 331–339 (1994).
sleep behavioral phenotypes. Methods Enzymol. 552, 33. Davis, F. C. & Mannion, J. Entrainment of hamster pup 56. Gaylor, E. E., Burnham, M. M., Goodlin-​Jones, B. L.
309–324 (2015). circadian rhythms by prenatal melatonin injections to & Anders, T. F. A longitudinal follow-​up study of young
6. Roenneberg, T. & Merrow, M. Entrainment of the the mother. Am. J. Physiol. 255, R439–R448 (1988). children’s sleep patterns using a developmental
human circadian clock. Cold Spring Harb. Symp. Quant. 34. Seron-​Ferre, M. et al. Circadian rhythms in the fetus. classification system. Behav. Sleep Med. 3, 44–61
Biol. 72, 293–299 (2007). Mol. Cell. Endocrinol. 349, 68–75 (2012). (2005).
7. Foster, R. G. et al. Sleep and circadian rhythm disruption 35. Seron-​Ferre, M., Riffo, R., Valenzuela, G. J. 57. Touchette, E. et al. Associations between sleep
in social jetlag and mental illness. Prog. Mol. Biol. Transl & Germain, A. M. Twenty-​four-hour pattern of cortisol duration patterns and behavioral/cognitive functioning
Sci. 119, 325–346 (2013). in the human fetus at term. Am. J. Obstet. Gynecol. at school entry. Sleep 30, 1213–1219 (2007).
8. Werner, H., Lebourgeois, M. K., Geiger, A. & Jenni, O. G. 184, 1278–1283 (2001). 58. Kobayashi, K. et al. Poor toddler-​age sleep schedules
Assessment of chronotype in four- to eleven-​year-old 36. Kennaway, D. J., Goble, F. C. & Stamp, G. E. predict school-​age behavioral disorders in a longitudinal
children: reliability and validity of the Children’s Factors influencing the development of melatonin survey. Brain Dev. 37, 572–578 (2015).
Chronotype Questionnaire (CCTQ). Chronobiol. Int. 26, rhythmicity in humans. J. Clin. Endocrinol. Metab. 81, 59. Gregory, A. M. et al. Prospective longitudinal
992–1014 (2009). 1525–1532 (1996). associations between persistent sleep problems
9. Simpkin, C. T. et al. Chronotype is associated with 37. Bisanti, L., Olsen, J., Basso, O., Thonneau, P. in childhood and anxiety and depression disorders in
the timing of the circadian clock and sleep in toddlers. & Karmaus, W. Shift work and subfecundity: a adulthood. J. Abnorm. Child Psychol. 33, 157–163
J. Sleep Res. 23, 397–405 (2014). European multicenter study. European Study Group (2005).
10. Crowley, S. J. et al. A longitudinal assessment of on Infertility and Subfecundity. J. Occup. Environ. 60. Gregory, A. M., Caspi, A., Moffitt, T. E. & Poulton, R.
sleep timing, circadian phase, and phase angle Med. 38, 352–358 (1996). Sleep problems in childhood predict neuropsychological
of entrainment across human adolescence. PLOS ONE 38. Aspholm, R. et al. Spontaneous abortions among functioning in adolescence. Pediatrics 123, 1171–1176
9, e112199 (2014). Finnish flight attendants. J. Occup. Environ. Med. 41, (2009).
11. Chen, C. Y. et al. Effects of aging on circadian patterns 486–491 (1999). 61. Hysing, M., Sivertsen, B., Garthus-​Niegel, S.
of gene expression in the human prefrontal cortex. 39. Cone, J. E., Vaughan, L. M., Huete, A. & Samuels, S. J. & Eberhard-​Gran, M. Pediatric sleep problems and
Proc. Natl Acad. Sci. USA 113, 206–211 (2015). Reproductive health outcomes among female flight social-​emotional problems. A population-​based study.
This is the largest-​scale study of circadian rhythms attendants: an exploratory study. J. Occup. Environ. Infant Behav. Dev. 42, 111–118 (2016).
of gene expression in human post-​mortem Med. 40, 210–216 (1998). 62. Gregory, A. M., Eley, T. C., O’Connor, T. G. & Plomin, R.
brain tissue, revealing age-​related decline in 40. Mahoney, M. M. Shift work, jet lag, and female Etiologies of associations between childhood sleep
the expression of core circadian genes and the reproduction. Int. J. Endocrinol. 2010, 813764 and behavioral problems in a large twin sample.
emergence of other rhythmic pathways in older (2010). J. Am. Acad. Child Adolesc. Psychiatry 43, 744–751
subjects. 41. Varcoe, T. J. et al. Characterisation of the maternal (2004).
12. Cornelissen, G. & Otsuka, K. Chronobiology of aging: response to chronic phase shifts during gestation in 63. Bendova, Z., Sumova, A. & Illnerova, H. Development
a mini-​review. Gerontology 63, 118–128 (2017). the rat: implications for fetal metabolic programming. of circadian rhythmicity and photoperiodic response
13. Drake, C. L., Roehrs, T., Richardson, G., Walsh, J. K. PLOS ONE 8, e53800 (2013). in subdivisions of the rat suprachiasmatic nucleus.
& Roth, T. Shift work sleep disorder: prevalence and 42. Torres-​Farfan, C. et al. Maternal melatonin selectively Brain Res. Dev. Brain Res. 148, 105–112 (2004).
consequences beyond that of symptomatic day workers. inhibits cortisol production in the primate fetal adrenal 64. Duncan, M. J., Banister, M. J. & Reppert, S. M.
Sleep 27, 1453–1462 (2004). gland. J. Physiol. 554, 841–856 (2004). Developmental appearance of light–dark entrainment
14. Ramin, C. et al. Night shift work at specific age ranges 43. Seron-​Ferre, M. et al. Impact of chronodisruption in the rat. Brain Res. 369, 326–330 (1986).
and chronic disease risk factors. Occup. Environ. Med. during primate pregnancy on the maternal and 65. Fahrenkrug, J., Nielsen, H. S. & Hannibal, J.
72, 100–107 (2015). newborn temperature rhythms. PLOS ONE 8, e57710 Expression of melanopsin during development of the
15. Bedrosian, T. A. & Nelson, R. J. Timing of light exposure (2013). rat retina. Neuroreport 15, 781–784 (2004).
affects mood and brain circuits. Transl Psychiatry 7, 44. Matsumoto, T. et al. Circadian myometrial and 66. Gonzalez-​Menendez, I., Contreras, F., Cernuda-​
e1017 (2017). endocrine rhythms in the pregnant rhesus macaque: Cernuda, R. & Garcia-​Fernandez, J. M. Daily rhythm
16. Bedrosian, T. A. & Nelson, R. J. Influence of the effects of constant light and timed melatonin infusion. of melanopsin-​expressing cells in the mouse retina.
modern light environment on mood. Mol. Psychiatry Am. J. Obstet. Gynecol. 165, 1777–1784 (1991). Front. Cell Neurosci. 3, 3 (2009).
18, 751–757 (2013). 45. Novakova, M., Sladek, M. & Sumova, A. Exposure 67. Landgraf, D., Koch, C. E. & Oster, H. Embryonic
17. Nakata, A. et al. Association of sickness absence with of pregnant rats to restricted feeding schedule development of circadian clocks in the mammalian
poor sleep and depressive symptoms in shift workers. synchronizes the SCN clocks of their fetuses under suprachiasmatic nuclei. Front. Neuroanat. 8, 143
Chronobiol. Int. 21, 899–912 (2004). constant light but not under a light-​dark regime. (2014).
18. Roth, T. Shift work disorder: overview and diagnosis. J. Biol. Rhythms 25, 350–360 (2010). 68. Leard, L. E., Macdonald, E. S., Heller, H. C.
J. Clin. Psychiatry 73, e09 (2012). 46. Vilches, N. et al. Gestational chronodisruption impairs & Kilduff, T. S. Ontogeny of photic-​induced c-​fos mRNA
19. Wright, K. P. Jr, Bogan, R. K. & Wyatt, J. K. Shift work hippocampal expression of NMDA receptor subunits expression in rat suprachiasmatic nuclei. Neuroreport
and the assessment and management of shift work Grin1b/Grin3a and spatial memory in the adult 5, 2683–2687 (1994).
disorder (SWD). Sleep Med. Rev. 17, 41–54 (2012). offspring. PLOS ONE 9, e91313 (2014). 69. Tarttelin, E. E. et al. Expression of opsin genes early in
20. Benca, R. et al. Biological rhythms, higher brain 47. Voiculescu, S. E. et al. Behavioral and molecular ocular development of humans and mice. Exp. Eye
function, and behavior: gaps, opportunities, and effects of prenatal continuous light exposure in the Res. 76, 393–396 (2003).
challenges. Brain Res. Rev. 62, 57–70 (2009). adult rat. Brain Res. 1650, 51–59 (2016). 70. Rivkees, S. A. The development of circadian rhythms:
21. Poggiogalle, E., Jamshed, H. & Peterson, C. M. 48. Smarr, B. L., Grant, A. D., Perez, L., Zucker, I. from animals to Humans. Sleep Med. Clin. 2, 331–341
Circadian regulation of glucose, lipid, and energy & Kriegsfeld, L. J. Maternal and early-​life circadian (2007).
metabolism in humans. Metabolism 84, 11–27 (2018). disruption have long-​lasting negative consequences 71. Guyer, C. et al. Very preterm infants show earlier
22. Chaudhari, A., Gupta, R., Makwana, K. & Kondratov, R. on offspring development and adult behavior in mice. emergence of 24-hour sleep-​wake rhythms compared
Circadian clocks, diets and aging. Nutr. Healthy Aging Sci. Rep. 7, 3326 (2017). to term infants. Early Hum. Dev. 91, 37–42 (2015).
4, 101–112 (2017). 49. Cisse, Y. M., Russart, K. L. & Nelson, R. J. Parental 72. Boivin, D. B., Duffy, J. F., Kronauer, R. E. & Czeisler, C. A.
23. Rivkees, S. A. Developing circadian rhythmicity in exposure to dim light at night prior to mating alters Dose-​response relationships for resetting of human
infants. Pediatr. Endocrinol. Rev. 1, 38–45 (2003). offspring adaptive immunity. Sci. Rep. 7, 45497 circadian clock by light. Nature 379, 540–542 (1996).
24. Swaab, D. F. Development of the human hypothalamus. (2017). 73. Rivkees, S. A., Hofman, P. L. & Fortman, J. Newborn
Neurochem. Res. 20, 509–519 (1995). 50. Cisse, Y. M., Russart, K. L. G. & Nelson, R. J. primate infants are entrained by low intensity lighting.
25. VanDunk, C., Hunter, L. A. & Gray, P. A. Development, Depressive-​like behavior is elevated among offspring Proc. Natl Acad. Sci. USA 94, 292–297 (1997).
maturation, and necessity of transcription factors in of parents exposed to dim light at night prior to 74. Shanahan, T. L. & Czeisler, C. A. Physiological effects
the mouse suprachiasmatic nucleus. J. Neurosci. 31, mating. Psychoneuroendocrinology 83, 182–186 of light on the human circadian pacemaker. Semin.
6457–6467 (2011). (2017). Perinatol. 24, 299–320 (2000).
26. Reppert, S. M. Pre-​natal development of a 51. Lunn, R. M. et al. Health consequences of electric 75. Watanabe, S. et al. Designing artificial environments
hypothalamic biological clock. Progress Brain Res. 93, lighting practices in the modern world: a report on for preterm infants based on circadian studies on
119–131; discussion 132 (1992). the National Toxicology Program’s workshop on shift pregnant uterus. Front. Endocrinol. 4, 113 (2013).
27. Reppert, S. M. Interaction between the circadian work at night, artificial light at night, and circadian 76. Morag, I. & Ohlsson, A. Cycled light in the intensive
clocks of mother and fetus. Ciba Found. Symp. 183, disruption. Sci. Total Environ. 607–608, 1073–1084 care unit for preterm and low birth weight infants.
198–207; discussion 207–111 (1995). (2017). Cochrane Database Syst. Rev. 8, CD006982 (2013).
28. Reppert, S. M., Weaver, D. R. & Rivkees, S. A. 52. Kennaway, D. J., Stamp, G. E. & Goble, F. C. 77. Vasquez-​Ruiz, S. et al. A light/dark cycle in the NICU
Maternal communication of circadian phase to the Development of melatonin production in infants and accelerates body weight gain and shortens time to
developing mammal. Psychoneuroendocrinology 13, the impact of prematurity. J. Clin. Endocrinol. Metab. discharge in preterm infants. Early Hum. Dev. 90,
63–78 (1988). 75, 367–369 (1992). 535–540 (2014).

NAtuRe Reviews | NEuRosCiEnCE volume 20 | JANUARY 2019 | 61


Reviews

78. Mirmiran, M. & Ariagno, R. L. Influence of light in the with autistic disorder. Biol. Psychiatry 57, 134–138 expression not predicted from copy number. J. Med.
NICU on the development of circadian rhythms in (2005). Genet. 46, 86–93 (2009).
preterm infants. Semin. Perinatol. 24, 247–257 100. Nir, I. et al. Brief report: circadian melatonin, thyroid-​ 124. Nishimura, Y. et al. Genome-​wide expression profiling of
(2000). stimulating hormone, prolactin, and cortisol levels in lymphoblastoid cell lines distinguishes different forms
79. Boo, N. Y., Chee, S. C. & Rohana, J. Randomized serum of young adults with autism. J. Autism Dev. of autism and reveals shared pathways. Hum. Mol.
controlled study of the effects of different durations Disord. 25, 641–654 (1995). Genet. 16, 1682–1698 (2007).
of light exposure on weight gain by preterm infants in 101. Kulman, G. et al. Evidence of pineal endocrine 125. Dindot, S. V., Antalffy, B. A., Bhattacharjee, M. B.
a neonatal intensive care unit. Acta Paediatr. 91, hypofunction in autistic children. Neuro Endocrinol. & Beaudet, A. L. The Angelman syndrome ubiquitin
674–679 (2002). Lett. 21, 31–34 (2000). ligase localizes to the synapse and nucleus, and
80. Brandon, D. H., Holditch-​Davis, D. & Belyea, M. 102. Melke, J. et al. Abnormal melatonin synthesis in maternal deficiency results in abnormal dendritic
Preterm infants born at less than 31 weeks’ gestation autism spectrum disorders. Mol. Psychiatry 13, spine morphology. Hum. Mol. Genet. 17, 111–118
have improved growth in cycled light compared with 90–98 (2008). (2008).
continuous near darkness. J. Pediatr. 140, 192–199 103. Simonneaux, V. & Ribelayga, C. Generation of the 126. Yamasaki, K. et al. Neurons but not glial cells
(2002). melatonin endocrine message in mammals: a review of show reciprocal imprinting of sense and antisense
This randomized intervention study reports that the complex regulation of melatonin synthesis by transcripts of Ube3a. Hum. Mol. Genet. 12, 837–847
preterm infants (<31 weeks) receiving cycled light norepinephrine, peptides, and other pineal transmitters. (2003).
(11 hours on and 11 hours off, with 1 transition Pharmacol. Rev. 55, 325–395 (2003). 127. Shi, S. Q., Bichell, T. J., Ihrie, R. A. & Johnson, C. H.
hour for shift changes) during hospital care 104. Jin, Y., Choi, J., Won, J. & Hong, Y. The relationship Ube3a imprinting impairs circadian robustness in
(~4 weeks) gained weight faster than age-​matched between autism spectrum disorder and melatonin Angelman syndrome models. Curr. Biol. 25, 537–545
infants receiving near darkness. during fetal development. Molecules 23, E198 (2018). (2015).
81. Guyer, C. et al. Cycled light exposure reduces fussing 105. Gringras, P., Nir, T., Breddy, J., Frydman-​Marom, A. 128. Boudreau, E. A. et al. Review of disrupted sleep
and crying in very preterm infants. Pediatrics 130, & Findling, R. L. Efficacy and safety of pediatric patterns in Smith–Magenis syndrome and normal
e145–e151 (2012). prolonged-​release melatonin for insomnia in children melatonin secretion in a patient with an atypical
82. Hao, H. & Rivkees, S. A. The biological clock of very with autism spectrum disorder. J. Am. Acad. Child interstitial 17p11.2 deletion. Am. J. Med. Genet. A
premature primate infants is responsive to light. Adolesc. Psychiatry 56, 948–957 (2017). 149A, 1382–1391 (2009).
Proc. Natl Acad. Sci. USA 96, 2426–2429 (1999). 106. Tordjman, S. et al. Autism as a disorder of biological 129. Greenberg, F. et al. Multi-​disciplinary clinical study
83. Lebel, V., Aita, M., Johnston, C., Heon, M. & Dupuis, F. and behavioral rhythms: toward new therapeutic of Smith–Magenis syndrome (deletion 17p11.2).
Effects of cycled lighting versus continuous near perspectives. Front. Pediatr. 3, 1 (2015). Am. J. Med. Genet. 62, 247–254 (1996).
darkness on physiological stability and motor activity 107. Hoel, E. P., Albantakis, L., Cirelli, C. & Tononi, G. 130. Gropman, A. L., Elsea, S., Duncan, W. C. Jr
level in preterm infants. Adv. Neonatal Care 17, Synaptic refinement during development and its & Smith, A. C. New developments in Smith–Magenis
282–291 (2017). effect on slow-​wave activity: a computational study. syndrome (del 17p11.2). Curr. Opin. Neurol. 20,
84. Mann, N. P., Haddow, R., Stokes, L., Goodley, S. J. Neurophysiol. 115, 2199–2213 (2016). 125–134 (2007).
& Rutter, N. Effect of night and day on preterm infants 108. Goldstone, A. et al. The mediating role of cortical 131. Smith, A. C., Dykens, E. & Greenberg, F. Sleep
in a newborn nursery: randomised trial. Br. Med. thickness and gray matter volume on sleep slow-​wave disturbance in Smith–Magenis syndrome (del 17p11.2).
J. (Clin. Res. Ed.) 293, (1265–1267 (1986). activity during adolescence. Brain Struct. Funct. 223, Am. J. Med. Genet. 81, 186–191 (1998).
85. Rivkees, S. A., Mayes, L., Jacobs, H. & Gross, I. 669–685 (2017). 132. Boone, P. M. et al. Abnormal circadian rhythm of
Rest-​activity patterns of premature infants are 109. Cassidy, S. B., Schwartz, S., Miller, J. L. & Driscoll, D. J. melatonin in Smith–Magenis syndrome patients with
regulated by cycled lighting. Pediatrics 113, 833–839 Prader–Willi syndrome. Genet. Med. 14, 10–26 RAI1 point mutations. Am. J. Med. Genet. A 155A,
(2004). (2012). 2024–2027 (2011).
86. Philipsen, A., Hornyak, M. & Riemann, D. Sleep 110. Sahoo, T. et al. Prader–Willi phenotype caused by 133. Walz, K. et al. Behavioral characterization of mouse
and sleep disorders in adults with attention deficit/ paternal deficiency for the HBII-85C/D box small models for Smith–Magenis syndrome and dup(17)
hyperactivity disorder. Sleep Med. Rev. 10, 399–405 nucleolar RNA cluster. Nat. Genet. 40, 719–721 (p11.2p11.2). Hum. Mol. Genet. 13, 367–378
(2006). (2008). (2004).
87. Coogan, A. N. & McGowan, N. M. A systematic review 111. Lassi, G. et al. Deletion of the Snord116/SNORD116 134. Lacaria, M., Gu, W. & Lupski, J. R. Circadian
of circadian function, chronotype and chronotherapy alters sleep in mice and patients with Prader–Willi abnormalities in mouse models of Smith–Magenis
in attention deficit hyperactivity disorder. Atten. Defic. syndrome. Sleep 39, 637–644 (2016). syndrome: evidence for involvement of RAI1.
Hyperact. Disord. 9, 129–147 (2017). 112. Butler, J. V. et al. Prevalence of, and risk factors for, Am. J. Med. Genet. A 161A, 1561–1568 (2013).
88. Rybak, Y. E., McNeely, H. E., Mackenzie, B. E., Jain, U. R. physical ill-​health in people with Prader–Willi syndrome: 135. Williams, S. R., Zies, D., Mullegama, S. V.,
& Levitan, R. D. Seasonality and circadian preference in a population-​based study. Dev. Med. Child Neurol. 44, Grotewiel, M. S. & Elsea, S. H. Smith–Magenis
adult attention-​deficit/hyperactivity disorder: clinical 248–255 (2002). syndrome results in disruption of CLOCK gene
and neuropsychological correlates. Compr. Psychiatry 113. Cavaille, J. et al. Identification of brain-​specific and transcription and reveals an integral role for RAI1 in
48, 562–571 (2007). imprinted small nucleolar RNA genes exhibiting an the maintenance of circadian rhythmicity. Am. J. Hum.
89. Baird, A. L., Coogan, A. N., Siddiqui, A., Donev, R. M. unusual genomic organization. Proc. Natl Acad. Sci. Genet. 90, 941–949 (2012).
& Thome, J. Adult attention-​deficit hyperactivity USA 97, 14311–14316 (2000). 136. De Leersnyder, H. et al. β1-adrenergic antagonists
disorder is associated with alterations in circadian 114. Galiveti, C. R., Raabe, C. A., Konthur, Z. & improve sleep and behavioural disturbances in a
rhythms at the behavioural, endocrine and molecular Rozhdestvensky, T. S. Differential regulation of non-​ circadian disorder, Smith–Magenis syndrome. J. Med.
levels. Mol. Psychiatry 17, 988–995 (2012). protein coding RNAs from Prader–Willi syndrome Genet. 38, 586–590 (2001).
90. Fargason, R. E. et al. Correcting delayed circadian locus. Sci. Rep. 4, 6445 (2014). 137. De Leersnyder, H. et al. β1-adrenergic antagonists
phase with bright light therapy predicts improvement 115. Runte, M. et al. The IC-​SNURF-SNRPN transcript and melatonin reset the clock and restore sleep in a
in ADHD symptoms: a pilot study. J. Psychiatr. Res. serves as a host for multiple small nucleolar RNA circadian disorder, Smith–Magenis syndrome. J. Med.
91, 105–110 (2017). species and as an antisense RNA for UBE3A. Hum. Genet. 40, 74–78 (2003).
91. Molina-​Carballo, A. et al. Methylphenidate effects on Mol. Genet. 10, 2687–2700 (2001). 138. De Leersnyder, H., Claustrat, B., Munnich, A.
blood serotonin and melatonin levels may help to 116. Powell, W. T. et al. A Prader–Willi locus lncRNA cloud & Verloes, A. Circadian rhythm disorder in a rare
synchronise biological rhythms in children with ADHD. modulates diurnal genes and energy expenditure. disease: Smith–Magenis syndrome. Mol. Cell
J. Psychiatr. Res. 47, 377–383 (2013). Hum. Mol. Genet. 22, 4318–4328 (2013). Endocrinol. 252, 88–91 (2006).
92. Wang, S. M. et al. Modafinil for the treatment of 117. Gossan, N. C. et al. The E3 ubiquitin ligase UBE3A is 139. Elsea, S. H. & Girirajan, S. Smith–Magenis syndrome.
attention-​deficit/hyperactivity disorder: a meta-​analysis. an integral component of the molecular circadian clock Eur. J. Hum. Genet. 16, 412–421 (2008).
J. Psychiatr. Res. 84, 292–300 (2017). through regulating the BMAL1 transcription factor. 140. Logan, R. W. et al. Impact of sleep and circadian
93. Gerrard, P. & Malcolm, R. Mechanisms of modafinil: Nucleic Acids Res. 42, 5765–5775 (2014). rhythms on addiction vulnerability in adolescents. Biol.
a review of current research. Neuropsychiatr. Dis. Treat. 118. Solter, D. Differential imprinting and expression of Psychiatry 83, 987–996 (2017).
3, 349–364 (2007). maternal and paternal genomes. Annu. Rev. Genet. 141. Roenneberg, T. et al. A marker for the end of
94. Cortesi, F., Giannotti, F., Ivanenko, A. & Johnson, K. 22, 127–146 (1988). adolescence. Curr. Biol. 14, R1038–R1039 (2004).
Sleep in children with autistic spectrum disorder. 119. Martins-​Taylor, K. et al. Imprinted expression of 142. Hagenauer, M. H. & Lee, T. M. The neuroendocrine
Sleep Med. 11, 659–664 (2010). UBE3A in non-​neuronal cells from a Prader–Willi control of the circadian system: adolescent chronotype.
95. Richdale, A. L. & Schreck, K. A. Sleep problems in syndrome patient with an atypical deletion. Hum. Mol. Front. Neuroendocrinol. 33, 211–229 (2012).
autism spectrum disorders: prevalence, nature, and Genet. 23, 2364–2373 (2014). 143. Hagenauer, M. H. & Lee, T. M. Adolescent sleep
possible biopsychosocial aetiologies. Sleep Med. Rev. 120. Rougeulle, C., Glatt, H. & Lalande, M. The Angelman patterns in humans and laboratory animals. Horm.
13, 403–411 (2009). syndrome candidate gene, UBE3A/E6-AP, is imprinted Behav. 64, 270–279 (2013).
96. Giannotti, F. et al. An investigation of sleep in brain. Nat. Genet. 17, 14–15 (1997). 144. Cain, N. & Gradisar, M. Electronic media use and
characteristics, EEG abnormalities and epilepsy in 121. Baron, C. A. et al. Genomic and functional profiling of sleep in school-​aged children and adolescents:
developmentally regressed and non-​regressed children duplicated chromosome 15 cell lines reveal regulatory a review. Sleep Med. 11, 735–742 (2010).
with autism. J. Autism Dev. Disord. 38, 1888–1897 alterations in UBE3A-​associated ubiquitin-​proteasome 145. Crowley, S. J., Cain, S. W., Burns, A. C., Acebo, C.
(2008). pathway processes. Hum. Mol. Genet. 15, 853–869 & Carskadon, M. A. Increased sensitivity of the
97. Takase, M., Taira, M. & Sasaki, H. Sleep–wake rhythm (2006). circadian system to light in early/mid-​puberty. J. Clin.
of autistic children. Psychiatry Clin. Neurosci. 52, 122. Herzing, L. B., Cook, E. H. Jr & Ledbetter, D. H. Endocrinol. Metab. 100, 4067–4073 (2015).
181–182 (1998). Allele-​specific expression analysis by RNA-​FISH This study shows that endogenous melatonin levels
98. Hayashi, E. Seasonal changes in sleep and behavioral demonstrates preferential maternal expression of of prepubertal children and early adolescents are
problems in a pubescent case with autism. Psychiatry UBE3A and imprint maintenance within 15q11-q13 more sensitive to the suppressive effects of acute
Clin. Neurosci. 55, 223–224 (2001). duplications. Hum. Mol. Genet. 11, 1707–1718 light exposure during the evening than melatonin
99. Tordjman, S., Anderson, G. M., Pichard, N., (2002). levels in late adolescents.
Charbuy, H. & Touitou, Y. Nocturnal excretion of 123. Hogart, A. et al. Chromosome 15q11-13 duplication 146. Paruthi, S. et al. Recommended amount of sleep for
6-sulphatoxymelatonin in children and adolescents syndrome brain reveals epigenetic alterations in gene pediatric populations: a consensus statement of the

62 | JANUARY 2019 | volume 20 www.nature.com/nrn


Reviews

American Academy of Sleep Medicine. J. Clin. Sleep 169. Forbes, E. E. et al. Healthy adolescents’ neural circadian rhythm in bipolar disorder: a systematic
Med. 12, 785–786 (2016). response to reward: associations with puberty, positive review. Sleep Med. Rev. 34, 46–58 (2017).
147. Basch, C. E., Basch, C. H., Ruggles, K. V. & Rajan, S. affect, and depressive symptoms. J. Am. Acad. Child 192. Winthorst, W. H. et al. Seasonal affective disorder
Prevalence of sleep duration on an average school Adolesc. Psychiatry 49, 162–172 (2010). and non-​seasonal affective disorders: results from the
night among 4 nationally representative successive 170. Padmanabhan, A., Geier, C. F., Ordaz, S. J., Teslovich, T. NESDA study. BJPsych Open 3, 196–203 (2017).
samples of American high school students, & Luna, B. Developmental changes in brain function 193. Medici, C. R., Vestergaard, C. H., Hadzi-​Pavlovic, D.,
2007–2013. Prev. Chron. Dis. 11, E216 (2014). underlying the influence of reward processing on Munk-​Jorgensen, P. & Parker, G. Seasonal variations
148. Touitou, Y. Adolescent sleep misalignment: a chronic inhibitory control. Dev. Cogn. Neurosci. 1, 517–529 in hospital admissions for mania: examining for
jet lag and a matter of public health. J. Physiol. Paris (2011). associations with weather variables over time. J. Affect.
107, 323–326 (2013). 171. Frischknecht, R. & Gundelfinger, E. D. The brain’s Disord. 205, 81–86 (2016).
149. Hasler, B. P. et al. Weekend-​weekday advances in sleep extracellular matrix and its role in synaptic plasticity. 194. Young, J. W. & Dulcis, D. Investigating the mechanism(s)
timing are associated with altered reward-​related Adv. Exp. Med. Biol. 970, 153–171 (2012). underlying switching between states in bipolar disorder.
brain function in healthy adolescents. Biol. Psychol. 172. Luciana, M. & Collins, P. F. Incentive motivation, Eur. J. Pharmacol. 759, 151–162 (2015).
91, 334–341 (2012). cognitive control, and the adolescent brain: is it time 195. Raniti, M. B. et al. Sleep duration and sleep quality:
150. Skeldon, A. C., Phillips, A. J. & Dijk, D. J. The effects of for a paradigm shift? Child Dev. Perspect. 6, 392–399 associations with depressive symptoms across
self-​selected light–dark cycles and social constraints (2012). adolescence. Behav. Sleep Med. 15, 198–215 (2017).
on human sleep and circadian timing: a modeling 173. Padmanabhan, A. & Luna, B. Developmental imaging 196. Lewy, A. J. Depressive disorders may more commonly
approach. Sci. Rep. 7, 45158 (2017). genetics: linking dopamine function to adolescent be related to circadian phase delays rather than
151. Crowley, S. J. & Carskadon, M. A. Modifications to behavior. Brain Cogn. 89, 27–38 (2014). advances: time will tell. Sleep Med. 11, 117–118
weekend recovery sleep delay circadian phase in older 174. Ernst, M. & Fudge, J. L. A developmental (2010).
adolescents. Chronobiol. Int. 27, 1469–1492 (2010). neurobiological model of motivated behavior: anatomy, 197. Addington, J. & Heinssen, R. Prediction and prevention
152. Wittmann, M., Dinich, J., Merrow, M. & Roenneberg, T. connectivity and ontogeny of the triadic nodes. of psychosis in youth at clinical high risk. Annu. Rev.
Social jetlag: misalignment of biological and social Neurosci. Biobehav. Rev. 33, 367–382 (2009). Clin. Psychol. 8, 269–289 (2012).
time. Chronobiol. Int. 23, 497–509 (2006). 175. Somerville, L. H., Jones, R. M. & Casey, B. J. A time of 198. Lunsford-​Avery, J. R. et al. Adolescents at clinical-​
153. Chen, K. & Kandel, D. B. The natural history of drug change: behavioral and neural correlates of adolescent high risk for psychosis: circadian rhythm disturbances
use from adolescence to the mid-​thirties in a general sensitivity to appetitive and aversive environmental predict worsened prognosis at 1-year follow-​up.
population sample. Am. J. Publ. Health 85, 41–47 cues. Brain Cogn. 72, 124–133 (2010). Schizophr. Res. 189, 37–42 (2017).
(1995). 176. Logan, R. W., Williams, W. P. & McClung, C. A. 199. Robillard, R. et al. Sleep–wake profiles predict
154. Falcon, E. & McClung, C. A. A role for the circadian Circadian rhythms and addiction: mechanistic longitudinal changes in manic symptoms and memory
genes in drug addiction. Neuropharmacology 56, insights and future directions. Behav. Neurosci. 128, in young people with mood disorders. J. Sleep Res.
91–96 (2009). (Suppl. 1). 387–412 (2014). 25, 549–555 (2016).
155. O’Brien, E. M. & Mindell, J. A. Sleep and risk-​taking 177. Hasler, B. P., Forbes, E. E. & Franzen, P. L. Time-​of-day 200. Crews, F. T., Vetreno, R. P., Broadwater, M. A.
behavior in adolescents. Behav. Sleep Med. 3, differences and short-​term stability of the neural & Robinson, D. L. Adolescent alcohol exposure
113–133 (2005). response to monetary reward: a pilot study. Psychiatry persistently impacts adult neurobiology and behavior.
156. Pasch, K. E., Laska, M. N., Lytle, L. A. & Moe, S. G. Res. 224, 22–27 (2014). Pharmacol. Rev. 68, 1074–1109 (2016).
Adolescent sleep, risk behaviors, and depressive 178. Hasler, B. P. et al. Chronotype and diurnal patterns of 201. Wei, Y., Krishnan, G. P. & Bazhenov, M. Synaptic
symptoms: are they linked? Am. J. Health Behav. 34, positive affect and affective neural circuitry in primary mechanisms of memory consolidation during sleep
237–248 (2010). insomnia. J. Sleep Res. 21, 515–526 (2012). slow oscillations. J. Neurosci. 36, 4231–4247 (2016).
157. McKnight-​Eily, L. R. et al. Relationships between hours 179. Hasler, B. P., Sitnick, S. L., Shaw, D. S. & Forbes, E. E. 202. Walker, M. P. & Stickgold, R. Sleep-​dependent
of sleep and health-​risk behaviors in US adolescent An altered neural response to reward may contribute learning and memory consolidation. Neuron 44,
students. Prev. Med. 53, 271–273 (2011). to alcohol problems among late adolescents with an 121–133 (2004).
158. Paiva, T., Gaspar, T. & Matos, M. G. Mutual relations evening chronotype. Psychiatry Res. 214, 357–364 203. Havekes, R., Meerlo, P. & Abel, T. Animal studies on
between sleep deprivation, sleep stealers and risk (2013). the role of sleep in memory: from behavioral
behaviours in adolescents. Sleep Sci. 9, 7–13 (2016). 180. Mullin, B. C. et al. Sleep deprivation amplifies striatal performance to molecular mechanisms. Curr. Top.
159. Sivertsen, B., Skogen, J. C., Jakobsen, R. & Hysing, M. activation to monetary reward. Psychol. Med. 43, Behav. Neurosci. 25, 183–206 (2015).
Sleep and use of alcohol and drug in adolescence. 2215–2225 (2013). 204. Marshall, L., Helgadottir, H., Molle, M. & Born, J.
A large population-​based study of Norwegian 181. Miller, M. B., Donahue, M. L., Carey, K. B. Boosting slow oscillations during sleep potentiates
adolescents aged 16 to 19 years. Drug Alcohol & Scott-​Sheldon, L. A. J. Insomnia treatment in the memory. Nature 444, 610–613 (2006).
Depend. 149, 180–186 (2015). context of alcohol use disorder: a systematic review and This study establishes the importance of slow-​wave
160. Thomas, A. G., Monahan, K. C., Lukowski, A. F. meta-​analysis. Drug Alcohol Depend. 181, 200–207 sleep during non-​REM episodes that occur during
& Cauffman, E. Sleep problems across development: (2017). the early night and shows that it promotes the
a pathway to adolescent risk taking through working 182. Merikangas, K. R. et al. Lifetime prevalence of mental retention of hippocampus-​dependent declarative
memory. J. Youth Adolesc. 44, 447–464 (2015). disorders in U. S. adolescents: results from the memories.
161. Wheaton, A. G., Olsen, E. O., Miller, G. F. & Croft, J. B. National Comorbidity Survey Replication — Adolescent 205. Selemon, L. D. & Zecevic, N. Schizophrenia: a tale of
Sleep duration and injury-​related risk behaviors among Supplement (NCS-​A). J. Am. Acad. Child Adolesc. two critical periods for prefrontal cortical development.
high school students — United States, 2007–2013. Psychiatry 49, 980–989 (2010). Transl Psychiatry 5, e623 (2015).
MMWR Morb. Mortal. Wkly Rep. 65, 337–341 183. Lamont, E. W., Coutu, D. L., Cermakian, N. 206. Keshavan, M. S., Anderson, S. & Pettegrew, J. W.
(2016). & Boivin, D. B. Circadian rhythms and clock genes in Is schizophrenia due to excessive synaptic pruning in
162. Pasch, K. E., Latimer, L. A., Cance, J. D., Moe, S. G. psychotic disorders. Isr. J. Psychiatry Relat. Sci. 47, the prefrontal cortex? The Feinberg hypothesis
& Lytle, L. A. Longitudinal bi-​directional relationships 27–35 (2010). revisited. J. Psychiatr. Res. 28, 239–265 (1994).
between sleep and youth substance use. J. Youth 184. McClung, C. A. How might circadian rhythms control 207. Billeh, Y. N. et al. Effects of chronic sleep restriction
Adolesc. 41, 1184–1196 (2012). mood? Let me count the ways. Biol. Psychiatry 74, during early adolescence on the adult pattern of
163. Hasler, B. P., Martin, C. S., Wood, D. S., Rosario, B. 242–249 (2013). connectivity of mouse secondary motor cortex.
& Clark, D. B. A longitudinal study of insomnia 185. Mansour, H. A. et al. Circadian phase variation in eNeuro https://doi.org/10.1523/ENEURO.0053-
and other sleep complaints in adolescents with and bipolar I disorder. Chronobiol. Int. 22, 571–584 16-2016 (2016).
without alcohol use disorders. Alcohol Clin. Exp. Res. (2005). 208. Carskadon, M. A. Patterns of sleep and sleepiness in
38, 2225–2233 (2014). 186. McCarthy, M. J. & Welsh, D. K. Cellular circadian adolescents. Pediatrician 17, 5–12 (1990).
164. Hasler, B. P., Kirisci, L. & Clark, D. B. Restless sleep and clocks in mood disorders. J. Biol. Rhythms 27, 209. Johnson, E. O., Roth, T., Schultz, L. & Breslau, N.
variable sleep timing during late childhood accelerate 339–352 (2012). Epidemiology of DSM-​IV insomnia in adolescence:
the onset of alcohol and other drug involvement. 187. McClung, C. A. Circadian rhythms and mood lifetime prevalence, chronicity, and an emergent gender
J. Stud. Alcohol Drugs 77, 649–655 (2016). regulation: insights from pre-​clinical models. Eur. difference. Pediatrics 117, e247–e256 (2006).
165. Hasler, B. P., Soehner, A. M. & Clark, D. B. Sleep and Neuropsychopharmacol. 21, S683–S693 (2011). 210. Sivertsen, B. et al. Delayed sleep phase syndrome in
circadian contributions to adolescent alcohol use 188. Wulff, K., Dijk, D. J., Middleton, B., Foster, R. G. adolescents: prevalence and correlates in a large
disorder. Alcohol 49, 377–387 (2015). & Joyce, E. M. Sleep and circadian rhythm disruption population based study. BMC Publ. Health 13, 1163
166. Tavernier, R., Munroe, M. & Willoughby, T. Perceived in schizophrenia. Br. J. Psychiatry 200, 308–316 (2013).
morningness-​eveningness predicts academic (2012). 211. Owens, J. A., Belon, K. & Moss, P. Impact of delaying
adjustment and substance use across university, but This study reports notable circadian misalignment school start time on adolescent sleep, mood, and
social jetlag is not to blame. Chronobiol. Int. 32, (phase delays and advances) of sleep–wake behavior. Arch. Pediatr. Adolesc. Med. 164, 608–614
1233–1245 (2015). cycles and melatonin rhythms in patients with (2010).
167. Hasler, B. P., Casement, M. D., Sitnick, S. L., Shaw, D. S. schizophrenia despite other factors, including 212. Wirz-​Justice, A. & Terman, M. Chronotherapeutics
& Forbes, E. E. Eveningness among late adolescent mood, cognitive status and pharmacological (light and wake therapy) as a class of interventions
males predicts neural reactivity to reward and alcohol treatments. for affective disorders. Handb. Clin. Neurol. 106,
dependence two years later. Behav. Brain Res. 327, 189. Frank, E., Swartz, H. A. & Kupfer, D. J. Interpersonal 697–713 (2012).
112–120 (2017). and social rhythm therapy: managing the chaos of 213. Srinivasan, V., De Berardis, D., Shillcutt, S. D.
This study shows that an evening chronotype bipolar disorder. Biol. Psychiatry 48, 593–604 & Brzezinski, A. Role of melatonin in mood disorders
during late adolescence (20 years of age) is (2000). and the antidepressant effects of agomelatine. Expert
associated with strength of activation in the VS to 190. Malkoff-​Schwartz, S. et al. Stressful life events and social Opin. Investig. Drugs 21, 1503–1522 (2012).
reward outcome, which was positively correlated rhythm disruption in the onset of manic and depressive 214. Wu, J. C. et al. Rapid and sustained antidepressant
with alcohol dependence 2 years later. bipolar episodes: a preliminary investigation. Arch. Gen. response with sleep deprivation and chronotherapy in
168. Mednick, S. C., Christakis, N. A. & Fowler, J. H. Psychiatry 55, 702–707 (1998). bipolar disorder. Biol. Psychiatry 66, 298–301 (2009).
The spread of sleep loss influences drug use in 191. Melo, M. C. A., Abreu, R. L. C., Linhares Neto, V. B., This study demonstrates, in a cohort of 49 patients
adolescent social networks. PLOS ONE 5, e9775 (2010). de Bruin, P. F. C. & de Bruin, V. M. S. Chronotype and with bipolar disorder, the ability of non-​invasive

NAtuRe Reviews | NEuRosCiEnCE volume 20 | JANUARY 2019 | 63


Reviews

circadian therapy (that is, bright light therapy and 237. Karasek, M. Melatonin, human aging, and age-​related 260. Rothman, S. M. & Mattson, M. P. Sleep disturbances in
sleep phase advance) and sleep (that is, acute diseases. Exp. Gerontol. 39, 1723–1729 (2004). Alzheimer’s and Parkinson’s diseases. Neuromolecular
sleep deprivation) to improve and sustain the 238. Zeitzer, J. M. et al. Do plasma melatonin Med. 14, 194–204 (2012).
therapeutic efficacy of lithium and antidepressants concentrations decline with age? Am. J. Med. 107, 261. Vitiello, M. V., Prinz, P. N., Williams, D. E.,
on mood symptoms. 432–436 (1999). Frommlet, M. S. & Ries, R. K. Sleep disturbances
215. Li, J., Lu, W. Q., Beesley, S., Loudon, A. S. & Meng, Q. J. 239. Czeisler, C. A. et al. Association of sleep–wake habits in patients with mild-​stage Alzheimer’s disease.
Lithium impacts on the amplitude and period of the in older people with changes in output of circadian J. Gerontol. 45, M131–M138 (1990).
molecular circadian clockwork. PLOS ONE 7, e33292 pacemaker. Lancet 340, 933–936 (1992). 262. Stopa, E. G. et al. Pathologic evaluation of the
(2012). This landmark study suggests that changes in sleep human suprachiasmatic nucleus in severe dementia.
216. Johansson, A. S., Brask, J., Owe-​Larsson, B., Hetta, J. timing and consolidation driven by the circadian J. Neuropathol. Exp. Neurol. 58, 29–39 (1999).
& Lundkvist, G. B. Valproic acid phase shifts the system underlie sleep disturbances in elderly 263. Cermakian, N., Lamont, E. W., Boudreau, P.
rhythmic expression of Period2::Luciferase. J. Biol. individuals. & Boivin, D. B. Circadian clock gene expression in brain
Rhythms 26, 541–551 (2011). 240. Wang, J. L. et al. Suprachiasmatic neuron numbers regions of Alzheimer ‘s disease patients and control
217. Sprouse, J., Braselton, J. & Reynolds, L. Fluoxetine and rest–activity circadian rhythms in older humans. subjects. J. Biol. Rhythms 26, 160–170 (2011).
modulates the circadian biological clock via phase Ann. Neurol. 78, 317–322 (2015). 264. Logan, R. W. et al. NAD+ cellular redox and SIRT1
advances of suprachiasmatic nucleus neuronal firing. 241. Chang, H. C. & Guarente, L. SIRT1 mediates central regulate the diurnal rhythms of tyrosine hydroxylase
Biol. Psychiatry 60, 896–899 (2006). circadian control in the SCN by a mechanism that and conditioned cocaine reward. Mol. Psychiatry.
218. Duncan, W. C. Jr et al. Motor-​activity markers of decays with aging. Cell 153, 1448–1460 (2013). https://doi.org/10.1038/s41380-018-0061-1 (2018).
circadian timekeeping are related to ketamine’s This mouse study reveals that reductions of sirtuin This study demonstrates mechanistic links between
rapid antidepressant properties. Biol. Psychiatry 82, activity in the SCN leads to a gradual age-​dependent circadian regulation of SIRT1-dependent metabolic
361–369 (2017). dampening of SCN activity and locomotor rhythms. signalling and dopaminergic neurotransmission in
In a study of 51 patients with major depressive 242. Yamazaki, S. et al. Effects of aging on central and the mouse ventral tegmental area, which is
disorder or bipolar disorder, circadian activity peripheral mammalian clocks. Proc. Natl Acad. Sci. important for drug reward.
patterns prior to and following the administration USA 99, 10801–10806 (2002). 265. Belaidi, A. A. et al. Marked age-​related changes in
of ketamine are linked to therapeutic response of 243. Swaab, D. F., Fliers, E. & Partiman, T. S. The brain iron homeostasis in amyloid protein precursor
mood symptoms. suprachiasmatic nucleus of the human brain in knockout mice. Neurotherapeutics https://doi.org/
219. Kragh, M. et al. Predictors of response to combined relation to sex, age and senile dementia. Brain Res. 10.1007/s13311-018-0656-x (2018).
wake and light therapy in treatment-​resistant inpatients 342, 37–44 (1985). 266. Uranga, R. M. & Salvador, G. A. Unraveling the burden
with depression. Chronobiol. Int. https://doi.org/ 244. Zhou, J. N., Hofman, M. A. & Swaab, D. F. VIP of iron in neurodegeneration: intersections with
10.1080/07420528.2018.1468341 (2018). neurons in the human SCN in relation to sex, age, and amyloid beta peptide pathology. Oxid. Med. Cell.
220. Adan, A. & Natale, V. Gender differences in Alzheimer’s disease. Neurobiol. Aging 16, 571–576 Longev. 2018, 2850341 (2018).
morningness–eveningness preference. Chronobiol. Int. (1995). 267. Graham, D. G. Oxidative pathways for catecholamines
19, 709–720 (2002). 245. Roozendaal, B., van Gool, W. A., Swaab, D. F., in the genesis of neuromelanin and cytotoxic quinones.
221. Caci, H., Deschaux, O., Adan, A. & Natale, V. Hoogendijk, J. E. & Mirmiran, M. Changes in Mol. Pharmacol. 14, 633–643 (1978).
Comparing three morningness scales: age and gender vasopressin cells of the rat suprachiasmatic 268. Tse, D. C., McCreery, R. L. & Adams, R. N. Potential
effects, structure and cut-​off criteria. Sleep Med. 10, nucleus with aging. Brain Res. 409, 259–264 oxidative pathways of brain catecholamines. J. Med.
240–245 (2009). (1987). Chem. 19, 37–40 (1976).
222. Duarte, L. L. et al. Chronotype ontogeny related 246. Tsukahara, S., Tanaka, S., Ishida, K., Hoshi, N. 269. Burbulla, L. F. et al. Dopamine oxidation mediates
to gender. Braz. J. Med. Biol. Res. 47, 316–320 & Kitagawa, H. Age-​related change and its sex mitochondrial and lysosomal dysfunction in Parkinson’s
(2014). differences in histoarchitecture of the hypothalamic disease. Science 357, 1255–1261 (2017).
223. Paine, S. J., Gander, P. H. & Travier, N. suprachiasmatic nucleus of F344/N rats. Exp. Gerontol. This study demonstrates that mitochondrial stress
The epidemiology of morningness/eveningness: 40, 147–155 (2005). promotes the accumulation of oxidized dopamine,
influence of age, gender, ethnicity, and socioeconomic 247. Nygard, M., Hill, R. H., Wikstrom, M. A. consequently increasing α-​synuclein aggregates in
factors in adults (30–49 years). J. Biol. Rhythms 21, & Kristensson, K. Age-​related changes in electro­ neurons derived from patients with PD, an effect
68–76 (2006). physiological properties of the mouse supra­ selective to human neurons.
224. Patke, A. et al. Mutation of the human circadian clock chiasmatic nucleus in vitro. Brain Res. Bull. 65, 270. Barone, P. et al. The PRIAMO study: a multicenter
gene CRY1 in familial delayed sleep phase disorder. 149–154 (2005). assessment of nonmotor symptoms and their impact
Cell 169, 203–215 (2017). 248. Farajnia, S. et al. Evidence for neuronal desynchrony in on quality of life in Parkinson’s disease. Mov. Disord.
225. Kripke, D. F. et al. Circadian polymorphisms in night the aged suprachiasmatic nucleus clock. J. Neurosci. 24, 1641–1649 (2009).
owls, in bipolars, and in non-24-hour sleep cycles. 32, 5891–5899 (2012). 271. Chaudhuri, K. R. et al. International multicenter pilot
Psychiatry Invest. 11, 345–362 (2014). 249. Nakamura, T. J. et al. Age-​related decline in circadian study of the first comprehensive self-​completed
226. Murray, J. M. et al. Prevalence of circadian output. J. Neurosci. 31, 10201–10205 (2011). nonmotor symptoms questionnaire for Parkinson’s
misalignment and its association with depressive 250. Tranah, G. J. et al. Circadian activity rhythms and risk disease: the NMSQuest study. Mov. Disord. 21,
symptoms in delayed sleep phase disorder. Sleep 40, of incident dementia and mild cognitive impairment in 916–923 (2006).
zsw002 (2017). older women. Ann. Neurol. 70, 722–732 (2011). 272. Goodman, A. O., Morton, A. J. & Barker, R. A.
227. James, S. M., Honn, K. A., Gaddameedhi, S. 251. Chen, H. F., Huang, C. Q., You, C., Wang, Z. R. Identifying sleep disturbances in Huntington’s disease
& Van Dongen, H. P. A. Shift work: disrupted circadian & Si-​qing, H. Polymorphism of CLOCK gene using a simple disease-​focused questionnaire. PLOS
rhythms and sleep-​implications for health and rs 4580704C>G is associated with susceptibility of Curr. 2, RRN1189 (2010).
well-​being. Curr. Sleep Med. Rep. 3, 104–112 (2017). Alzheimer’s disease in a Chinese population. Arch. 273. Aziz, N. A., Anguelova, G. V., Marinus, J., Lammers, G. J.
228. Hall, A. L., Franche, R. L. & Koehoorn, M. Examining Med. Res. 44, 203–207 (2013). & Roos, R. A. Sleep and circadian rhythm alterations
exposure assessment in shift work research: a study 252. Chen, Q., Peng, X. D., Huang, C. Q., Hu, X. Y. correlate with depression and cognitive impairment in
on depression among nurses. Ann. Work Expo. Health & Zhang, X. M. Association between ARNTL (BMAL1) Huntington’s disease. Parkinsonism Relat. Disord. 16,
62, 182–194 (2018). rs2278749 polymorphism T>C and susceptibility to 345–350 (2010).
229. Kang, M. Y., Kwon, H. J., Choi, K. H., Kang, C. W. Alzheimer disease in a Chinese population. Genet. 274. Iranzo, A. et al. Neurodegenerative disorder risk in
& Kim, H. The relationship between shift work Mol. Res. 14, 18515–18522 (2015). idiopathic REM sleep behavior disorder: study in 174
and mental health among electronics workers in 253. Gu, Z. et al. Association of ARNTL and PER1 genes patients. PLOS ONE 9, e89741 (2014).
South Korea: a cross-​sectional study. PLOS ONE 12, with Parkinson’s disease: a case-​control study of Han 275. Schenck, C. H., Boeve, B. F. & Mahowald, M. W.
e0188019 (2017). Chinese. Sci. Rep. 5, 15891 (2015). Delayed emergence of a parkinsonian disorder or
230. McNeely, E. et al. The self-​reported health of U. S. 254. Videnovic, A. & Golombek, D. Circadian dysregulation dementia in 81% of older men initially diagnosed with
flight attendants compared to the general population. in Parkinson’s disease. Neurobiol. Sleep Circadian idiopathic rapid eye movement sleep behavior
Environ. Health 13, 13 (2014). Rhythms 2, 53–58 (2017). disorder: a 16-year update on a previously reported
231. McNeely, E. et al. Cancer prevalence among flight 255. Breen, D. P. et al. Sleep and circadian rhythm series. Sleep Med. 14, 744–748 (2013).
attendants compared to the general population. regulation in early Parkinson disease. JAMA Neurol. 276. Kudo, T., Loh, D. H., Truong, D., Wu, Y. & Colwell, C. S.
Environ. Health 17, 49 (2018). 71, 589–595 (2014). Circadian dysfunction in a mouse model of Parkinson’s
232. Grajewski, B., Whelan, E. A., Nguyen, M. M., Kwan, L. This study shows significantly increased sleep disease. Exp. Neurol. 232, 66–75 (2011).
& Cole, R. J. Sleep disturbance in female flight latency, reduced sleep efficiency and reduced REM, 277. Grippo, R. M., Purohit, A. M., Zhang, Q., Zweifel, L. S.
attendants and teachers. Aerosp. Med. Hum. Perform. along with reduced melatonin levels and a loss of & Guler, A. D. Direct midbrain dopamine input to the
87, 638–645 (2016). BMAL1 serum expression rhythms, in a cohort suprachiasmatic nucleus accelerates circadian
233. McNeely, E., Mordukhovich, I., Tideman, S., Gale, S. of 239 individuals with PD. entrainment. Curr. Biol. 27, 2465–2475 (2017).
& Coull, B. Estimating the health consequences of 256. Gros, P. & Videnovic, A. Sleep and circadian rhythm This study shows re-​entrainment of the SCN to
flight attendant work: comparing flight attendant disorders in Parkinson’s disease. Curr. Sleep Med. phase shifts in the light–dark schedule is facilitated
health to the general population in a cross-​sectional Rep. 3, 222–234 (2017). by dopaminergic projections from the ventral
study. BMC Publ. Health 18, 346 (2018). 257. Uchida, K., Okamoto, N., Ohara, K. & Morita, Y. tegmental area to the SCN and is dependent
234. Feinsilver, S. H. & Hernandez, A. B. Sleep in the Daily rhythm of serum melatonin in patients with on dopamine 1 receptor-​expressing neurons in
elderly: unanswered questions. Clin. Geriatr. Med. 33, dementia of the degenerate type. Brain Res. 717, the SCN.
579–596 (2017). 154–159 (1996). 278. Korshunov, K. S., Blakemore, L. J. & Trombley, P. Q.
235. Mattis, J. & Sehgal, A. Circadian rhythms, sleep, and 258. Videnovic, A. et al. Circadian melatonin rhythm and Dopamine: a modulator of circadian rhythms in the
disorders of aging. Trends Endocrinol. Metab. 27, excessive daytime sleepiness in Parkinson disease. central nervous system. Front. Cell. Neurosci. 11, 91
192–203 (2016). JAMA Neurol. 71, 463–469 (2014). (2017).
236. Musiek, E. S. & Holtzman, D. M. Mechanisms linking 259. Claassen, D. O. & Kutscher, S. J. Sleep disturbances 279. Fifel, K. & Cooper, H. M. Loss of dopamine disrupts
circadian clocks, sleep, and neurodegeneration. in Parkinson’s disease patients and management circadian rhythms in a mouse model of Parkinson’s
Science 354, 1004–1008 (2016). options. Nat. Sci. Sleep 3, 125–133 (2011). disease. Neurobiol. Dis. 71, 359–369 (2014).

64 | JANUARY 2019 | volume 20 www.nature.com/nrn


Reviews

280. Gnanasekaran, G. “Sundowning” as a biological 300. Figueiro, M. G. et al. Tailored lighting intervention 320. Jones, S. E. et al. Genome-​wide association analyses in
phenomenon: current understandings and future improves measures of sleep, depression, and agitation 128,266 individuals identifies new morningness and
directions: an update. Aging Clin. Exp. Res. 28, in persons with Alzheimer’s disease and related sleep duration loci. PLOS Genet. 12, e1006125 (2016).
383–392 (2016). dementia living in long-​term care facilities. Clin. Interv. 321. Lane, J. M. et al. Genome-​wide association analysis
281. Bedrosian, T. A. & Nelson, R. J. Sundowning syndrome Aging 9, 1527–1537 (2014). identifies novel loci for chronotype in 100,420
in aging and dementia: research in mouse models. 301. Riemersma-​van der Lek, R. F. et al. Effect of bright individuals from the UK Biobank. Nat. Commun. 7,
Exp. Neurol. 243, 67–73 (2013). light and melatonin on cognitive and noncognitive 10889 (2016).
282. Volicer, L., Harper, D. G., Manning, B. C., Goldstein, R. function in elderly residents of group care facilities: a 322. LeGates, T. A., Fernandez, D. C. & Hattar, S. Light as a
& Satlin, A. Sundowning and circadian rhythms in randomized controlled trial. JAMA 299, 2642–2655 central modulator of circadian rhythms, sleep and
Alzheimer’s disease. Am. J. Psychiatry 158, 704–711 (2008). affect. Nat. Rev. Neurosci. 15, 443–454 (2014).
(2001). This double-​blinded randomized controlled trial 323. LeGates, T. A. et al. Aberrant light directly impairs
This study shows that patients with AD have less of 189 elderly residents of a group care facility mood and learning through melanopsin-​expressing
diurnal activity, more nocturnal activity and phase demonstrates that a combination of melatonin neurons. Nature 491, 594–598 (2012).
delays in body-​temperature rhythms, and that supplementation and scheduled light exposure 324. Fernandez, D. C. et al. Light affects mood and learning
the severity of sundowning is positively correlated improves cognitive symptoms of dementia and through distinct retina-​brain pathways. Cell 175,
with lower-​amplitude and more-​phase-delayed reduces aggression. 71–84 (2018).
rhythms. 302. Singer, C. et al. A multicenter, placebo-​controlled trial Using mice, this study demonstrates that light
283. Todd, W. D. et al. A hypothalamic circuit for the of melatonin for sleep disturbance in Alzheimer’s input to the brain reaches neural circuits related
circadian control of aggression. Nat. Neurosci. 21, disease. Sleep 26, 893–901 (2003). to depression and learning through distinct
717–724 (2018). 303. Coogan, A. N. et al. The circadian system in Alzheimer’s projections from the retina, revealing a potential
This chemogenetic study in mice reveals a neural disease: disturbances, mechanisms, and opportunities. mechanism by which aberrant light exposure (for
circuit involving projections from the SCN to Biol. Psychiatry 74, 333–339 (2013). example, light at night) affects mood and cognition.
the ventromedial hypothalamus that drives 304. Arey, R. N. et al. An important role for cholecystokinin, 325. Librodo, P., Buckley, M., Luk, M. & Bisso, A.
time-​of-day-​dependent aggressive behaviours. a CLOCK target gene, in the development and Chronotherapeutic drug delivery. J. Infus. Nurs. 38,
284. Shokri-​Kojori, E. et al. β-​Amyloid accumulation in the treatment of manic-​like behaviors. Mol. Psychiatry S18–S23 (2015).
human brain after one night of sleep deprivation. 19, 342–350 (2014). 326. Ruben, M. D. et al. A database of tissue-​specific
Proc. Natl Acad. Sci. USA 115, 4483–4488 (2018). 305. Coque, L. et al. Specific role of VTA dopamine rhythmically expressed human genes has potential
285. Xie, L. et al. Sleep drives metabolite clearance from neuronal firing rates and morphology in the reversal applications in circadian medicine. Sci. Transl Med.
the adult brain. Science 342, 373–377 (2013). of anxiety-​related, but not depression-​related 10, eaat8806 (2018).
A landmark study in mice demonstrating that behavior in the Clock∆19 mouse model of mania. 327. Melkani, G. C. & Panda, S. Time-​restricted feeding
wakefulness suppresses the outflow of cerebral Neuropsychopharmacology 36, 1478–1488 for prevention and treatment of cardiometabolic
spinal fluid from the brain, whereas sleep leads to (2011). disorders. J. Physiol. 595, 3691–3700 (2017).
an increase in convective fluxes of fluid, effectively 306. Dzirasa, K. et al. Lithium ameliorates nucleus 328. Mindikoglu, A. L., Opekun, A. R., Gagan, S. K.
promoting the clearance of neurotoxic metabolics, accumbens phase-​signaling dysfunction in a genetic & Devaraj, S. Impact of time-​restricted feeding and
including Aβ. mouse model of mania. J. Neurosci. 30, 16314–16323 dawn-​to-sunset fasting on circadian rhythm, obesity,
286. Da Mesquita, S. et al. Functional aspects of meningeal (2010). metabolic syndrome, and nonalcoholic fatty liver
lymphatics in ageing and Alzheimer’s disease. Nature 307. McClung, C. A. Circadian rhythms, the mesolimbic disease. Gastroenterol. Res. Pract. 2017, 3932491
560, 185–191 (2018). dopaminergic circuit, and drug addiction. (2017).
287. Kang, J. E. et al. Amyloid-​β dynamics are regulated ScientificWorldJournal 7, 194–202 (2007). 329. Stewart, K. T., McEachron, D. L., Rosenwasser, A. M.
by orexin and the sleep-​wake cycle. Science 326, 308. McClung, C. A. et al. Regulation of dopaminergic & Adler, N. T. Lithium lengthens circadian period but
1005–1007 (2009). transmission and cocaine reward by the Clock gene. fails to counteract behavioral helplessness in rats.
288. Tabuchi, M. et al. Sleep interacts with Aβ to modulate Proc. Natl Acad. Sci. USA 102, 9377–9381 (2005). Biol. Psychiatry 30, 515–518 (1991).
intrinsic neuronal excitability. Curr. Biol. 25, 702–712 309. Ozburn, A. R. et al. Direct regulation of diurnal Drd3 330. McCarthy, M. J. et al. Genetic and clinical factors
(2015). expression and cocaine reward by NPAS2. Biol. predict lithium’s effects on PER2 gene expression
289. Faghih, M. M. & Sharp, M. K. Is bulk flow plausible in Psychiatry 77, 425–433 (2015). rhythms in cells from bipolar disorder patients. Transl
perivascular, paravascular and paravenous channels? 310. Ozburn, A. R. et al. NPAS2 regulation of anxiety-​like Psychiatry 3, e318 (2013).
Fluids Barriers CNS 15, 17 (2018). behavior and GABAA receptors. Front. Mol. Neurosci. 331. Sprouse, J., Reynolds, L., Braselton, J. & Schmidt, A.
290. McClung, C. A. Mind your rhythms: an important role 10, 360 (2017). Serotonin-​induced phase advances of SCN neuronal
for circadian genes in neuroprotection. J. Clin. Invest. 311. Ozburn, A. R., Larson, E. B., Self, D. W. & McClung, C. A. firing in vitro: a possible role for 5-HT5A receptors?
123, 4994–4996 (2013). Cocaine self-​administration behaviors in Clock∆19 Synapse 54, 111–118 (2004).
291. Musiek, E. S. et al. Circadian clock proteins regulate mice. Psychopharmacology 223, 169–177 (2012). 332. Zhang, R., Lahens, N. F., Ballance, H. I., Hughes, M. E.
neuronal redox homeostasis and neurodegeneration. 312. Parekh, P. K. et al. Altered GluA1 (Gria1) function and & Hogenesch, J. B. A circadian gene expression atlas in
J. Clin. Invest. 123, 5389–5400 (2013). accumbal synaptic plasticity in the Clock∆19 model mammals: implications for biology and medicine. Proc.
This preclinical investigation demonstrates that of bipolar mania. Biol. Psychiatry. https://doi.org/ Natl Acad. Sci. USA 111, 16219–16224 (2014).
components of the molecular clock control redox 10.1016/j.biopsych.2017.06.022 (2017). 333. Dokucu, M. E., Yu, L. & Taghert, P. H. Lithium- and
state and, when disrupted, lead to oxidative stress 313. Roybal, K. et al. Mania-​like behavior induced by valproate-​induced alterations in circadian locomotor
and cell death related to neurodegeneration. disruption of CLOCK. Proc. Natl Acad. Sci. USA 104, behavior in Drosophila. Neuropsychopharmacology
292. Kondratov, R. V., Vykhovanets, O., Kondratova, A. A. 6406–6411 (2007). 30, 2216–2224 (2005).
& Antoch, M. P. Antioxidant N-​acetyl-L-​cysteine This study demonstrates mice harbouring a 334. Sathyanarayana, A. et al. Sleep quality prediction from
ameliorates symptoms of premature aging associated mutation in Clock display a behavioural repertoire wearable data using deep learning. JMIR Mhealth
with the deficiency of the circadian protein BMAL1. similar to human mania and provides evidence for Uhealth 4, e125 (2016).
Aging 1, 979–987 (2009). CLOCK specifically in the ventral tegmental area to
293. Chauhan, R., Chen, K. F., Kent, B. A. & Crowther, D. C. be important for these behaviours via modulation Acknowledgements
Central and peripheral circadian clocks and their of dopamine cell firing. Figures were initially designed using BioRender software
role in Alzheimer’s disease. Dis. Model. Mech. 10, 314. Sidor, M. M. et al. Daytime spikes in dopaminergic (biorender.io). Funding for the studies discussed from the
1187–1199 (2017). activity drive rapid mood-​cycling in mice. Mol. authors’ group was obtained from NARSAD, International
294. Homolak, J., Mudrovcic, M., Vukic, B. & Toljan, K. Psychiatry 20, 1406–1419 (2015). Mental Health Research Organization (IMHRO), MH106460,
Circadian rhythm and Alzheimer’s disease. Med. Sci. 315. Spencer, S. et al. A mutation in CLOCK leads to altered DA039865, DA037636, MH082876 and DA023988
6, 52 (2018). dopamine receptor function. J. Neurochem. 123, to C.A.M. and from NARSAD, DA038654 and DA041872 to
295. Huang, Z. et al. Circadian rhythm dysfunction 124–134 (2012). R.W.L.
accelerates disease progression in a mouse model 316. Neufeld-​Cohen, A. et al. Circadian control of
with amyotrophic lateral sclerosis. Front. Neurol. 9, oscillations in mitochondrial rate-​limiting enzymes Author contributions
218 (2018). and nutrient utilization by PERIOD proteins. Proc. Both authors researched data for article, made substantial
296. Skjerve, A., Bjorvatn, B. & Holsten, F. Light therapy for Natl Acad. Sci. USA 113, E1673–E1682 (2016). contributions to the discussion of the content, wrote the article
behavioural and psychological symptoms of dementia. 317. O’Neill, J. S. & Feeney, K. A. Circadian redox and and reviewed or edited the manuscript before submission.
Int. J. Geriatr. Psychiatry 19, 516–522 (2004). metabolic oscillations in mammalian systems.
297. Yamadera, H. et al. Effects of bright light on cognitive Antioxid. Redox Signal. 20, 2966–2981 (2014). Competing interests
and sleep-​wake (circadian) rhythm disturbances in 318. Marcheva, B. et al. Circadian clocks and metabolism. The authors declare no competing interests.
Alzheimer-​type dementia. Psychiatry Clin. Neurosci. Handb. Exp. Pharmacol. 217, 127–155 (2013).
54, 352–353 (2000). 319. Braun, R. et al. Universal method for robust detection Publisher’s note
298. Hanford, N. & Figueiro, M. Light therapy and of circadian state from gene expression. Proc. Natl Springer Nature remains neutral with regard to jurisdictional
Alzheimer’s disease and related dementia: past, Acad. Sci. USA 115, E9247–E9256 (2018). claims in published maps and institutional affiliations.
present, and future. J. Alzheimers Dis. 33, 913–922 This study demonstrates the use of a set of
(2013). computational algorithms called TimeSignature Reviewer information
299. De Lepeleire, J., Bouwen, A., De Coninck, L. to predict the endogenous phase of an individual Nature Reviews Neuroscience thanks A. Phillips, K. Wulff and
& Buntinx, F. Insufficient lighting in nursing homes. based on the particular gene signatures from the the other anonymous reviewer(s) for their contribution to the
J. Am. Med. Dir. Assoc. 8, 314–317 (2007). blood acquired at a single timepoint. peer review of this work.

NAtuRe Reviews | NEuRosCiEnCE volume 20 | JANUARY 2019 | 65

You might also like