You are on page 1of 16

REVIEWS

Transcriptional architecture
of the mammalian circadian clock
Joseph S. Takahashi
Abstract | Circadian clocks are endogenous oscillators that control 24‑hour physiological
and behavioural processes in organisms. These cell-autonomous clocks are composed of a
transcription–translation-based autoregulatory feedback loop. With the development of
next-generation sequencing approaches, biochemical and genomic insights into circadian
function have recently come into focus. Genome-wide analyses of the clock transcriptional
feedback loop have revealed a global circadian regulation of processes such as transcription
factor occupancy, RNA polymerase II recruitment and initiation, nascent transcription, and
chromatin remodelling. The genomic targets of circadian clocks are pervasive and are intimately
linked to the regulation of metabolism, cell growth and physiology.

Convergent evolution
Energetic cycles governed by the solar day have driven how is this hierarchy coupled and integrated? And
The process by which the evolution of life on Earth. To adapt and respond opti‑ what are the functions of clocks in tissue-specific and
unrelated organisms mally to the daily environmental cycles, living systems cell-­specific contexts?
­

independently evolve similar utilize internal timing systems that resonate with the The molecular mechanism of the circadian clock in
traits as a result of adapting
24‑hour day 1 (FIG. 1a,b). Across life forms, the internal mammals is generated by a cell-autonomous transcrip‑
to similar environments or
selective pressures. timing system can be seen at the transcriptional level, tional autoregulatory feedback loop. The core clock genes
giving rise to divergent gene networks in different organ‑ include CLOCK and BMAL1, which encode activators,
Network motif isms that oscillate on a 24‑hour basis2. Circadian clocks, and PER1, PER2, CRY1 and CRY2, which encode repres‑
As defined by Uri Alon, a small as these endogenous timers are known, seem to be the sors. Current topics of active investigation include the
set of recurring regulatory
patterns in a network.
product of convergent evolution, as the design principles components of this core circadian clock gene network,
of circadian transcriptional circuits share a common how environmental and systemic inputs impinge on the
Oscillations ­ etwork motif3 (that is, negative feedback with delay
n core clock mechanism20,21, and how the core clock network
Repetitive variations of (FIG. 1c)), although the specific components are diver‑ regulates downstream output pathways. Recent work has
variables over time with a
gent across the kingdoms of life2,4–7. The link to the daily provided insight into the genomic targets of the core clock
stable frequency or period.
solar energetic cycle is most obvious in photosynthetic pathway and has led to some surprises concerning the
organisms8 but is also a conserved motif in the nutrient pervasiveness of circadian regulation in gene expression.
cycles of higher organisms9–11. Whether circadian clocks A substantial fraction (~5–20%) of genes expressed in
evolved to adapt to the energetic demands of the envi‑ any particular cell or tissue have been found to undergo
ronment or to avoid the detrimental consequences of circadian oscillations at the mRNA level22; however, this
solar radiation remains a topic of debate1,6,11. level of regulation (that is, steady-state mRNA rhythms)
Irrespective of why circadian clocks evolved, it is evi‑ is only one of a multitude of layers of circadian regula‑
dent today that cell-autonomous clocks are ubiquitous. tion in gene expression. Virtually every regulatory stage
In mammals, for example, the neuro-centric view of a in gene expression that has been examined in detail —
master clock located only in the brain is no longer via‑ including transcription, splicing, termination, polyadeny‑
Howard Hughes Medical
Institute, Department of ble. Instead, the discovery of clock genes12 in the 1990s lation, nuclear export, microRNA regulation, translation
Neuroscience, University of showed that virtually all cells express these genes and and RNA degradation — has revealed additional ­layers
Texas Southwestern Medical have the capacity to generate circadian oscillations13,14. of circadian control23,24. Thus, both transcriptional
Center, 5323 Harry Hines The circadian clock in mammals is now conceptual‑ and post-transcriptional regulation of circadian gene
Boulevard, NA4.118, Dallas,
Texas 75390–9111, USA.
ized as a hierarchical system in which a clock located ­expression are key nodal points of control.
joseph.takahashi@ in the hypothalamic suprachiasmatic nucleus acts as a Our understanding of the mechanism of the circadian
utsouthwestern.edu master pacemaker to synchronize or entrain peripheral clock in mammals has been informed by progress on
doi:10.1038/nrg.2016.150 clocks distributed throughout the body 15–19 (FIG. 1d). This two fronts: first, by the biochemical definition of com‑
Published online 19 Dec 2016 system-­wide architecture raises the following questions: ponents of circadian activator and repressor complexes,

NATURE REVIEWS | GENETICS ADVANCE ONLINE PUBLICATION | 1


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

a b c
Behavioural rythms Negative
Light Dark elements
(Energy harvesting, (Energy storage,
DNA damage) DNA repair)

Sleep Wake Delay


Fast Feed
Positive
elements
RE Clock genes

d SCN in vitro SCN in vivo


50
0.4

Bioluminescence
40 0.3

30 0.2

0.1
20
Bioluminescence (103 counts per min)

0 1 2 3 4 5 6 7 0
Day 0 1 2 3 4
Day
Lung
90

75 Fibroblasts
Per2 promoter

Bioluminescence
60 24
0 1 2 3 4
Day

Liver
100 6
0 1 2 3 4
80 Day

60

40

20
0 1 2 3 4
Day

Figure 1 | Circadian rhythms are adaptations of organismal physiology to resonate with the daily solar energetic
Nature Reviews | Genetics
cycle on earth. a | The Earth’s 24‑hour rotation leads to a diurnal cycle of light and darkness that drive an energy
harvesting and energy storage daily rhythm. Solar irradiation also imposes a cycle of DNA damage and recovery.
b | Animals have behavioural rhythms of sleep–wake, and feeding and fasting occur on a 24‑hour basis in synchrony with
the solar day. c | A conserved network motif of circadian clocks involves a transcription–translation negative-feedback
loop with delay. d | In mammals, circadian clocks are cell autonomous and are found in all major organ systems and tissues
of the body. A hierarchical organization exists in which the hypothalamic suprachiasmatic nucleus (SCN) acts as a master
pacemaker to synchronize behavioural and physiological rhythms throughout the body. RE, regulatory element. Part a and
part b are reproduced from REF. 11, Nature Publishing Group. Part d is adapted from REF. 15, Nature Publishing Group.

and, second, by the structural determination of key The core circadian clock gene network
domains of clock proteins (BOX 1). This Review focuses Core components of the mammalian circadian clock.
on the molecular, biochemical and genomic advances in In mammals, the mechanism of the circadian clock con‑
E‑boxes our understanding of circadian regulation in mammals. sists of the activators CLOCK (and its paralogue NPAS2)
Also known as enhancer boxes, I highlight work on the genomic targets of the core cir‑ and its heterodimeric partner BMAL1 (also known as
these are short DNA regulatory
cadian transcriptional pathway, as well as studies reveal‑ ARNTL), which are basic helix–loop–helix (bHLH)–
elements in some eukaryotic
promoters that act as binding ing global circadian oscillations in RNA polymerase II PER-ARNT-SIM (PAS) transcription factors38,39. The
sites for transcription factors (RNAPII)-mediated transcription and their relation to CLOCK–BMAL1 complex binds to regulatory elements
and have thus been found to the dynamic circadian regulation of chromatin state and containing E‑boxes in a set of rhythmic genes that encode
regulate gene expression. chromatin modifiers. For non-mammalian model sys‑ the repressor proteins period (encoded by PER1, PER2
For CLOCK–BMAL1, the
consensus sequence for this
tems and aspects of circadian biology focused on photic and PER3) and cryptochrome (encoded by CRY1 and
cis-regulatory element is regulation, metabolism and immunity, there are other CRY2)39–41 (FIG. 2). In mice, CLOCK–BMAL1 activation
CACGTG. excellent reviews8,25–37. occurs in the daytime, leading to the transcription of the

2 | ADVANCE ONLINE PUBLICATION www.nature.com/nrg


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Per and Cry genes in the afternoon and the accumula‑ casein kinase 1δ (CK1δ) and CK1ε42,43, and trans­locate
tion of the PER and CRY proteins in the late afternoon into the nucleus at night, where they interact with
or evening 42. The PER and CRY proteins interact with CLOCK–BMAL1 to repress their own transcription42,44.
each other as well as with the serine/threonine kinases As repression progresses, Per and Cry transcription

Box 1 | Structural biology of clock proteins


Structural biology will have an essential role in achieving a mechanistic understanding of clock function154,155. The 3D
structures of several circadian proteins and their complexes have been solved using X‑ray crystallography. These structures
include the basic helix–loop–helix (bHLH), PAS‑A and PAS‑B domains of the heterodimeric transcriptional activator
complex CLOCK–BMAL1 (Protein Data Bank identifier: 4F3L)156 (see the figure, part a); the PAS‑A and PAS‑B domains
of the mammalian period (PER) proteins157,158; the photolyase homology region (PHR) of mammalian cryptochrome 1
(CRY1)159; the PHR of CRY2 in complex with the F-box protein FBXL3 (REF. 160); and the PER2‑CBD (CRY-binding domain)
in complex with CRY1 (REF. 161) or CRY2 (PDB ID: 4U8H)162 (see the figure, part b).
The structure of the CLOCK–BMAL1 complex is asymmetric, with CLOCK wrapping around BMAL1. The PAS‑A and
PAS‑B domains of the two subunits interact with their complementary domains but with two different interfaces: the PAS‑A
domains have symmetrical interactions involving the β‑sheet surfaces, whereas the PAS‑B domains have a head‑to‑tail
interaction in which the β‑sheet surface of BMAL1 interacts with the α‑helical surface of CLOCK. The PAS‑A and PAS‑B
domain configuration of BMAL1 (REF. 156) (but not CLOCK) is identical to that seen in the PAS‑A and PAS‑B domains of the
mammalian PER proteins157,158, indicating that PER could interact with CLOCK and replace BMAL1 (REF. 163). Interestingly,
a crystal structure of another member of the bHLH–PER-ARNT-SIM (PAS) heterodimeric protein family, HIF2α (also known as
EPAS‑1)‑ARNT, reveals a completely different overall domain arrangement compared to CLOCK–BMAL1. However, specific
domain interface interactions are preserved between the respective bHLH, PAS‑A and PAS‑B domains of the two subunits164
(see REF. 165 for a comparison of these bHLH–PAS structures). The recognition of the CACGTG E‑box DNA-binding motif of
the CLOCK–BMAL1 bHLH region is similar to that seen for other bHLH proteins166.
The crystal structures of mammalian CRY2 have been determined in five functional states: 1) the PHR by itself (apo); 2) the
FAD-bound CRY2 PHR; 3) the CRY2–FBXL3–SKP1 complex; 4) the KL001‑bound CRY2 PHR; and 5) CRY2 in complex with
PER2‑CBD160,162,167. Both the CRY1 and CRY2 apo structures adopt
the canonical photolyase fold and can be aligned most closely to the a
Drosophila (6–4) photolyase159,160. However, unlike (6–4) photolyases and PAS-B
cryptochromes from plants and Drosophila melanogaster, the mammalian
CRY1 and CRY2 PHRs do not have the FAD cofactor bound. In CRY2, FAD
PAS-B C
can bind with low affinity, and the FAD-binding pocket has a more open
conformation, perhaps explaining the lower affinity for FAD160. CRY2 in
complex with FBXL3 reveals an extensive interface with the FBXL3
leucine-rich repeat (LRR) domain. Moreover, most interestingly, the
carboxy‑terminal tail of FBXL3 inserts into the FAD-binding pocket, PAS-A
revealing a novel substrate recognition motif by the SCFFBXL3 ubiquitin PAS-A
ligase160. The small-molecule stabilizer of CRY, KL001 (REF. 168), binds to
the FAD pocket, but also extends beyond the pocket where the FBXL3 bHLH
C‑terminal tail enters; this binding therefore blocks the FBXL3–CRY2
interaction to stabilize CRY167. Finally, the known interaction between bHLH
CRY and PER proteins, which mutually stabilize one another, has been
illuminated by the recent crystal structures of the CRY1–PER2‑CBD161 N
and CRY2–PER2‑CBD162 complexes. The PER2‑CBD forms an extended CLOCK – BMAL1
conformation on CRY2 that begins at a secondary pocket critical for
CLOCK–BMAL1 binding and winds around the CRY2 C‑terminal helix to N
sterically hinder the recognition of CRY2 by the SCFFBXL3 ubiquitin ligase
(see the figure, part b). Thus, the overlapping interfaces on CRY2 for
PER2‑CBD and FBXL3 provide a structural basis for the competition of PER b
proteins and FBXL3 for controlling the stability of CRY162.
α-helical αβ photolyase
The existing crystal structures raise the question of how PER and CRY domain domain
N
interact with CLOCK–BMAL1 as a quaternary complex. In different
contexts, CRY1 can interact directly with CLOCK–BMAL1 without PER
in vitro169, although in other experiments PER has been shown to be
essential for interaction with CLOCK–BMAL1 in vivo163. It seems probable Zn2+
that PER is involved in reducing the occupancy of CLOCK–BMAL1 on N
DNA170 as seen in D. melanogaster171, whereas CRY1 may be able to
repress CLOCK–BMAL1 on DNA170,172. Because the native (CLOCK– C
BMAL1)–(PER–CRY) quaternary complexes are megadalton in size and
involve other interacting proteins, and important domains of these
proteins are intrinsically disordered, the solving of these complexes will C
probably require a combination of crystallography, NMR and mPER2-CBD
cryoelectron microscopy methods. mCRY2-PHR

Box figure part a reproduced with permission from REF. 156, AAAS. Box figure part b reproduced from REF. 162.
Nature Reviews | Genetics

NATURE REVIEWS | GENETICS ADVANCE ONLINE PUBLICATION | 3


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Cytoplasm
Dbp Per2 Cry1 Bmal1
Nucleus

RORs REV-ERBs

pre-mRNA
RORE Nfil3
NFIL3

RORE Bmal1 RORs


REV-ERBs

Inputs D-box Rora/Rorb 0 4 8 12 16 20 24 28 32 36 40 44


Circadian time (hours)
BMAL1 CLOCK
DBP

D-box E-box Nr1d1/Nr1d2

E-box Dbp
β-TrCP
CK1ε/δ SCF
Repression CRE SRE D-box E-box Per1
+Ub 26S
PERs
CRE SRE D-box E-box Per2
Proteasome

RORE E-box Cry1


+Ub 26S
FBXL21
CRYs
E-box Cry2 SCF

CRY PER Nuclear translocation CRY PER

CK1ε/δ CK1ε/δ

BMAL1 CLOCK

E-box Ccg

RORE Ccg Clock outputs/rhythmic biological processes

CRY
FBXL21 D-box Ccg

AMPK Proteasome
+Ub 26S
FBXL3
SCF

Figure 2 | The circadian gene network in mammals. At the core of the regulate a rhythm in the ROR nuclear receptors. TheseReviews
Nature three interlocked
| Genetics
network, the basic helix–loop–helix (bHLH)–PER-ARNT-SIM (PAS) transcriptional feedback loops represent the three major transcriptional
transcription factors CLOCK and BMAL1 activate the Per1, Per2, Cry1 and regulators of the majority of cycling genes. Different combinations of these
Cry2 genes, whose protein products interact and repress their own factors generate different phases of transcriptional rhythms, as exemplified
transcription. The stability of the period (PER) and cryptochrome (CRY) by the graph (top right) showing the RNA profiles of Dbp, Per2, Cry1 and
proteins are regulated by parallel E3 ubiquitin ligase pathways. CLOCK and Bmal1 in the mouse liver. Additional rhythmic output genes (so-called
BMAL1 also regulate the nuclear receptors REV-ERBα and REV-ERBβ clock-controlled genes or Ccgs) are transcriptionally regulated by the three
(encoded by Nr1d1 and Nr1d2, respectively), which rhythmically repress the loops acting on E‑boxes, RevDR2 and ROR-binding elements (ROREs) and
transcription of Bmal1 and Nfil3 (which encodes nuclear factor, interleukin‑3 D‑boxes in the regulatory regions of target genes. AMPK, 5ʹ AMP-activated
regulated) that is driven by the activators retinoic acid-related orphan protein kinase; CK1, casein kinase 1; CRE, cAMP response element; FBX,
receptor-α (RORα) and RORβ (encoded by Rora and Rorb, respectively). F-box protein; SCF, SKP1–cullin–F‑box protein; SRE, serum response
NFIL3 in turn represses the PAR-bZip factor DBP (D‑box binding protein) to element; Ub, ubiquitin.

4 | ADVANCE ONLINE PUBLICATION www.nature.com/nrg


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

RevDR2 and retinoic declines and the PER and CRY protein levels decrease; REV-ERB–ROR loop52,53 (FIG. 2). Together, these three
acid-related orphan the half-lives of PER and CRY proteins are relatively interlocking transcriptional feedback loops can generate
receptor (ROR)-binding short and specific E3 ligase complexes target these pro‑ cycles of transcription with various phases of expression
elements teins for ubiquitylation and subsequent degradation by depending on the combination of cis-­elements (that is,
(ROREs). The cis-regulatory
elements for the nuclear
the proteasome43,45. Once negative-feedback repression is E‑boxes, ROREs and D‑boxes) in the promoters and
receptors REV-ERB and ROR. relieved by turnover of the repressor complex, transcrip‑ enhancers of specific target genes54 (FIG. 2).
The consensus sequence of tion by CLOCK–BMAL1 can begin anew to start a new
RevDR2/RORE motifs involve cycle of transcription the next morning. Importantly, Core clock protein interactions with chromatin and
RGGTCA half-sites preceded by
core clock genes, such as those that encode PER2 and chromatin-modifying complexes. At the beginning
an (A/T)-rich region.
its regulatory kinase CK1δ, have been found to underlie of the transcription cycle, the activators CLOCK and
D‑boxes human sleep-timing disorders46,47, thus demonstrating a BMAL1 interact with the histone acetyltransferases
Short cis-regulatory elements clear role for clock genes in humans (BOX 2). (HATs) p300 and CREB-binding protein (CBP), respec‑
for the PAR-zip transcription In addition to the PER and CRY target genes, the tively, to acetylate histones and provide an accessible
factors DBP (D-box binding
protein), HLF (hepatic
CLOCK–BMAL1 complex also activates the nuclear chromatin state for transcription55–58. CLOCK has
leukaemia factor) and TEF receptors REV-ERBα and REV-ERBβ (encoded by been reported to have intrinsic HAT activity and to
(thyrotroph embryonic factor). NR1D1 and NR1D2, respectively), which compete at acetylate histone H3 on Lys9 (H3K9) and Lys14 resi­
The common sequence of a RevDR2 and retinoic acid-related orphan receptor (ROR)- dues (H3K14) 59,60. The NAD +-dependent histone
D‑box is TTATG(C/T)AA.
binding elements (ROREs) with RORα, RORβ and RORγ deacetylase (HDAC) sirtuin 1 (SIRT1) associates with
Phases of expression (encoded by RORA, RORB and RORC, respectively)48–50 CLOCK, BMAL1 and PER2 (REFS 61,62), and a circadian
The time of the day when peak (FIG. 2). The rhythmic expression of REV-ERBα and REV- rhythm in NAD+ levels driven by the expression of the
gene expression occurs. ERβ leads to the repression of BMAL1 and CLOCK, CLOCK–BMAL1 target gene Nampt in turn leads to a
which induces a rhythm in these genes that is in anti- rhythm in SIRT1 activity that feeds back to inhibit the
phase with PER expression rhythms48. These inter‑ ­CLOCK–BMAL1 complex 63,64.
locked feedback loops of activators and repressors form CLOCK and BMAL1 have also been shown to be
a canonical positive-feedback and negative-feedback associated with the methyltransferase MLL1 (mixed lin‑
gene network motif 51. A third CLOCK–BMAL1‑driven eage leukaemia 1)65. Interestingly, MLL1 interacts with
transcriptional loop involves the PAR-bZip (proline the Δ19 activation domain of CLOCK and is thought to
and acidic amino acid-rich basic leucine zipper) fac‑ be important for the recruitment of CLOCK to DNA65.
tors DBP (D‑box binding protein), TEF (thyrotroph Moreover, MLL1 is associated with rhythmic H3K4 tri‑
embryonic factor) and HLF (hepatic leukaemia factor). methylation (H3K4me3)65, a common histone marker
These proteins interact at sites containing D‑boxes with associated with active gene promoters. In addition,
the repressor NFIL3 (nuclear factor, interleukin‑3 regu­ MLL1 is acetylated and the metabolic regulator SIRT1
lated; also known as E4BP4), which is driven by the rhythmically deacetylates MLL1, thereby affecting its
methyltransferase activity 66. The related methyltrans‑
ferase MLL3 exhibits circadian occupancy and is associ‑
Box 2 | Human genetics of clock genes
ated with promoters bearing rhythmic H3K4me3 marks;
Validation of the core circadian transcriptional mechanism elucidated in rodents (FIG. 2) however, MLL3 recruitment to DNA does not depend on
has come from the Mendelian disorder familial advanced sleep phase disorder (FASPD), BMAL1 or CLOCK67.
in which the timing of sleep–wake cycles in affected individuals are shifted several CLOCK and BMAL1 also interact with the
hours earlier than normal173. Dominant mutations in the genes that encode period 2 lysine-specific histone demethylases JARID1A and
(PER2) and the serine/threonine kinases casein kinase 1δ (CSNK1D) have been shown
LSD1, which are both thought to promote the recruit‑
to be linked to FASPD and to be causative in humans46,47. In large-scale human
genome-wide association studies (GWAS) for fasting glucose levels and increased risk
ment of CLOCK–BMAL1 to PER promoters to enhance
of type 2 diabetes mellitus, common variants in the MTNR1B locus, which encodes transcription68,69. BMAL1 interacts with TRAP150 (thy‑
melatonin receptor 1B, and in CRY2, which encodes cryptochrome 2, have been roid hormone receptor-associated protein 150), a co‑­
documented174–177. Melatonin signalling is a major output pathway of the activator that promotes CLOCK–BMAL1 binding to
suprachiasmatic nucleus (SCN) in mammals, and levels of melatonin are circadian, with circadian target genes and links CLOCK–BMAL1 to the
high levels during night-time and low levels during daytime178, reflecting a physiological Mediator complex subunit MED1 to recruit the general
signal of darkness. The MTNR1B mutations have been proposed to act in pancreatic transcriptional machinery 70. RNAPII is both recruited
islets to impair insulin secretion, but this hypothesis remains a subject of debate179–181. and initiated on a circadian basis as discussed in more
Recently, large-scale GWAS for morningness and sleep-timing preference (chronotype) detail below 71,72.
have identified several genes known to be associated with circadian rhythms in
During the transition from activation to repression,
animals182–184. These genes include RGS16, a regulator of G protein signalling that is
enriched in the SCN and has a long period phenotype with a loss‑of‑function
the CLOCK–BMAL1 complex interacts with the PER–
mutation185; the core clock gene PER2; VIP (which encodes vasoactive intestinal CRY repressor complex at the beginning of the night
polypeptide), a neuropeptide in the SCN involved in oscillator coupling186; HCRTR2, as the levels of PER and CRY accumulate and trans‑
the receptor for hypocretin or orexin that is linked to narcolepsy in humans and dogs187; locate to the nucleus42. The PER complex is large and
and FBXL3, an F‑box protein that is the recognition domain for an E3 ubiquitin ligase contains at least 25 proteins 73–75. Initially, CLOCK
complex that degrades the CRY proteins86. GWAS using sleep duration as a phenotype and BMAL1 are associated with two components of
have identified a few loci (PAX8, VRK2 and UFL1)183,188,189, but their relation to sleep the Mi‑2–nucleosome remodelling and deacetylase
mechanisms remains obscure. Overall, it is reassuring that numerous GWAS hits are (NuRD) transcriptional co-repressor complex: CHD4
associated with genes with known circadian and sleep functions, thus fortifying the (chromodomain-helicase-DNA-binding protein 4)
notion that circadian transcriptional mechanisms are likely to apply to human biology.
and MTA2 (metastasis-associated 1 family member 2).

NATURE REVIEWS | GENETICS ADVANCE ONLINE PUBLICATION | 5


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Interestingly, PER complexes without CLOCK–BMAL1 kinase (AMPK)-mediated phosphorylation, which


lack CHD4 and MTA2 but contain other members of the provides a link to metabolism via nutrient-responsive
NuRD complex, such as MBD2 (methyl-CpG-binding AMPK activity 88.
domain protein 2), GATAD2A (GATA zinc finger The stability and turnover of PER proteins is primarily
domain-­containing protein 2A), HDAC1 and RBAP48 regulated by several mechanisms, including by CK1δ and
(retinoblastoma-­binding protein p48)75. This finding CK1ε43,81,84,89, by the balance between these two protein
indicates that during this transition phase there is a kinases and phosphatase 1 (REF. 90), and by the stoichiom‑
split NuRD complex that then comes together to form etry of PER proteins and these kinases91,92. Recent work
a reconstituted NuRD co-repressor complex to inhibit has shown that a dual kinase, multisite ­phosphoswitch
CLOCK–BMAL1 transcription. The CLOCK–BMAL1 regu­lates PER stability and its ­ubiquitylation by the
complex also recruits polycomb repressive complex 2 β‑TrCP E3 ubiquitin ligase complex 93.
(PRC2), leading to the trimethylation of H3K27 at tar‑ For CRY proteins, their stability and turnover are
get genes and thereby inhibiting gene expression76,77. regulated by the balance of the two paralogues FBXL3
In  addition, CLOCK–BMAL1 recruits the DNA and FBXL21 (REFS 94,95) (FIG. 2). Contrary to expec‑
damage-­binding protein 1 (DDB1)–cullin 4 (CUL4) tation, FBXL21 antagonizes FBXL3 to reduce the
ubiquitin ligase complex to circadian target genes to pro‑ E3 ligase activity of SCFFBXL3 on CRY1 and CRY2 to sta‑
mote H2B mono­ubiquitylation at E‑box sites to cova‑ bilize them95. Although SCFFBXL21 does possess E3 ligase
lently mark genes for subsequent recruitment of the PER activity, it is less efficient than SCFFBXL3 in the ubiqui‑
complex 78. At the peak of repression during the middle tylation of CRY95. However, FBXL21 interacts more
of the night, the PER complex is also associated with strongly with CRY than FBXL3, which protects CRY
the transcriptional repressor complexes SIN3–HDAC79 from SCFFBXL3 activity 95. The subcellular localization of
and HP1γ–SUV39H histone methyltransferase80, which FBXL3 and FBXL21 also differ, with FBXL3 acting as
contribute to CLOCK–BMAL1 repression through the the primary E3 ligase for CRY in the nucleus, whereas
deacetylation of histone H3K9 and H4K5 as well as FBXL21 acts as the primary E3 ligase for CRY in the
dimethylation and trimethylation of H3K9. The PER cytoplasm95. Finally, in genetic experiments exploring
repressor complex also includes NONO, a protein the inter­action of the CK1εtau mutant with the FBXL3Afh
involved in RNA processing, and WDR5, a component mutant, a broad range of periodicities were seen
of a histone methyltransferase complex 73. At later times depending on the combination of alleles96, indicating
during the repression phase, PER complexes include the that the PER and CRY pathways independently affect
RNA heli­cases DDX5 and DHX9, as well as the large the ultimate period length in vivo.
subunit of RNAPII and SETX, a helicase that promotes
transcriptional termination74. Thus, the recruitment Genome-wide clock architecture
of several chromatin-modifying complexes as co‑­ Although previous work has examined the transcrip‑
repressors is neces­sary for PER-mediated and prob‑ tional regulation of CLOCK–BMAL1 target genes such
Period ably CRY-mediated repression of CLOCK–BMAL1. as Dbp, Per1 and Nr1d1 in detail, the identity of the cir‑
The length of the circadian
During the circadian cycle, transcriptional activation cadian cistrome has only been recently defined. Using
rhythm measured from specific
phase points in each cycle; for and repression are accompanied by dynamic epigenetic chromatin immunoprecipitation followed by sequen­
example, the peak‑to‑peak transitions in chromatin state that poise the genome for cing (ChIP–seq), several research groups have identified
interval. rhythmic gene expression. genome-wide DNA-binding sites for BMAL1, CLOCK,
NPAS2, PER1, PER2, CRY1 and CRY2 in mouse liver
Phosphoswitch
In the case of the PER2 protein,
Regulation of circadian period length by repressor during the circadian cycle71,97–100.
a phosphoswitch model is stability. The stability of the core repressor proteins is
proposed whereby two an important determining factor in the dynamics of Genomic targets of the core circadian activators.
competing phosphorylation the transcriptional feedback loop, which in turn regu­ Occupancy of CLOCK–BMAL1 at the Dbp locus,
sites (one for the FASPD site
lates the period length of the oscillation. The stability which encodes a transcription factor that modulates
and one for the β‑TrCP binding
site) determine whether PER2 of PER and CRY is regulated by their phosphorylation the expression of clock output genes, is circadian, with
has a fast or slow degradation state and by their ubiquitylation by E3 ligase complexes peak binding occurring during the middle of the day 101.
rate. (FIG. 2). For example, the tau mutation in CSNK1E (which At the single-cell level, CLOCK and BMAL1 recruitment
encodes CK1ε) shortens the circadian period length to E‑boxes in Dbp are interdependent and circadian102.
Cistrome
The in vivo genome-wide
in vivo81. Moreover, this mutation causes the destabili‑ BMAL1 binding is dynamic and stochastic, and fluctu‑
location of transcription zation of PER proteins by phosphorylating a regulatory ations in binding depend on the proteasome102, which
factor-binding sites. site that marks the PER proteins for ubiquitylation by is consistent with a ‘kamikaze’ model of t­ ranscriptional
β‑TrCP (also known as F-box/WD repeat-containing activation 103. On a genome-wide level, CLOCK and
‘Kamikaze’ model of
protein 1A) and proteas­omal degradation43,82–84. By con‑ BMAL1 occupancy also peak within a narrow window
transcriptional activation
Describes the ubiquitin-­ trast, mutations in the gene encoding the F‑box protein of time during the middle of the day, and the phasing of
dependent proteolysis of FBXL3, which is one of four subunits of the SCF (SKP1– CLOCK–BMAL1 occupancy is not tightly phase-locked
transcriptional activators, cullin–F‑box protein) E3 ligase complex, lengthen with the RNA expression of their target genes71,97,98. This
which suggests a role for the circadian period and stabilize the CRY proteins flexible phase relationship can be partially explained by
activator degradation in RNA
polymerase II elongation and
by decreasing ubiquitylation and subsequent protea­ the combinatorial action of other cis-acting elements
the requirement for reloading somal degradation85–87. In addition, CRY ubiquity­lation (ROREs and D‑boxes) within the tripartite clock gene
of newly synthesized activators. by SCFFBXL3 is triggered by 5ʹ AMP-activated protein network54 (FIG. 2). For example, Cry1 expression, which

6 | ADVANCE ONLINE PUBLICATION www.nature.com/nrg


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

a b BMAL1 PER1 CRY1 CRY2


Scale 5 kb mm9
chr11: 68910000 68915000 0 4 8 12 16 20 K 0 4 8 12 16 20 K 0 4 8 12 16 20 K 0 4 8 12 16 20 K
0
4
8
BMAL1 12
16
20
KO
0
4
8
CLOCK 12
16
20
KO
0
4
8
PER1 12
16
0
n = 4,653 20
20
KO
0 0 20
4 n = 5,952
8 PER2
PER2 12
16 CLOCK 0 4 8 12 16 20 K
20
KO
0 0 4 8 12 16 20 K
4
8
CRY1 12
16
20
KO 0 15
0 n = 10,110
4
CRY2 8
12
16
20
KO

UCSC genes Per1


Mammal cons 0 15
n = 4,636

0 20
n = 7,392
0 20
n = 16,506
c

660 PER1 (4,653) CRY1 (16,506)


172
BMAL1 (5,952) 48
194 196
46 1011
48 39 54 9 5 79
182 10 50 5,662
748 194 55
490 789
173 186
209
10 284
8 1,837
10 1,444
216
126 839 2,130
456 65 6 241 CRY2 (10,110)
148 4 11 5 150 85 230
34 10 18
134 2 9 74 20
45 775
8 22
197 PER2 (7,392)
727
264
CLOCK (4,636)
300

Figure 3 | The circadian cistrome of the mouse liver. a | University of from left to right. Knockout (K) mouse control isNature
shownReviews | Genetics
on the far right of
California Santa Cruz (UCSC) genome browser view of chromatin each panel. The number of peaks in the genome is indicated at the bottom
immuno­precipitation followed by sequencing (ChIP–seq) profiles of of each panel. The blue–red gradient indicates the coverage for all binding
circadian transcription factors at the Per1 gene at six circadian times (CTs) sites in the genome. The number of binding sites (n) in the genome are
of the day (0, 4, 8, 12, 16 and 20 hours). The colours of the wiggle plots of indicated at the bottom of each heat map. c | Chow–Ruskey diagram
ChIP–seq occupancy indicate the following transcriptional regulators: showing the six‑way overlap of BMAL1, CLOCK, PER1, PER2, CRY1 and
BMAL1 (blue), CLOCK (green), PER1 (orange), PER2 (gold), CRY1 (red) and CRY2 peaks in the genome. The red circle in the middle represents the
CRY2 (pink). KO indicates a ChIP–seq sample from a knockout mouse for overlap of all six factors. Lighter shades of red, orange and yellow
each transcriptional regulator. b | Heat map views of genome-wide DNA represent fewer overlaps of subsets. The boundaries for each protein
binding for BMAL1, CLOCK, PER1, PER2, CRY1 and CRY2 measured over are colour coded: BMAL1 (blue), CLOCK (green), PER1 (orange), PER2
500 bp fragments encompassing the binding sites. Each peak in the (brown), CRY1 (red) and CRY2 (pink), and the areas of each domain are
genome is represented as a horizontal line, ordered vertically by signal proportional to the number of binding sites. Adapted with permission
strength. Six time points are shown beginning at CT0 and ending at CT20 from REF. 71, AAAS.

NATURE REVIEWS | GENETICS ADVANCE ONLINE PUBLICATION | 7


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

a Intron Exon b d
0 8 16 24 32 40 0 8 16 24 32 40 Exon (2,037)
Intron (1,371) RNAPII–
8WG16
0 4 8 12 16 20

913 458 1,579

c Intron Exon
0 0

18 6 18 6

12 12

CT15.1 CT7.8
400 250
Frequency

0 0
0 8 16 24 0 8 16 24
n = 1,371
Circadian time (hours)

0 10
n = 7,353

CT14.5
1,250 n = 4,043

0
0 8 16 24
n = 2,037 Circadian time (hours)

Figure 4 | Whole-transcriptome RNA sequencing analysis of circadian gene expression in the mouse liver. a | Heat
Nature Reviews | Genetics
map view of intron (left) and exon (right) RNA cycling genes. Each gene is represented as a horizontal line, ordered
vertically by phase in hours indicated at the top. Twelve time points are shown beginning at circadian time zero (CT0) and
ending at CT44 from left to right. n indicates the number of cycling genes in each category. b | Venn diagram of intron and
exon cycling genes showing only 22% overlap. c | The phase distribution of cycling genes. The phase of each transcript
rhythm is represented in a dot plot (top) and histogram plot (bottom). d | Heat map view of cycling RNA polymerase II
(RNAPII)–8WG16 occupancy across the genome of the liver with a peak occurring at night-time. Part a and part b are
reproduced and part c and part d are adapted with permission from REF. 71, AAAS.

peaks much later in phase than Per genes, is regulated by analyses, CLOCK and BMAL1 bind to more than 4,600
both E‑boxes and ROREs104. and 5,900 sites, respectively, corresponding to ~3,000
Circadian time In an effort to view the dynamics of the occupancy unique genes in the liver 71 (FIG. 3b). At highly expressed
(CT). A standard of time based of the core circadian activator complexes in a genome- target genes, CLOCK–BMAL1 DNA-binding motifs
on the free-running period of a wide context, one study assessed all core clock proteins are enriched for a tandem E‑box motif 97. Native liver
rhythm (oscillation). By in the mouse liver 71. At Per1, the activators BMAL1 extracts showed that CLOCK–BMAL1 bind to tandem
convention, CT0 is the
beginning of the subjective day
and CLOCK bind in a cyclic manner at the promoter E‑box sites coopera­tively with higher affinity to more
and CT12 is the beginning of between circadian time zero (CT0) and CT12, with maxi­ effectively compete with other E‑box-binding factors
the subjective night. mal binding observed at CT8 (FIG. 3a). In genome-wide such as USF1 (upstream stimulatory factor 1)105.

8 | ADVANCE ONLINE PUBLICATION www.nature.com/nrg


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

a Scale 10 kb mm9
chr11: 68910000 68915000 68920000 68925000
0
4
BMAL1 8
12
16
20
0
4
H3K4me1 8
12
16
20
0
4
H3K4me3 8
12
16
20
0
4
H3K9ac 8
12
16
20
0
4
H3K27ac 8
12
16
20
0
4
H3K36me3 8
12
16
20
0
4
H3K79me2 8
12
16
20

Per1
UCSC genes

Mammal cons

b
CT0 CT4 CT8 CT12 CT16 CT20
RNAPII–8WG16
Expressed genes Unexpressed genes Intron cycling genes Non-cycling genes
4 4 4 4

3 3 3 3
Coverage

2 2 2 2

1 1 1 1

0 0 0 0
−3 −2 −1 0 1 2 3 −3 −2 −1 0 1 2 3 −3 −2 −1 0 1 2 3 −3 −2 −1 0 1 2 3
Distance from TSS (kb) Distance from TSS (kb) Distance from TSS (kb) Distance from TSS (kb)

H3K4me3
Expressed genes Unexpressed genes Intron cycling genes Non-cycling genes

8 8 8 8

6 6 6 6
Coverage

4 4 4 4

2 2 2 2

0 0 0 0
−3 −2 −1 0 1 2 3 −3 −2 −1 0 1 2 3 −3 −2 −1 0 1 2 3 −3 −2 −1 0 1 2 3
Distance from TSS (kb) Distance from TSS (kb) Distance from TSS (kb) Distance from TSS (kb)

Nature Reviews | Genetics


NATURE REVIEWS | GENETICS ADVANCE ONLINE PUBLICATION | 9
©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Global Run‑On sequencing The hepatic steroid receptor co-activator 2 (SRC2; circadian gene expression112. Thus, the circadian cis‑
(GRO-seq). A nuclear run‑on also known as N-CoA2) is recruited to DNA in a trome is highly dynamic and involves the core circadian
assay method that captures diurnal cycle and overlaps extensively with CLOCK– E‑box transcriptional regulators and is interlinked via
nascent transcripts from BMAL1 sites106. SRC2 acts as a potent transcriptional REV-ERBα/β as well as DBP and NFIL3 to secondary
initiated RNA polymerase II
followed by next-generation
co-­activator of CLOCK–BMAL1, and null Src2 muta‑ and tertiary gene targets with a wide range of phases of
sequencing. tions in mice lead to an abnormally early phase of the gene expression.
activity rhythm106, indicating that SRC2 has a role within
Zeitgeber time the circadian clock system. Genome-wide circadian RNA expression. To examine
(ZT). A standard of time based
In addition to the liver, the circadian cistrome has the functional consequences of the circadian cistrome,
on the period of an
environmental synchronizer or
recently been described in mouse pancreatic islets107. In a RNA sequencing (RNA-seq) has been used to profile
zeitgeber, such as the 24‑hour comparison of the occupancy of BMAL1 between the cycling genes in the mouse liver 71,98,99. For example, using
diurnal cycle of light and liver and pancreas, ~40% of the target genes overlapped, both the intron RNA signal as a proxy for pre-mRNA
darkness. Under standard with disparate binding sites clustering at distal regula‑ and the exon RNA signal as a proxy for mRNA113,114,
light–dark cycles, the time of
‘lights on’ usually defines
tory regions of tissue-specific genes. Interestingly, even strand-specific whole-transcriptome RNA-seq revealed
zeitgeber time zero (ZT0) for within the overlapping set of target genes, BMAL1 was a robust circadian oscillation at the Dbp and Per2 genes71.
diurnal organisms and the time found to bind to distinct enhancers in the liver and pan‑ As was previously observed in microarray experiments,
of ‘lights off’ defines ZT12 for creas107, consistent with the existence of tissue-specific thousands of genes cycle at the steady-state mRNA level
nocturnal animals.
circadian cis-regulatory modules108. in the liver as well as in many other tissues22,115. As more
Nascent-seq regions in the mouse have been profiled, more than
Next-generation sequencing of Genomic targets of the core circadian repressors. In con‑ half of all genes have been found to cycle at the mRNA
nascent RNA transcripts based trast to the activators, the repressors PER1, PER2 and level in at least one tissue22. Importantly, the identified
on chromatin-associated RNA CRY2 bind genome-wide at night between CT12 cycling genes are enriched for drug targets; more than
transcripts, nuclear run‑on
global transcripts, such as
and CT20, peaking between CT15 and CT17 (REF. 71) half of the top selling 100 drugs on the market in the
GRO-seq or PRO-seq, or RNA (FIG. 3b). CRY1 has a bimodal pattern, with a major peak at United States have a circadian target gene22. This finding
sequencing of transcripts CT0 (FIG. 3b). The repressors CRY1 and CRY2 bind to sig‑ emphasizes the importance of optimizing the timing of
immunoprecipitated with RNA nificantly more sites in the genome than PER1 and PER2, drug administration for circadian drug targets, which
polymerase II.
with many thousands of these sites being independent of can both enhance efficacy while at the same time reduce
CLOCK–BMAL1 and containing DNA-binding motifs side effects.
for nuclear receptors71 (FIG. 3c). Consistent with this find‑ A wide range of cycling RNA species has been found
ing, CRY1 interacts with several nuclear receptors and using RNA-seq. In addition to cycling protein-coding
is a co-repressor for the glucocorticoid receptor 109. The transcripts, oscillating long intergenic non-coding RNA,
nuclear receptors REV-ERBα and REV-ERBβ also occupy microRNA and antisense RNA transcripts have been
DNA rhythmically, with a peak in the daytime110,111. REV- found in the mouse liver 71,98,99,116. Using Global Run‑On
ERBα competes with RORα for DNA binding at RORE ­s equencing (GRO-seq) 117 to identify enhancer RNA
and RevDR2 sites to regulate circadian target genes such (eRNA) transcripts118, thousands of oscillating enhancers
as Bmal1 and Cry1, where it recruits nuclear receptor have been found, and the phase of these cycling eRNAs
corepressor 1 (N‑CoR) and HDAC3 to repress gene corresponds with the phases of transcription of nearby
expression50,111. Interestingly, REV-ERBα also interacts genes119. A motif analysis of cycling enhancers at different
at other genomic sites that are enriched for tissue-spe‑ circadian phases revealed that there was phase-­specific
cific transcription factors such as hepatocyte nuclear enrichment of transcription factor motifs and that four
factor 6 (HNF6) in the liver 50. At these sites, the action of sets of motifs were enriched at different times119. E‑box
REV-ERBα is independent of its DNA-binding domain motifs were enriched between zeitgeber time six (ZT6)
and defines a different set of target genes that are directly and ZT9, corresponding to the peak of CLOCK–BMAL1
involved in metabolism. Finally, SRC2 has been proposed occupancy. D‑box motifs were enriched between ZT9
to modulate an oscillation in chromatin a­ ccessibility and ZT15, corresponding to the expression of Dbp and
— by recruiting PBAF (polybromo-­associated Brg/ the low point of Nfil3 repression. RORE and REV‑DR2
Brahma-associated factors) members of the SWI–SNF motifs were enriched between ZT18 and ZT24, corre‑
(switch–sucrose non-­fermentable) ­complex — to pro‑ sponding with the low point of REV-ERB repression of
mote REV-ERB loading and enhance the amplitude of the RORs. ETS (E26 transformation-specific)-binding
motifs were enriched between ZT0 and ZT3 (REF. 119),
which may correspond to GM129 (also known as
◀ Figure 5 | Circadian chromatin states in the mouse liver. a | University of California circadian-­a ssociated transcriptional repressor or
Santa Cruz (UCSC) genome browser view of histone methylation and acetylation at the CHRONO) repression120–122. Overall, the circadian cis‑
Per1 gene at six circadian times (CTs) of the day (0, 4, 8, 12, 16 and 20 hours). The colours of tromic results agree well with their transcriptional read‑
the wiggle plots of chromatin immunoprecipitation followed by sequencing (ChIP–seq) outs as assessed by eRNA and gene expression ana­lysis:
signal indicate the following: BMAL1 occupancy (dark blue), monomethylation of Lys4 at
the tripartite core circadian transcriptional feedback
histone H3 (H3K4me1; red), H3K4me3 (pink), H3K9ac (light blue), H3K27ac (orange),
H3K36me3 (light green), and H3K79me2 (dark green). b | Binding profiles of RNA loop (E‑box, D‑box and RORE) drives the majority of
polymerase II (RNAPII)–8WG16 and H3K4me3 at the transcription start site (TSS) ± 3 kb ­circadian gene e­ xpression on a genome-wide level.
in 12,680 expressed genes (first column), 8,945 unexpressed genes (second column), Using whole-transcriptome RNA-seq or nascent-seq
1,371 intron cycling genes (third column) and 5,839 non-cycling genes (last column). combined with mRNA-seq, two divergent sets of cycling
Cons, conservation. Adapted with permission from REF. 71, AAAS. genes have been found: cycling nascent transcripts

10 | ADVANCE ONLINE PUBLICATION www.nature.com/nrg


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Poised Derepression Activation Transcription Repression


30
CLOCK–BMAL1 CLOCK–BMAL1 Nascent transcription Repression phase
CRY1 repressed Activation phase
RNAPII-Ser5P p300 recruitment
Poised state H3K9 acetylation
Derepression
PER1
15.9
20 PER2
RNAPII– 17.3
8WG16
NPAS2 14.5
% of peaks

BMAL1 8.1
6.1 CRY2
15.4
CBP
H3K9ac 20.1
RNAPII- p300 RNA
CRY1 6.6 CLOCK K36me3
Ser5 5.0 (intron) H3K4me3
10 P0.6 0.4 7.3 17.8 19.8
15.1

H3K27ac K79me2
H3K4me1 17.1
5.8 22.4

CBP

0
0 2 4 6 8 10 12 14 16 18 20 22 24
Circadian time (hours)

Figure 6 | Circadian transcriptional landscape in the mouse liver. Histograms showing the phase distributions of each
Nature
factor as a function of time of day. ac, acetylation; CBP, CREB-binding protein; CRY, cryptochrome; Reviews | Genetics
me, methylation;
NPAS2, paralogue of CLOCK; PER, period; RNAPII, RNA polymerase II; Ser5P, phosphorylation on Ser5. Reproduced with
permission from REF. 71, AAAS.

and cycling mRNA levels71,98 (FIG. 4a). Only ~20% of Genome-wide circadian regulation of transcription,
cycling mRNAs also had a corresponding cycling nas‑ RNAPII and chromatin states. In the mouse liver,
cent transcript 71,98, indicating that other RNA processes cycling pre-mRNA levels (using the intron RNA sig‑
are likely to be contributing to steady-state circadian nal) were coordinately expressed in time, with a peak at
mRNA expression (FIG. 4b). Several different classes of CT15 (FIG. 4c). Thus, there is a peak of globally coordi‑
cycling transcripts have been observed, including the nated circadian transcription that occurs at the begin‑
following: first, cycling transcription and cycling steady- ning of the night during the active or awake portion of
state mRNA; second, cycling transcription and no cycling the mouse circadian rhythm, which is not seen at the
mRNA; and third, no detectable cycling transcription, mRNA level (which has a bimodal pattern). To further
but cycling mRNA levels. The first class is typical for the analyse the global rhythms in nascent transcription,
core circadian genes and their high-amplitude direct tar‑ RNAPII occupancy has been measured during the cir‑
gets. The second class is typically seen in highly expressed cadian cycle71,72. The large subunit of RNAPII (POLR2A)
genes in the liver that have mRNAs with long half-lives, contains a carboxy‑­terminal domain (CTD) that is post-­
such as albumin. The third class represents transcripts translationally modified at different stages of transcrip‑
that could become rhythmic via circadian variation in tion127–129. RNAPII is recruited into the pre-initiation
mRNA turnover, RNA splicing 123, polyadenylation124, complex in a hypophosphorylated state (recognized by
microRNA regulation116 and other post-transcriptional the 8WG16 antibody)130. Surprisingly, RNAPII–8WG16
steps. Whether rhythmic translation or protein levels ulti‑ occupancy is highly circadian in the mouse liver, with a
mately occur has also been revealing. For example, rhyth‑ peak that correlated with the intron RNA peak (FIG. 4d).
mic diurnal polyadenylation occurs in 2.3% of expressed The first step in the initiation of RNAPII involves phos‑
genes in the liver, and one-fifth of these genes (third phorylation on Ser5 (Ser5P) of the CTD of RNAPII130,131.
class) lead to rhythmic protein expression that correlate RNAPII‑Ser5P occupancy is also circadian71, indicating
with polyA tail length in the absence of rhythmic mRNA that both recruitment and initiation of RNAPII are
­levels124. Ribosomal profiling has also revealed substantial under circadian control. An antibody against the RPB2
numbers of examples of circadian translation in U2OS subunit of RNAPII (POLR2B) revealed that RNAPII
cells and liver cells125,126. Interestingly, there are at least occupancy for promoters and gene bodies is rhythmic
150 examples of translationally driven oscillations in the and co‑varies over time72. Rhythmic loading of RNAPII
liver for which no mRNA rhythms are present 125. Thus, to promoters rather than a rhythmic transition from a
both transcriptional and post-transcriptional regulation paused state to productive elongation accounts for the
are critical for circadian gene expression. majority of rhythmic transcription in the mouse liver 72.

NATURE REVIEWS | GENETICS ADVANCE ONLINE PUBLICATION | 11


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Enhancer Promoter

CRY1

C B C B CLOCK–BMAL1

TF
TSS

C B

C B TF

TSS

p300/CBP p300/CBP

TSS
C B TF C B TF

GTFs RNAPII
+ Recruitment

Ser5
Initiation
GTFs
C B C B TF
RNAPII
RNAPII RNAPII
TSS

Pause
Ser5
DSIF
NELF
C B C B TF GTFs
RNAPII
RNAPII RNAPII
TSS Pause region

P-TEFb
Release

NELF Ser5
Ser2

C B C B TF GTFs
RNAPII
RNAPII RNAPII
TSS P-TEFb
eRNA
Nascent RNA Elongation

Figure 7 | Circadian transcriptional regulation at enhancer and promoter sites during the CLOCK–BMAL1 Nature Reviewsactivation
| Genetics
phase during the daytime. CLOCK and BMAL1 act as pioneer transcription factors at enhancer sites. Circadian
recruitment and initiation (as seen by RNA polymerase II (RNAPII)‑Ser5 marks) of RNAPII occurs at both distal enhancers and
proximal promoter sites. However, the circadian regulation of pausing factors (negative elongation factor (NELF) and DRB-
sensitivity-inducing factor (DSIF)) and pause release factors (positive transcription elongation factor b (P‑TEFb)) remain to
be determined. CBP, CREB-binding protein; CRY1, cryptochrome 1; eRNA, enhancer RNA; GTF, general transcription factor;
TF, transcription factor; TSS, transcription start site. Figure adapted from REF. 129, Nature Publishing Group.

12 | ADVANCE ONLINE PUBLICATION www.nature.com/nrg


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Box 3 | Four-dimensional regulation of the genome peaks along with p300 and CBP recruitment and are
accompanied by histone H3K4me1 and H3K9ac acti‑
It is well established that chromatin interactions occur in three dimensions190–192. vation marks; 4) a temporally coordinated transcrip‑
However, temporal changes in 3D genomic interactions or in both space and time tional activation phase at CT12–CT15 during which
(the fourth dimension) are an emerging area (‘the 4D nucleome’ (REF. 193)). Circadian RNAPII recruitment and pre-mRNA transcript expres‑
cycles in chromosome organization have been observed using circular chromosome
sion peak, followed by histone H3K4me3 and H3K27ac
conformation capture (4C) methods at the Dbp locus, a well-known CLOCK–BMAL1
target gene144. Although chromosome interactions surrounding the Dbp locus were marks; 5) a classic repression phase at CT15–CT18 dur‑
not reported to change during the circadian cycle, the frequency of long-range ing which PER1, PER2 and CRY2 occupancy peaks and
inter-chromosomal interactions did vary. Inter-chromosomal interactions were highest transcription declines; and 6) a transition between the
at the peak of Dbp expression and were lowest at the opposite phase, when Dbp end of repression and the beginning of the next cycle.
expression was low. The Dbp circadian interactome was dependent on Bmal1 expression Interestingly, the co‑­activator CBP interacts with the
because Bmal1–/– knockout cells had lower inter-chromosomal interactions similar to repressor PER2 at the end of the repression phase, and
that seen at the trough of Dbp rhythms144. 4C analysis of a BMAL1‑bound enhancer CBP occupancy has a second peak at this time71, perhaps
upstream of the circadian gene Nr1d1 (which encodes the nuclear receptor REV-ERBα) maintaining an open chromatin state during the tran‑
also found that interactions were largely stable during the circadian cycle148. However, sition from repression to activation. CRY1 occupancy
a global analysis of cohesin and CTCF revealed that cohesin–CTCF co‑binding sites
also bridges this transition period and interacts with
insulated cycling genes with different phases of expression and that cohesin–non-CTCF
sites were associated with high-amplitude circadian cycling148. In addition, genome-wide CLOCK–BMAL1 to begin the next cycle.
chromosome conformation capture (Hi‑C) combined with RNA sequencing expression In contrast to embryonic stem cells, for which
was used in human fibroblasts to search for circadian interactions, and a Laplacian 30% of genes are paused and carry bivalent histone
analysis framework was implemented to quantify dynamic structure–function marks139,140, genes that are not expressed in the mouse
relationships in the genome145. Using multicolour 3D‑FISH, CLOCK and PER2, which have liver are not enriched for paused RNAPII71 (FIG. 5b).
rhythmic and anti-phasic structure–function dynamics, were shown to also have a Interestingly, the histone H3K4me3 promoter mark
correlated rhythm in spatial distance between them145. Finally, using 4C to probe the H19 lags RNAPII occupancy by approximately 1 hour in
imprinting control region in human embryonic stem cells, a large inter-chromosomal cycling genes, and the H3K36me3 elongation mark
interactome was identified146. The interactome was organized by poly(ADP-ribose) lags by approximately 3 hours72, consistent with the
polymerase 1 (PARP1) and CTCF, and, interestingly, active loci enriched for circadian
idea that these marks arise as a consequence of RNAPII
genes were found to be recruited to repressed lamina-associated domains in the nuclear
periphery. In serum-synchronized HCT116 cells, there was a circadian rhythm of transcription. Thus, circadian transcriptional regulators
recruitment of circadian loci to the nuclear envelope that was dependent on PARP1 and are clearly involved in RNAPII recruitment and initi­
CTCF. This recruitment was associated with the acquisition of dimethylation of Lys9 at ation, as are the histone modifications associated with
histone H3 (H3K9me2) repressive marks and the attenuation of transcription. Thus, these events. However, additional steps such as promoter
circadian regulation of transcription also involves the recruitment and repression of proximal pausing, pause release and productive elonga‑
circadian loci to lamina-associated domains in the nuclear periphery146. tion probably have additional regulatory roles (FIG. 7). At
enhancer sites, CLOCK–BMAL1 is a pioneer transcrip‑
tion factor and can bind nucleosomes to promote rhyth‑
Given these circadian rhythms of RNAPII, chromatin mic chromatin remodelling and incorporation of the
states associated with RNAPII transcription have been histone variant H2A.Z141 (FIG. 7). Cycling enhancers also
examined during the circadian cycle71,72,99. FIGURE  5 display circadian rhythms in eRNA expression in con‑
illustrates genome-level views of histone modi­fications junction with tissue-specific transcription factors107,119.
that are characteristic of promoter, enhancer and elon‑ In the case of cycling genes, RNAPII recruitment, initi­
gation states132–138. Histone H3K4me3, H3K9ac and ation and productive elongation are all rhythmic and
H3K27ac modifications are enriched at promoters occur in phase, leading to associated rhythms in histone
and undergo circadian rhythms in histone modifications modifications72. In the case of expressed but non-­cycling
at Per1 (FIG. 5a). On a genome level, circadian rhythms genes, RNAPII recruitment and initiation are also rhyth‑
in RNAPII occupancy as well as histone H3K4me3, mic, leading to rhythmic histone modifications at the
H3K9ac and H3K27ac modifications are present in promoter. However, subsequent steps in pausing, pause
Laplacian analysis the majority of expressed genes irrespective of whether release and productive elongation may no longer be
A framework used in many RNA cycling occurs71. Both RNAPII occupancy and under circadian regulation71. Because RNAPII pausing
disciplines to quantify H3K4me3 marks are highly dynamic and show circa‑ at promoter-proximal regions and pause release are
topologies in which dian oscillations that peak during the early night (CT12) major regulatory steps in transcription, it will be impor‑
autonomous entities reach a
consensus without a central
for all expressed genes, cycling intron RNA genes and tant in the future to determine whether pausing factors
direction. In the 4D nucleome, non-cycling genes (FIG. 5b). such as negative elongation factor (NELF) and DRB-
it is used to describe the In summary, in the mouse liver, there is a highly sensitivity-inducing factor (DSIF), as well as the positive
underlying topology of stereo­typed, time-dependent pattern of core transcrip‑ transcription elongation factor‑b (P‑TEFb) complex, are
genome-wide chromosome
tion factor binding, RNAPII occupancy, transcription under circadian control129,142 (FIG. 7).
conformation capture
interactions in the genome. and chromatin states (FIG. 6). There are six distinctive
phases of the circadian cycle: 1) a poised state at the Conclusions
Lamina-associated domains beginning of the cycle (at CT0) during which CLOCK– The circadian modulation of transcription occurs on a
Regions of condensed BMAL1 and CRY1 bind to E‑box sites in a transcrip‑ genome-wide scale far greater than that seen previously
chromatin that are bound by
the nuclear lamina and are
tionally silent state; 2) a derepression phase at CT4 as by gene expression profiling. The circadian clock in the
enriched for repressive histone the levels of CRY1 decline; 3) a major activation phase liver regulates RNAPII occupancy on a genome-wide
marks such as H3K9me2. at CT8 during which CLOCK–BMAL1 occupancy basis and drives the genome-wide circadian modulation

NATURE REVIEWS | GENETICS ADVANCE ONLINE PUBLICATION | 13


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

of chromatin states. These circadian rhythms in the (BOX 3). Thus, the four-dimensional regulation of the
chromatin state poise the genome for transcription on a genome and its nuclear architecture are ripe for study.
daily basis to coincide with the daily metabolic demands In addition to the pervasive circadian regulation of
of the organism and to optimize the energy utilization of metabolism11,149, strong links between the circadian gene
gene expression143. In addition to the circadian modu‑ network and cell growth and cancer 150–152, immune func‑
lation of chromatin states, the 3D architecture of the tion29,30 and signalling 153 are apparent, and components
nucleus and the interaction of active and repressive chro‑ of this network will be important t­ herapeutic targets in
mosomal domains undergo circadian oscillations144–148 the future.

1. Pittendrigh, C. S. Temporal organization: reflections 23. Lim, C. & Allada, R. Emerging roles for post- 47. Xu, Y. et al. Functional consequences of a CKIδ
of a Darwinian clock-watcher. Annu. Rev. Physiol. 55, transcriptional regulation in circadian clocks. mutation causing familial advanced sleep phase
16–54 (1993). Nat. Neurosci. 16, 1544–1550 (2013). syndrome. Nature 434, 640–644 (2005).
2. Dunlap, J. C. Molecular bases for circadian clocks. Cell 24. Kojima, S. & Green, C. B. Circadian genomics reveal a 48. Preitner, N. et al. The orphan nuclear receptor REV-
96, 271–290 (1999). role for post-transcriptional regulation in mammals. ERBα controls circadian transcription within the
3. Alon, U. Network motifs: theory and experimental Biochemistry 54, 124–133 (2015). positive limb of the mammalian circadian oscillator.
approaches. Nat. Rev. Genet. 8, 450–461 (2007). 25. Allada, R. & Chung, B. Y. Circadian organization of Cell 110, 251–260 (2002).
4. Young, M. W. & Kay, S. A. Time zones: a comparative behavior and physiology in Drosophila. Annu. Rev. 49. Sato, T. et al. A functional genomics strategy reveals
genetics of circadian clocks. Nat. Rev. Genet. 2, Physiol. 72, 605–624 (2010). Rora as a component of the mammalian circadian
702–715 (2001). 26. Hardin, P. E. Molecular genetic analysis of circadian clock. Neuron 43, 527–537 (2004).
5. Bell-Pedersen, D. et al. Circadian rhythms from timekeeping in Drosophila. Adv. Genet. 74, 141–173 50. Zhang, Y. et al. Discrete functions of nuclear receptor
multiple oscillators: lessons from diverse organisms. (2011). Rev-erbα couple metabolism to the clock. Science
Nat. Rev. Genet. 6, 544–556 (2005). 27. Kramer, A. & Merrow, M. (eds) Circadian Clocks 348, 1488–1492 (2015).
6. Rosbash, M. The implications of multiple circadian (Springer, 2013). 51. Novak, B. & Tyson, J. J. Design principles of
clock origins. PLoS Biol. 7, e62 (2009). 28. Eckel-Mahan, K. & Sassone-Corsi, P. Metabolism biochemical oscillators. Nat. Rev. Mol. Cell Biol. 9,
7. Doherty, C. J. & Kay, S. A. Circadian control of global and the circadian clock converge. Physiol. Rev. 93, 981–991 (2008).
gene expression patterns. Annu. Rev. Genet. 44, 107–135 (2013). 52. Mitsui, S., Yamaguchi, S., Matsuo, T., Ishida, Y. &
419–444 (2010). 29. Scheiermann, C., Kunisaki, Y. & Frenette, P. S. Okamura, H. Antagonistic role of E4BP4 and PAR
8. Greenham, K. & McClung, C. R. Integrating circadian Circadian control of the immune system. Nat. Rev. proteins in the circadian oscillatory mechanism.
dynamics with physiological processes in plants. Immunol. 13, 190–198 (2013). Genes Dev. 15, 995–1006 (2001).
Nat. Rev. Genet. 16, 598–610 (2015). 30. Curtis, A. M., Bellet, M. M., Sassone-Corsi, P. & 53. Gachon, F. et al. The loss of circadian PAR bZip
9. Rutter, J., Reick, M. & McKnight, S. L. Metabolism O’Neill, L. A. J. Circadian clock proteins and immunity. transcription factors results in epilepsy. Genes Dev.
and the control of circadian rhythms. Annu. Rev. Immunity 40, 178–186 (2014). 18, 1397–1412 (2004).
Biochem. 71, 307–331 (2002). 31. Johnson, C. H. & Egli, M. Metabolic compensation 54. Ueda, H. R. et al. System-level identification of
10. Tu, B. P. & McKnight, S. L. Metabolic cycles as an and circadian resilience in prokaryotic cyanobacteria. transcriptional circuits underlying mammalian
underlying basis of biological oscillations. Nat. Rev. Annu. Rev. Biochem. 83, 221–247 (2014). circadian clocks. Nat. Genet. 37, 187–192 (2005).
Mol. Cell Biol. 7, 696–701 (2006). 32. LeGates, T. A., Fernandez, D. C. & Hattar, S. Light 55. Etchegaray, J. P., Lee, C., Wade, P. A. & Reppert, S. M.
11. Bass, J. Circadian topology of metabolism. Nature as a central modulator of circadian rhythms, sleep Rhythmic histone acetylation underlies transcription in
491, 348–356 (2012). and affect. Nat. Rev. Neurosci. 15, 443–454 the mammalian circadian clock. Nature 421, 177–182
12. Takahashi, J. S. Finding new clock components: past (2014). (2003).
and future. J. Biol. Rhythms 19, 339–347 (2004). 33. Reddy, A. B. & Rey, G. Metabolic and 56. Curtis, A. M. et al. Histone acetyltransferase-
13. Lowrey, P. & Takahashi, J. Mammalian circadian nontranscriptional circadian clocks: eukaryotes. dependent chromatin remodeling and the vascular
biology: elucidating genome-wide levels of temporal Annu. Rev. Biochem. 83, 165–189 (2014). clock. J. Biol. Chem. 279, 7091–7097 (2004).
organization. Annu. Rev. Genomics Hum. Genet. 5, 34. Asher, G. & Sassone-Corsi, P. Time for food: the 57. Lee, J. et al. Dual modification of BMAL1 by
407–441 (2004). intimate interplay between nutrition, metabolism, SUMO2/3 and ubiquitin promotes circadian activation
14. Takahashi, J. S., Hong, H. K., Ko, C. H. & and the circadian clock. Cell 161, 84–92 (2015). of the CLOCK/BMAL1 complex. Mol. Cell. Biol. 28,
McDearmon, E. L. The genetics of mammalian 35. Cha, J., Zhou, M. & Liu, Y. Mechanism of the 6056–6065 (2008).
circadian order and disorder: implications for Neurospora circadian clock, a FREQUENCY-centric 58. Hosoda, H. et al. CBP/p300 is a cell type-specific
physiology and disease. Nat. Rev. Genet. 9, 764–775 view. Biochemistry 54, 150–156 (2015). modulator of CLOCK/BMAL1‑mediated transcription.
(2008). 36. Hurley, J., Loros, J. J. & Dunlap, J. C. Dissecting the Mol. Brain 2, 34 (2009).
15. Hastings, M. H., Reddy, A. B. & Maywood, E. S. mechanisms of the clock in Neurospora. Methods 59. Doi, M., Hirayama, J. & Sassone-Corsi, P. Circadian
A clockwork web: circadian timing in brain and Enzymol. 551, 29–52 (2015). regulator CLOCK is a histone acetyltransferase. Cell
periphery, in health and disease. Nat. Rev. Neurosci. 37. Shultzaberger, R. K., Boyd, J. S., Diamond, S., 125, 497–508 (2006).
4, 649–661 (2003). Greenspan, R. J. & Golden, S. S. Giving time purpose: 60. Hirayama, J. et al. CLOCK-mediated acetylation of
16. Welsh, D. K., Takahashi, J. S. & Kay, S. A. the Synechococcus elongatus clock in a broader BMAL1 controls circadian function. Nature 450,
Suprachiasmatic nucleus: cell autonomy and network context. Annu. Rev. Genet. 49, 485–505 1086–1090 (2007).
network properties. Annu. Rev. Physiol. 72, 551–577 (2015). 61. Nakahata, Y. et al. The NAD+-dependent deacetylase
(2010). 38. King, D. P. et al. Positional cloning of the mouse SIRT1 modulates CLOCK-mediated chromatin
17. Dibner, C., Schibler, U. & Albrecht, U. The mammalian circadian Clock gene. Cell 89, 641–653 (1997). remodeling and circadian control. Cell 134, 329–340
circadian timing system: organization and coordination 39. Gekakis, N. et al. Role of the CLOCK protein in the (2008).
of central and peripheral clocks. Annu. Rev. Physiol. mammalian circadian mechanism. Science 280, 62. Asher, G. et al. SIRT1 regulates circadian clock gene
72, 517–549 (2010). 1564–1569 (1998). expression through PER2 deacetylation. Cell 134,
18. Mohawk, J. A., Green, C. B. & Takahashi, J. S. Central 40. Kume, K. et al. mCRY1 and mCRY2 are essential 317–328 (2008).
and peripheral circadian clocks in mammals. Annu. components of the negative limb of the circadian 63. Nakahata, Y., Sahar, S., Astarita, G., Kaluzova, M.
Rev. Neurosci. 35, 445–462 (2012). clock feedback loop. Cell 98, 193–205 (1999). & Sassone-Corsi, P. Circadian control of the NAD+
19. Albrecht, U. Timing to perfection: the biology of 41. Shearman, L. P. et al. Interacting molecular loops salvage pathway by CLOCK–SIRT1. Science 324,
central and peripheral circadian clocks. Neuron 74, in the mammalian circadian clock. Science 288, 654–657 (2009).
246–260 (2012). 1013–1019 (2000). 64. Ramsey, K. M. et al. Circadian clock feedback cycle
20. Gerber, A. et al. Blood-borne circadian signal 42. Lee, C., Etchegaray, J. P., Cagampang, F. R., through NAMPT-mediated NAD+ biosynthesis.
stimulates daily oscillations in actin dynamics and Loudon, A. S. & Reppert, S. M. Posttranslational Science 324, 651–654 (2009).
SRF activity. Cell 152, 492–503 (2013). mechanisms regulate the mammalian circadian clock. 65. Katada, S. & Sassone-Corsi, P. The histone
21. Kornmann, B., Schaad, O., Bujard, H., Takahashi, J. S. Cell 107, 855–867 (2001). methyltransferase MLL1 permits the oscillation of
& Schibler, U. System-driven and oscillator-dependent 43. Gallego, M. & Virshup, D. M. Post-translational circadian gene expression. Nat. Struct. Mol. Biol. 17,
circadian transcription in mice with a conditionally modifications regulate the ticking of the circadian 1414–1421 (2010).
active liver clock. PLoS Biol. 5, e34 (2007). clock. Nat. Rev. Mol. Cell Biol. 8, 139–148 (2007). 66. Aguilar-Arnal, L., Katada, S., Orozco-Solis, R. &
22. Zhang, R., Lahens, N. F., Ballance, H. I., Hughes, M. E. 44. Lowrey, P. L. & Takahashi, J. S. Genetics of circadian Sassone-Corsi, P. NAD+–SIRT1 control of H3K4
& Hogenesch, J. B. A circadian gene expression atlas rhythms in mammalian model organisms. Adv. Genet. trimethylation through circadian deacetylation of
in mammals: implications for biology and medicine. 74, 175–230 (2011). MLL1. Nat. Struct. Mol. Biol. 22, 312–318 (2015).
Proc. Natl Acad. Sci. USA 45, 16219–16224 45. Preussner, M. & Heyd, F. Post-transcriptional control 67. Valekunja, U. K. et al. Histone methyltransferase
(2014). of the mammalian circadian clock: implications for MLL3 contributes to genome-scale circadian
This paper shows that 43% of protein-coding genes health and disease. Pflugers Arch. 468, 983–991 transcription. Proc. Natl Acad. Sci. USA 110,
in the genome show circadian oscillation in at least (2016). 1554–1559 (2013).
one tissue. The authors also estimate that 56% of 46. Toh, K. et al. An hPer2 phosphorylation site mutation 68. Ditacchio, L. et al. Histone lysine demethylase
the best-selling drugs directly target the products in familial advanced sleep phase syndrome. Science JARID1a activates CLOCK–BMAL1 and influences the
of rhythmic genes. 291, 1040–1043 (2001). circadian clock. Science 333, 1881–1885 (2011).

14 | ADVANCE ONLINE PUBLICATION www.nature.com/nrg


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

69. Nam, H. J. et al. Phosphorylation of LSD1 by PKCα is 93. Zhou, M., Kim, J. K., Eng, G. W. L., Forger, D. B. & 115. Hughes, M. E. et al. Harmonics of circadian gene
crucial for circadian rhythmicity and phase resetting. Virshup, D. M. A Period2 phosphoswitch regulates transcription in mammals. PLoS Genet. 5, e1000442
Mol. Cell 53, 791–805 (2014). and temperature compensates circadian period. (2009).
70. Lande-Diner, L., Boyault, C., Kim, J. Y. & Weitz, C. J. Mol. Cell 60, 1–13 (2015). 116. Du, N.‑H., Arpat, A. B., De Matos, M. & Gatfield, D.
A positive feedback loop links circadian clock factor 94. Hirano, A. et al. FBXL21 regulates oscillation of MicroRNAs shape circadian hepatic gene expression
CLOCK–BMAL1 to the basic transcriptional machinery. the circadian clock through ubiquitination and on a transcriptome-wide scale. eLife 3, e02510
Proc. Natl Acad. Sci. USA 110, 16021–16026 (2013). stabilization of cryptochromes. Cell 152, 1106–1118 (2014).
71. Koike, N. et al. Transcriptional architecture and (2013). 117. Core, L. J., Waterfall, J. J. & Lis, J. T. Nascent RNA
chromatin landscape of the core circadian clock in 95. Yoo, S.‑H. et al. Competing E3 ubiquitin ligases sequencing reveals widespread pausing and divergent
mammals. Science 338, 349–354 (2012). govern circadian periodicity by degradation of CRY initiation at human promoters. Science 322,
72. Le Martelot, G. et al. Genome-wide RNA polymerase II in nucleus and cytoplasm. Cell 152, 1091–1105 1845–1848 (2008).
profiles and RNA accumulation reveal kinetics of (2013). 118. Kim, T.‑K. & Shiekhattar, R. Architectural and
transcription and associated epigenetic changes References 94 and 95 identify FBXL21 as a functional commonalities between enhancers and
during diurnal cycles. PLoS Biol. 10, e1001442 component of a second E3 ubiquitin ligase complex promoters. Cell 162, 948–959 (2015).
(2012). for the CRY circadian repressor proteins. 119. Fang, B. et al. Circadian enhancers coordinate multiple
73. Brown, S. A. et al. PERIOD1‑associated proteins 96. Maywood, E. S. et al. Tuning the period of the phases of rhythmic gene transcription in vivo. Cell
modulate the negative limb of the mammalian mammalian circadian clock: additive and independent 159, 1140–1152 (2014).
circadian oscillator. Science 308, 693–696 effects of CK1εTau and Fbxl3Afh mutations on mouse This paper identifies circadian eRNAs using
(2005). circadian behavior and molecular pacemaking. GRO-seq.
74. Padmanabhan, K., Robles, M., Westerling, T. & J. Neurosci. 31, 1539–1544 (2011). 120. Anafi, R. C. et al. Machine learning helps identify
Weitz, C. Feedback regulation of transcriptional 97. Rey, G. et al. Genome-wide and phase-specific DNA- CHRONO as a circadian clock component. PLoS Biol.
termination by the mammalian circadian clock binding rhythms of BMAL1 control circadian output 12, e1001840 (2014).
PERIOD complex. Science 337, 599–602 (2012). functions in mouse liver. PLoS Biol. 9, e1000595 121. Annayev, Y. et al. Gene model 129 (Gm129) encodes
75. Kim, J. Y., Kwak, P. B. & Weitz, C. J. Specificity in (2011). a novel transcriptional repressor that modulates
circadian clock feedback from targeted reconstitution 98. Menet, J. S., Rodriguez, J., Abruzzi, K. C. & circadian gene expression. J. Biol. Chem. 289,
of the NuRD corepressor. Mol. Cell 56, 738–748 Rosbash, M. Nascent-seq reveals novel features of 5013–5024 (2014).
(2014). mouse circadian transcriptional regulation. eLife 1, 122. Goriki, A. et al. A novel protein, CHRONO, functions
76. Etchegaray, J. P. et al. The polycomb group protein e00011 (2012). as a core component of the mammalian circadian
EZH2 is required for mammalian circadian clock 99. Vollmers, C. et al. Circadian oscillations of protein- clock. PLoS Biol. 12, e1001839 (2014).
function. J. Biol. Chem. 281, 21209–21215 (2006). coding and regulatory RNAs in a highly dynamic 123. Preußner, M. et al. Rhythmic U2af26 alternative
77. Nguyen, K. D. et al. Circadian gene Bmal1 regulates mammalian liver epigenome. Cell Metab. 16, splicing controls PERIOD1 stability and the circadian
diurnal oscillations of Ly6Chi inflammatory monocytes. 833–845 (2012). clock in mice. Mol. Cell 54, 651–662 (2014).
Science 341, 1483–1488 (2013). 100. Yoshitane, H. et al. CLOCK-controlled polyphonic 124. Kojima, S., Sher-Chen, E. L. & Green, C. B. Circadian
78. Tamayo, A. G., Duong, H. A., Robles, M. S., Mann, M. regulation of circadian rhythms through canonical and control of mRNA polyadenylation dynamics regulates
& Weitz, C. J. Histone monoubiquitination by Clock– noncanonical E‑boxes. Mol. Cell. Biol. 34, 1776–1787 rhythmic protein expression. Genes Dev. 26,
Bmal1 complex marks Per1 and Per2 genes for (2014). 2724–2736 (2012).
circadian feedback. Nat. Struct. Mol. Biol. 22, References 71, 72 and 97–100 define the This paper presents an important example of
759–766 (2015). circadian transcriptome, cistrome and epigenome circadian post-transcriptional regulation.
79. Duong, H., Robles, M., Knutti, D. & Weitz, C. in the mouse liver. 125. Janich, P., Arpat, A. B., Castelo-Szekely, V., Lopes, M.
A molecular mechanism for circadian clock negative 101. Ripperger, J. A. & Schibler, U. Rhythmic CLOCK– & Gatfield, D. Ribosome profiling reveals the rhythmic
feedback. Science 332, 1436–1439 (2011). BMAL1 binding to multiple E‑box motifs drives liver translatome and circadian clock regulation by
80. Duong, H. A. & Weitz, C. J. Temporal orchestration circadian Dbp transcription and chromatin transitions. upstream open reading frames. Genome Res. 25,
of repressive chromatin modifiers by circadian clock Nat. Genet. 38, 369–374 (2006). 1848–1859 (2015).
Period complexes. Nat. Struct. Mol. Biol. 21, 102. Stratmann, M., Suter, D. M., Molina, N., Naef, F. & 126. Jang, C., Lahens, N. F., Hogenesch, J. B. & Sehgal, A.
126–132 (2014). Schibler, U. Circadian Dbp transcription relies on Ribosome profiling reveals an important role for
References 73–75 and 78–80 describe the PER highly dynamic BMAL1–CLOCK interaction with translational control in circadian gene expression.
repressor complex. E boxes and requires the proteasome. Mol. Cell 48, Genome Res. 25, 1836–1847 (2015).
81. Lowrey, P. L. et al. Positional syntenic cloning and 277–287 (2012). 127. Sims, R. J. III, Belotserkovskaya, R. & Reinberg, D.
functional characterization of the mammalian 103. Thomas, D. & Tyers, M. Transcriptional regulation: Elongation by RNA polymerase II: the short and long
circadian mutation tau. Science 288, 483–492 kamikaze activators. Curr. Biol. 10, R341–R343 of it. Genes Dev. 18, 2437–2468 (2004).
(2000). (2000). 128. Fuda, N., Ardehali, M. & Lis, J. Defining mechanisms
82. Shirogane, T., Jin, J., Ang, X. & Harper, J. SCFβ-TRCP 104. Ukai-Tadenuma, M. et al. Delay in feedback repression that regulate RNA polymerase II transcription in vivo.
controls clock-dependent transcription via casein by Cryptochrome 1 is required for circadian clock Nature 461, 186–192 (2009).
kinase 1‑dependent degradation of the mammalian function. Cell 144, 268–281 (2011). 129. Jonkers, I. & Lis, J. T. Getting up to speed with
period‑1 (Per1) protein. J. Biol. Chem. 280, 105. Shimomura, K. et al. Usf1, a suppressor of the transcription elongation by RNA polymerase II.
26863–26872 (2005). circadian Clock mutant, reveals the nature of the DNA- Nat. Rev. Mol. Cell Biol. 16, 167–177 (2015).
83. Reischl, S. et al. β‑TrCP1‑mediated degradation of binding of the CLOCK:BMAL1 complex in mice. eLife 130. Jones, J. C. et al. C‑terminal repeat domain kinase I
PERIOD2 is essential for circadian dynamics. 2, e00426 (2013). phosphorylates Ser2 and Ser5 of RNA polymerase II
J. Biol. Rhythms 22, 375–386 (2007). 106. Stashi, E. et al. SRC‑2 is an essential coactivator for C‑terminal domain repeats. J. Biol. Chem. 279,
84. Meng, Q. J. et al. Setting clock speed in mammals: orchestrating metabolism and circadian rhythm. 24957–24964 (2004).
the CK1ε tau mutation in mice accelerates circadian Cell Rep. 6, 633–645 (2014). 131. Chapman, R. D. et al. Transcribing RNA polymerase II
pacemakers by selectively destabilizing PERIOD 107. Perelis, M. et al. Pancreatic β cell enhancers regulate is phosphorylated at CTD residue serine‑7. Science
proteins. Neuron 58, 78–88 (2008). rhythmic transcription of genes controlling insulin 318, 1780–1782 (2007).
85. Busino, L. et al. SCFFbxl3 controls the oscillation secretion. Science 350, aac4250 (2015). 132. Kim, T. H. et al. A high-resolution map of active
of the circadian clock by directing the degradation of This paper defines the circadian cistrome in promoters in the human genome. Nature 436,
cryptochrome proteins. Science 316, 900–904 pancreatic β-cells. 876–880 (2005).
(2007). 108. Hardison, R. C. & Taylor, J. Genomic approaches 133. Guenther, M. G., Levine, S. S., Boyer, L. A.,
86. Siepka, S. et al. Circadian mutant Overtime reveals towards finding cis-regulatory modules in animals. Jaenisch, R. & Young, R. A. A chromatin landmark and
F‑box protein FBXL3 regulation of Cryptochrome Nat. Rev. Genet. 13, 469–483 (2012). transcription initiation at most promoters in human
and Period gene expression. Cell 129, 1011–1023 109. Lamia, K. et al. Cryptochromes mediate rhythmic cells. Cell 130, 77–88 (2007).
(2007). repression of the glucocorticoid receptor. Nature 480, 134. Barski, A. et al. High-resolution profiling of histone
87. Godinho, S. et al. The after-hours mutant reveals a 552–556 (2011). methylations in the human genome. Cell 129,
role for Fbxl3 in determining mammalian circadian 110. Cho, H. et al. Regulation of circadian behaviour and 823–837 (2007).
period. Science 316, 897–900 (2007). metabolism by REV-ERB‑α and REV-ERB‑β. Nature 135. Li, B., Carey, M. & Workman, J. L. The role of
88. Lamia, K. A. et al. AMPK regulates the circadian clock 485, 123–127 (2012). chromatin during transcription. Cell 128, 707–719
by cryptochrome phosphorylation and degradation. 111. Feng, D. et al. A circadian rhythm orchestrated by (2007).
Science 326, 437–440 (2009). histone deacetylase 3 controls hepatic lipid 136. Creyghton, M. P. et al. Histone H3K27ac separates
89. Etchegaray, J. P. et al. Casein kinase 1 delta regulates metabolism. Science 331, 1315–1319 (2011). active from poised enhancers and predicts
the pace of the mammalian circadian clock. Mol. Cell. 112. Zhu, B. et al. Coactivator-dependent oscillation of developmental state. Proc. Natl Acad. Sci. USA 107,
Biol. 29, 3853–3866 (2009). chromatin accessibility dictates circadian gene 21931–21936 (2010).
90. Lee, H. M. et al. The period of the circadian oscillator amplitude via REV-ERB loading. Mol. Cell 60, 137. Ong, C. T. & Corces, V. G. Enhancer function: new
is primarily determined by the balance between casein 769–783 (2015). insights into the regulation of tissue-specific gene
kinase 1 and protein phosphatase 1. Proc. Natl Acad. 113. Ameur, A. et al. Total RNA sequencing reveals nascent expression. Nat. Rev. Genet. 12, 283–293 (2011).
Sci. USA 108, 16451–16456 (2011). transcription and widespread co‑transcriptional 138. Rada-Iglesias, A. et al. A unique chromatin signature
91. Lee, Y., Chen, R., Lee, H. M. & Lee, C. Stoichiometric splicing in the human brain. Nat. Struct. Mol. Biol. 18, uncovers early developmental enhancers in humans.
relationship among clock proteins determines 1435–1440 (2011). Nature 470, 279–283 (2011).
robustness of circadian rhythms. J. Biol. Chem. 286, 114. Gaidatzis, D., Burger, L., Florescu, M. & Stadler, M. B. 139. Bernstein, B. E. et al. A bivalent chromatin structure
7033–7042 (2011). Analysis of intronic and exonic reads in RNA-seq marks key developmental genes in embryonic stem
92. D’Alessandro, M. et al. A tunable artificial circadian data characterizes transcriptional and post- cells. Cell 125, 315–326 (2006).
clock in clock-defective mice. Nat. Commun. 6, 8587 transcriptional regulation. Nat. Biotechnol. 33, 140. Rahl, P. B. et al. c‑Myc regulates transcriptional pause
(2015). 722–729 (2015). release. Cell 141, 432–445 (2010).

NATURE REVIEWS | GENETICS ADVANCE ONLINE PUBLICATION | 15


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

141. Menet, J. S., Pescatore, S. & Rosbash, M. substrate recognition for an SCF E3 ubiquitin studies to pathophysiology: the melatonin receptor 1B
CLOCK:BMAL1 is a pioneer-like transcription factor. ligase complex. (MT2) paradigm. Cell Metab. 24, 345–347 (2016).
Genes Dev. 28, 8–13 (2014). 161. Schmalen, I. et al. Interaction of circadian clock 182. Hu, Y. et al. GWAS of 89,283 individuals identifies
This paper shows that CLOCK and BMAL1 act as proteins CRY1 and PER2 Is modulated by zinc binding genetic variants associated with self-reporting of being
pioneer transcription factors. and disulfide bond formation. Cell 157, 1203–1215 a morning person. Nat. Commun. 7, 10448 (2016).
142. Adelman, K. & Lis, J. T. Promoter-proximal pausing of (2014). 183. Jones, S. E. et al. Genome-wide association analyses in
RNA polymerase II: emerging roles in metazoans. 162. Nangle, S. N. et al. Molecular assembly of the period– 128,266 individuals identifies new morningness and
Nat. Rev. Genet. 13, 720–731 (2012). cryptochrome circadian transcriptional repressor sleep duration loci. PLoS Genet. 12, e1006125
143. Wang, G.‑Z. et al. Cycling transcriptional networks complex. eLife 3, e03674 (2014). (2016).
reduce the synthetic cost of genomes. Cell Rep. 13, 163. Chen, R. et al. Rhythmic PER abundance defines a 184. Lane, J. M. et al. Genome-wide association analysis
1868–1880 (2015). critical nodal point for negative feedback within the identifies novel loci for chronotype in 100,420
144. Aguilar-Arnal, L. et al. Cycles in spatial and temporal circadian clock mechanism. Mol. Cell 36, 417–430 individuals from the UK Biobank. Nat. Commun. 7,
chromosomal organization driven by the circadian (2009). 10889 (2016).
clock. Nat. Struct. Mol. Biol. 20, 1206–1213 (2013). 164. Wu, D., Potluri, N., Lu, J., Kim, Y. & Rastinejad, F. References 182–184 report genome-wide
This paper is the first to provide evidence for Structural integration in hypoxia-inducible factors. association loci, including known circadian genes,
circadian rhythms in chromosomal conformation. Nature 524, 303–308 (2015). for morningness or early chronotype.
145. Chen, H. et al. Functional organization of the human 165. Wu, D. & Rastinejad, F. Structural characterization 185. Doi, M. et al. Circadian regulation of intracellular
4D nucleome. Proc. Natl Acad. Sci. USA 112, of mammalian bHLH–PAS transcription factors. G‑protein signalling mediates intercellular synchrony
8002–8007 (2015). Curr. Opin. Struct. Biol. 43, 1–9 (2016). and rhythmicity in the suprachiasmatic nucleus.
146. Zhao, H. et al. PARP1- and CTCF-mediated 166. Wang, Z., Wu, Y., Li, L. & Su, X.‑D. Intermolecular Nat. Commun. 2, 327 (2011).
interactions between active and repressed chromatin recognition revealed by the complex structure of human 186. Aton, S. J., Colwell, C. S., Harmar, A. J., Waschek, J.
at the lamina promote oscillating transcription. CLOCK–BMAL1 basic helix–loop–helix domains with & Herzog, E. D. Vasoactive intestinal polypeptide
Mol. Cell 59, 984–997 (2015). E‑box DNA. Cell Res. 23, 213–224 (2012). mediates circadian rhythmicity and synchrony in
This paper shows that there is a circadian rhythm 167. Nangle, S., Xing, W. & Zheng, N. Crystal structure of mammalian clock neurons. Nat. Neurosci. 8, 476–483
of recruitment of circadian loci to the nuclear mammalian cryptochrome in complex with a small (2005).
envelope that is dependent on PARP1 and CTCF. molecule competitor of its ubiquitin ligase. Cell Res. 187. Chabas, D., Taheri, S., Renier, C. & Mignot, E. The
147. Aguilar-Arnal, L. & Sassone-Corsi, P. Chromatin 23, 1417–1419 (2013). genetics of narcolepsy. Annu. Rev. Genomics Hum.
landscape and circadian dynamics: spatial and 168. Hirota, T. et al. Identification of small molecule Genet. 4, 459–483 (2003).
temporal organization of clock transcription. activators of cryptochrome. Science 337, 1094–1097 188. Gottlieb, D. J. et al. Novel loci associated with usual
Proc. Natl Acad. Sci. USA 112, 6863–6870 (2015). (2012). sleep duration: the CHARGE Consortium genome-wide
148. Xu, Y. et al. Long-range chromosome interactions 169. Ye, R., Selby, C., Ozturk, N., Annayev, Y. & Sancar, A. association study. Mol. Psychiatry 20, 1232–1239
mediated by cohesin shape circadian gene expression. Biochemical analysis of the canonical model for the (2015).
PLoS Genet. 12, e1005992 (2016). mammalian circadian clock. J. Biol. Chem. 286, 189. Spada, J. et al. Genome-wide association analysis of
149. Feng, D. & Lazar, M. A. Clocks, metabolism, and the 25891–25902 (2011). actigraphic sleep phenotypes in the LIFE Adult Study.
epigenome. Mol. Cell 47, 158–167 (2012). 170. Chiou, Y. Y. et al. Mammalian Period represses and J. Sleep Res. http://dx.doi.org/10.1111/jsr.12421
150. Sancar, A. et al. Circadian clock, cancer, and de‑represses transcription by displacing CLOCK– (2016).
chemotherapy. Biochemistry 54, 110–123 (2014). BMAL1 from promoters in a cryptochrome-dependent 190. Levine, M., Cattoglio, C. & Tjian, R. Looping back to
151. Masri, S., Kinouchi, K. & Sassone-Corsi, P. Circadian manner. Proc. Natl Acad. Sci. USA 113, E6072–E6079 leap forward: transcription enters a new era. Cell 157,
clocks, epigenetics, and cancer. Curr. Opin. Oncol. 27, (2016). 13–25 (2014).
50–56 (2015). 171. Menet, J. S., Abruzzi, K. C., Desrochers, J., 191. Dekker, J. & Mirny, L. The 3D genome as moderator of
152. Papagiannakopoulos, T. et al. Circadian rhythm Rodriguez, J. & Rosbash, M. Dynamic PER repression chromosomal communication. Cell 164, 1110–1121
disruption promotes lung tumorigenesis. Cell Metab. mechanisms in the Drosophila circadian clock: from (2016).
24, 324–331 (2016). on‑DNA to off-DNA. Genes Dev. 24, 358–367 192. Bonev, B. & Cavalli, G. Organization and function of
This paper demonstrates that circadian disruption (2010). the 3D genome. Nat. Rev. Genet. 17, 661–678
promotes tumorigenesis in a well-validated 172. Ye, R. et al. Dual modes of CLOCK:BMAL1 inhibition (2016).
genetically engineered mouse model of lung cancer. mediated by cryptochrome and Period proteins in 193. Cremer, T. et al. The 4D nucleome: evidence for a
153. Schibler, U. et al. Clock-talk: interactions between the mammalian circadian clock. Genes Dev. 28, dynamic nuclear landscape based on co‑aligned active
central and peripheral circadian oscillators in 1989–1998 (2014). and inactive nuclear compartments. FEBS Lett. 589,
mammals. Cold Spring Harb. Symp. Quant. Biol. 80, 173. Jones, C. R., Huang, A. L., Ptacek, L. J. & Fu, Y. H. 2931–2943 (2015).
223–232 (2015). Genetic basis of human circadian rhythm disorders.
154. Crane, B. R. & Young, M. W. Interactive features of Exp. Neurol. 243, 28–33 (2013). Acknowledgements
proteins composing eukaryotic circadian clocks. 174. Bouatia-Naji, N. et al. A variant near MTNR1B is The author thanks three anonymous reviewers for their con‑
Annu. Rev. Biochem. 83, 191–219 (2014). associated with increased fasting plasma glucose levels structive comments. Apologies to those whose work was not
155. Gustafson, C. L. & Partch, C. L. Emerging models for and type 2 diabetes risk. Nat. Genet. 41, 89–94 cited owing to content and length constraints. The author
the molecular basis of mammalian circadian timing. (2009). thanks N. Koike and T.-K. Kim for critical contributions to the
Biochemistry 54, 134–149 (2015). 175. Lyssenko, V. et al. Common variant in MTNR1B research presented here. This work was supported by
156. Huang, N. et al. Crystal structure of the heterodimeric associated with increased risk of type 2 diabetes the Howard Hughes Medical Institute and US National
CLOCK:BMAL1 transcriptional activator complex. and impaired early insulin secretion. Nat. Genet. 41, Institutes of Health (NIH) grants R01AG045795 (to J.S.T.)
Science 337, 189–194 (2012). 82–88 (2009). and R21MH107672 (to G. Konopka and J.S.T.).
This paper is the first to describe the crystal 176. Prokopenko, I. et al. Variants in MTNR1B influence
structure of a heterodimeric bHLH–PAS fasting glucose levels. Nat. Genet. 41, 77–81 Competing interests statement
transcription factor. (2009). The author declares competing interests: see Web version
157. Hennig, S. et al. Structural and functional analyses 177. Dupuis, J. et al. New genetic loci implicated in for details.
of PAS domain interactions of the clock proteins fasting glucose homeostasis and their impact on
Drosophila PERIOD and mouse PERIOD2. PLoS Biol. type 2 diabetes risk. Nat. Genet. 42, 105–116
7, e1000094 (2009). (2010). DATABASES
158. Kucera, N. et al. Unwinding the differences of the 178. Pevet, P. & Challet, E. Melatonin: both master clock CircaDB: http://circadb.hogeneschlab.org
mammalian PERIOD clock proteins from crystal output and internal time-giver in the circadian RSCB Protein Data Bank: http://www.rcsb.org/pdb/home/
structure to cellular function. Proc. Natl Acad. Sci. clocks network. J. Physiol. Paris 105, 170–182 home.do
USA 109, 3311–3316 (2012). (2011).
159. Czarna, A. et al. Structures of Drosophila cryptochrome 179. Bonnefond, A. et al. Rare MTNR1B variants impairing FURTHER INFORMATION
and mouse cryptochrome1 provide insight into melatonin receptor 1B function contribute to type 2 BioGPS Circadian Genomics Screen: http://biogps.org/
circadian function. Cell 153, 1394–1405 (2013). diabetes. Nat. Genet. 44, 297–301 (2012). plugin/406/circadian-genomics-screen/
160. Xing, W. et al. SCFFBXL3 ubiquitin ligase targets 180. Tuomi, T. et al. Increased melatonin signaling is a risk Circadian genes as listed in GeneCards: http://www.
cryptochromes at their cofactor pocket. Nature 496, factor for type 2 diabetes. Cell Metab. 23, 1067–1077 genecards.org/Search/Keyword?queryString=circadian%20
64–68 (2013). (2016). clock
This study reports the crystal structure of CRY2 181. Bonnefond, A., Karamitri, A., Jockers, R. & Froguel, P. ALL LINKS ARE ACTIVE IN THE ONLINE PDF
in complex with FBXL3, revealing a novel mode of The difficult journey from genome-wide association

16 | ADVANCE ONLINE PUBLICATION www.nature.com/nrg


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.

You might also like