You are on page 1of 6

Homer1a is a core brain molecular correlate

of sleep loss
Stéphanie Maret*, Stéphane Dorsaz*, Laure Gurcel*, Sylvain Pradervand†, Brice Petit*, Corinne Pfister*,
Otto Hagenbuchle†, Bruce F. O’Hara‡, Paul Franken*, and Mehdi Tafti*§
*Center for Integrative Genomics and †Lausanne DNA Array Facility, University of Lausanne, Génopode, CH-1015 Lausanne, Switzerland; and ‡Department
of Biology, University of Kentucky, Lexington, KY 40506-0225

Communicated by Michael Rosbash, Brandeis University, Waltham, MA, October 24, 2007 (received for review October 15, 2007)

Sleep is regulated by a homeostatic process that determines its need takes these interacting factors into account. We show that short-
and by a circadian process that determines its timing. By using sleep term sleep loss induces changes in brain gene expression for a few
deprivation and transcriptome profiling in inbred mouse strains, we genes only. These genes are all part of a highly specific pathway
show that genetic background affects susceptibility to sleep loss at involved in neuronal protection and recovery after waking-induced
the transcriptional level in a tissue-dependent manner. In the brain, glutamate overstimulation.
Homer1a expression best reflects the response to sleep loss. Time-
course gene expression analysis suggests that 2,032 brain transcripts Results
are under circadian control. However, only 391 remain rhythmic when We have previously reported (7) that the dynamics of sleep need
mice are sleep-deprived at four time points around the clock, sug- varies strongly among inbred mouse strains, with AKR/J (AK) mice
gesting that most diurnal changes in gene transcription are, in fact, showing a dramatic increase in delta power after 6-h sleep depri-
sleep–wake-dependent. By generating a transgenic mouse line, we vation, whereas DBA/2J (D2) mice have a blunted response.
show that in Homer1-expressing cells specifically, apart from Through quantitative linkage analysis, a significant quantitative
Homer1a, three other activity-induced genes (Ptgs2, Jph3, and Nptx2) trait locus (QTL) was identified on mouse chromosome 13 (Dps1:
are overexpressed after sleep loss. All four genes play a role in delta power in slow-wave sleep 1) that explains ⬎50% of variance
recovery from glutamate-induced neuronal hyperactivity. The consis- in delta power after sleep deprivation in BXD recombinant inbred
tent activation of Homer1a suggests a role for sleep in intracellular lines (RIs) derived from inbred mouse strains C57BL/6J (B6) and
calcium homeostasis for protecting and recovering from the neuronal D2 (7). The best polymorphic marker associated was D13Mit126 at
activation imposed by wakefulness. 46.5 cM (95% CI ⫽ 25 cM), suggesting a large QTL region (⬇38
Mb). However, based on the most recent high-resolution, single-
homeostasis 兩 microarray 兩 mRNA tagging 兩 sleep deprivation 兩 nucleotide polymorphism (SNP) genetic map of BXD RIs (10)
sleep function (http://gscan.well.ox.ac.uk/gs/strains.cgi), the smallest differential
region corresponds to an 11-Mb sequence flanked by SNPs
rs3669221 and rs397202. According to the Ensemble database (Mus
T wo main processes regulate sleep. A homeostatic process
regulates sleep need and intensity according to the time
spent awake or asleep. A circadian process regulates the appro-
musculus release 46), this region contains 33 known genes, 15
potential unknown coding sequences, and three pseudogenes.
priate timing of sleep and wakefulness across the 24-h day. A Among these, the short splice variant of the Homer1 (Homer1a)
highly reliable index of the homeostatic process is provided by gene is the only transcript in the region that was previously reported
the amplitude and prevalence of delta (1- to 4-Hz) oscillations to be among the up-regulated genes after sleep deprivation (11–14).
in the electroencephalogram (EEG) of nonrapid eye movement However, the specificity of this finding, compared with other gene
(NREM) sleep (hereafter, ‘‘delta power’’). Delta power is high expression changes after sleep loss, has not yet been established.
at sleep onset and decreases during sleep, in parallel with sleep To confirm and verify whether differences in sleep need between
depth. Sleep deprivations and naps induce a predictable increase mouse strains are regulated at the transcriptional level, mice of
or decrease, respectively, in delta power during subsequent three genotypes (AK, B6, and D2) were deprived of sleep by gentle
sleep. The interaction between homeostatic and circadian pro- handling [see supporting information (SI) Materials and Methods]
cesses is mathematically described in the two-process model of for 6 h, starting at light onset, and killed at the same time of day
sleep regulation, which provides a framework for prediction and [Zeitgeber time (ZT)6] with their non-sleep-deprived controls.
interpretation of a large body of experimental data (1). Because it is believed that sleep fulfills a brain-specific function, we
Among hypotheses concerning the physiological function of also sampled the liver as a peripheral reference organ. Microarray
waking-induced changes in sleep, the most compelling suggests that results were analyzed by a two-way ANOVA, with strain and
sleep plays a key role in synaptic plasticity (2, 3). More specifically, condition as main factors. Venn diagrams summarizing the data are
EEG delta power during NREM sleep has been shown to play a shown in Fig. 1 A and B and suggest that very few genes show
critical role in learning-induced plasticity (4–6). In general, the consistent changes in expression across genetic backgrounds. We
prediction is that local neural activation due to specific behavioral use the terms ‘‘consistent’’ and ‘‘reliable’’ hereafter only for tran-
(cognitive) demands imposes a burden on the brain which neces-
sitates sleep and which is reflected by the EEG delta power.
Author contributions: S.D. and L.G. contributed equally to this work; O.H., P.F., and M.T.
On the basis of mathematical modeling and experimental data, designed research; S.M., S.D., L.G., B.P., C.P., O.H., and B.F.O. performed research; S.M., S.D.,
we have shown that sleep need, as indexed by the EEG delta power, L.G., S.P., B.F.O., P.F., and M.T. analyzed data; and S.M., S.D., L.G., S.P., P.F., and M.T. wrote
is under genetic control (7), which is of direct relevance for the paper
explaining the interindividual vulnerability to sleep loss in human The authors declare no conflict of interest.
subjects (8, 9). However, deciphering the molecular bases of sleep Data deposition: The data reported in this paper have been deposited in the Gene
need is rendered difficult because the contributions of the homeo- Expression Omnibus (GEO) database, www.ncbi.nlm.nih.gov/geo (accession no. GSE9444).
static and circadian processes are difficult to separate and because §To whom correspondence should be addressed. E-mail: mehdi.tafti@unil.ch.
the impact of genetic background on brain gene expression is poorly This article contains supporting information online at www.pnas.org/cgi/content/full/
understood. From a series of gene-profiling experiments, we here 0710131104/DC1.
report a comprehensive transcriptome analysis that specifically © 2007 by The National Academy of Sciences of the USA

20090 –20095 兩 PNAS 兩 December 11, 2007 兩 vol. 104 兩 no. 50 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0710131104
part of a general stress–response pathway induced by enforced
wakefulness in most organs, and, therefore, is not specific to the
brain. The most significant decrease was found for the cold-induced
RNA binding protein (Cirbp) in both tissues, again suggesting a
general rather than a brain-specific implication (SI Tables 1 and 2).
In brain, the expression of Homer1a was most affected by both
genotype and sleep loss. In situ hybridization analysis confirmed the
strong induction of Homer1a transcript after sleep deprivation and
indicated a restricted up-regulation in cortex, striatum, and hip-
pocampus (Fig. 1C). Homer1 encodes several transcripts by alter-
native splicing, among which long forms are constitutively expressed
while two short forms, Homer1a and Ania-3, are activity-induced.
These postsynpatic density proteins bind calcium-signaling mole-
cules and have been implicated in synaptic plasticity.
In contrast to most other activity-induced genes, Homer1a over-
expression seemed highly specific. For instance, brain-derived
neurotrophic factor (Bdnf ), a plasticity-related gene, showed a
significant increase in AK and D2 only; the immediate early gene
Fos did not show a significant increase in D2; and Arc, Egr1, and
Egr3 were induced by sleep deprivation mainly in AK (SI Table 1).
Thus, comparisons among genotypes, and between brain and liver,
identified Homer1a as the most specific transcriptional index of the
whole brain in response to sleep loss.
We then analyzed the time course of Homer1a induction by
real-time quantitative RT-PCR in a dose–response experiment.
Mice of all three strains were sleep-deprived for 1, 3, and 6 h and
killed together with their time-matched controls. An additional
group, which was also sleep-deprived for 6 h, was killed 2 h into
recovery sleep. As shown in Fig. 1D, Homer1a is rapidly and
strongly induced by sleep deprivation in a dose-dependent manner.
Parallel to this increase, Homer1a expression decreases with in-

NEUROSCIENCE
creasing accumulated sleep in non-sleep-deprived animals, and its
relative level remains significantly higher than in the time-matched
controls after only 2 h of recovery sleep, indicating that the time
course of Homer1a expression closely parallels sleep need. How-
Fig. 1. Transcriptome analysis of the brain and liver after sleep deprivation
ever, the fold change of Homer1a expression after 6-h sleep
in three inbred mouse strains indicates that Homer 1a is specifically upregu-
lated in the brain. (A and B) Venn diagrams of a two-way ANOVA microarray
deprivation was similar between D2 and AK, which represent the
data analysis of total RNA extracted from brain (A) and liver (B) in AK, B6, and two extreme strains in their response to sleep deprivation, and was
D2 inbred mouse strains at ZT6 after a 6-h sleep deprivation. Results are significantly higher than in B6 mice (Fig. 1E). In accord with our
reported as a function of genetic background (genotype), experimental con- findings, sequence comparisons in public databases for the Homer1
dition (sleep deprivation), and their interaction. (C) Homer1a is overexpressed region indicated that D2 and AK share the same SNP haplotype,
after sleep deprivation (SD) in the cortex and caudate putamen (Left) and in which is different from that of B6, although no coding sequence
the hippocampus (Right). (D) Mean (⫾1 SEM) forebrain mRNA levels of differences between any of the three strains can be identified. This
Homer1a for AK, B6, and D2 mice after SDs of 1, 3, and 6 h (pink bars), with finding suggested that, even though Homer1a is the best candidate
their time-matched controls (gray bars). All SDs started at light onset (ZT0).
for the Dps1 QTL segregating with response to sleep loss between
Effects of recovery sleep on expression were assessed by allowing 2 h of
recovery after 6-h SD (ZT8). Homer1a expression was affected by SD, geno-
D2 and B6 mice, other genes may affect sleep need in other inbred
type, and time of day (P ⬍ 0.05; three-way ANOVA). (E) Fold change increase strains, or posttranscriptional or translational changes of Homer1a
of Homer1a expression level after 6 h of SD varies among strains. Ratios may differ between D2 and AK mice.
between SD and control mRNA levels at ZT 6 were calculated as described in
SI Materials and Methods, with their standard errors. The effect of SD on Time-of-Day Effects of Sleep Loss on Brain Transcriptional Changes. As
Homer1a expression is statistically different between B6 and the two other shown in Fig. 1D, the level of Homer1a expression under baseline
strains (⫹, P ⬍ 0.006; two-way ANOVA). conditions shows significant variation that might be either due to a
direct circadian effect or driven by the diurnal sleep–wake distri-
bution, as we suggest. To separate time-of-day and homeostatic
scripts either increased or decreased in all three inbred strains and effects on brain gene expression, we have simulated the time course
across experimental conditions. of sleep need (quantified as EEG delta power) according to
Sleep deprivation induced changes in only 42 brain probe sets in published assumptions and parameters (7) under entrained base-
B6, 92 in D2, and 188 in AK, among which 52 sets were affected by line conditions and after 6-h sleep deprivation at four time points
genotype. To our surprise, sleep loss induced almost three times around the 24-h day. If a gene is implicated in the processes
more transcriptional changes in liver compared with whole brain underlying sleep need, then changes in its expression can be
(Fig. 1 A). As in brain, among all transcripts with significant changes expected to parallel the predicted time course of sleep need (Fig.
in liver, ⬇50% changed as a function of genetic background. Thus, 2A). Under control conditions, sleep need is high at sleep onset,
for genes differentially expressed in each tissue of the three strains which coincides with light onset (ZT0), and during the latter part
in response to sleep deprivation, we have focused on those showing of the active period (ZT18), whereas sleep need is lowest between
a significant two-way ANOVA interaction (see SI Tables 1 and 2). ZT6 and ZT12 (i.e., dark onset) (Fig. 2 A). Under sleep deprivation
As reported in refs. 11, 15, and 16, sleep deprivation most signifi- conditions, these pronounced sleep–wake-dependent changes are
cantly up-regulated the expression of heat shock protein genes in expected to be strongly reduced. Thus, we sleep-deprived mice of
both brain and liver, strongly suggesting that this group of genes is the three strains for 6 h, starting at ZT0, ZT6, ZT12, or ZT18, and

Maret et al. PNAS 兩 December 11, 2007 兩 vol. 104 兩 no. 50 兩 20091
Fig. 2. Around-the-clock transcriptome
analysis of the brain after sleep deprivation
A ‘Sleep pressure’ B AK B6 D2 indicates that most changes are affected by
ZT 0 6 12 18 0 6 12 18 0 6 12 18 behavioral states. (A) Simulation of
sleep need in the around-the-clock sleep-
Homer1a control
deprivation experiment. Sleep-need dynam-
C Control SD
Cycling ics in the control (gray) were modeled
in SD mathematically according to Franken et al.
AK B6 D2 AK B6 D2 AK B6 D2 (7): lights on at ZT0, lights off at ZT12. The
ZT 6 ZT 6 increase in sleep-need during the four 6-h
ZT [h] sleep deprivations (SD) was determined by
ZT12
assuming a saturating exponential increase
Control
D ZT0
SD
ZT0 ZT18
during wakefulness (red dashed lines). The
red solid line connects values of sleep need
ZT 0
reached at the end of the SDs. (B and C) SD
cos coeff

ZT18 ZT6 ZT18 ZT6 ZT 6


affects the cycling pattern of gene expression
in the brain. Rhythmic transcripts were se-
lected, as outlined in SI Materials and Meth-
ZT12
ods, and their temporal expression patterns
ZT12
were aligned according to phase. (B) Enlarge-
sin coeff ment of heat map for Homer1a gene in con-
trol condition. For AK, B6, and D2, the four
time points (ZT0, 6, 12, and 18) are repre-
E Homer1a Bmal1 sented in triplicate. Green and red represent
ZT12
minimal and maximal expression levels, re-
spectively. (C Left) The 2,032 genes cycling in
min max
controls are depicted. The peak time of ex-
pression is indicated at left. (Center) The
ZT18 same genes are represented after SD. Note
mRNA Signal (100X)

that the rhythmic accumulation of most tran-


Egr3 Per2
scripts is severely blunted after SD. (Right)
The 391 probe sets for which cycling expres-
sion profiles are not affected by SD (with the
peak of expression indicated at right). (D)
Cycling genes shown in C are plotted accord-
ing to their amplitude and phase. Homer1a,
Arc Dbp outlined in green, is the most rhythmic tran-
script under control conditions. (E) Temporal
expression profiles of representative homeo-
static (Left) and circadian (Right) genes in
ZT 0 whole brain. Normalized expression signals
from microarray experiments are plotted as a
function of time for control (solid lines) and
ZT [h] SD (dashed lines) mice of each strain. Each
ZT 6 point is the mean ⫾ SEM of three pools of
three animals.

killed them together with their home-cage controls (n ⫽ 9 per strain as Homer1a, Arc, and Egr3 show a high amplitude variation (up to
per time per condition; i.e., total 216). A linear statistical model was 6-fold for Homer1a), closely following the diurnal sleep–wake
first used (see SI Materials and Methods) and identified 2,540 probe distribution, and no, or greatly reduced, variation after sleep
sets that were significantly changed at any time point (false-positive deprivation. Although the pattern of expression of circadian genes
rate; false discovery rate ⬍5%). These probe sets were then remains little affected by sleep deprivation, their relative levels can
assessed for time-of-day variation separately in the baseline and be significantly affected (unchanged for Bmal1, increased for Per2,
sleep-deprivation conditions. Under the baseline condition, ⬇8% and decreased for Dbp; Fig. 2E), as we also reported in refs. 17
(2,032; see SI Table 3) of probe sets detected in the brain showed and 18.
a significant time-of-day pattern of expression (Fig. 2 C and D), with Comparison of sleep-deprived and control mice at all four time
two major, opposite phases of peak expression (between ZT6 and points again revealed that ⬍2% (343 probe sets) of the expressed
ZT12 and between ZT18 and ZT0, respectively; Fig. 2D). Under the genes in the brain are up-regulated (⬎70%, or 249 probe sets) or
down-regulated (⬍30%, or 94 probe sets) by sleep loss (SI Table 5).
sleep-deprivation condition, only 391 (SI Table 4) of the 2,032 sets
Significant interaction between condition and time of day was
at baseline were still significantly affected by time of day, indicating
detected for 585 probe sets. Again, the most significantly overex-
that most others changed according to the sleep–wake distribution, pressed gene after sleep deprivation was Homer1a, followed by
rather than as a result of (or in addition to) a direct circadian effect. those belonging to stress–response and synaptic-plasticity gene
Among the 391 transcripts that maintained a significant cycling groups. The major functional gene groups that reduced their
pattern after sleep deprivation, a large majority reached maximum expression after sleep deprivation concern protein synthesis, mem-
levels of expression between ZT0 and ZT6 and between ZT12 and brane trafficking, and protein transport (SI Table 5) (12). Real-time
ZT18 (Fig. 2 C and D). Among all rhythmic transcripts under quantitative RT-PCR verification for 41 candidate genes (found
control conditions, Homer1a clearly showed the largest amplitude here and by others) at ZT6 confirmed our microarray findings at
of variation (Fig. 2 B, D, and E), beyond that of all canonical this reference time point (SI Table 6).
circadian genes. Time course of expression of six representative
genes under the two conditions is depicted in Fig. 2E. As predicted Cell-Specific Transcriptional Changes Due To Sleep Loss. Previous
by the simulation analysis (Fig. 2 A), activity-regulated genes such studies (12, 19, 20) investigated transcriptomes of different brain

20092 兩 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0710131104 Maret et al.


structures, rather than the whole brain. In contrast to most other
organs, the brain is a highly heterogeneous tissue, with specialized
A ~100kb
H1a Ania-3 H1b/c
regions and nuclei having very different functional roles. Whole- 1 2 3 4 5 6 7 8 9 10 Exons
brain transcriptome, therefore, suffers several potential limitations,
including dilution of low copy-number transcripts and missing those 5’ 6’
with opposing patterns of transcriptional changes among brain
regions. On the other hand transcriptome analysis of selected B 1 2 3 4 5
regions also has drawbacks because of our limited knowledge of BAC
exactly where functionally significant molecular changes can be

ZEO
Flag
PABP IRES eGFP
expected and because, even in well defined regions, cell types vary
greatly. To overcome these limitations, we chose to analyze changes Recombination
in mRNA profiles of those neurons that are selectively and specif- 179Kb 36Kb

ZEO
Flag
ically activated by sleep deprivation. To this end, we used a modified BAC PABP IRES eGFP

mRNA tagging technique originally established in Caenorhabditis Flip-out of Zeo


elegans and Drosophila (21–23).

Flag
We have shown here that Homer1a is the transcript most BAC PABP IRES eGFP

consistently increased by enforced wakefulness. The Homer1 gene


BXD-F2 pronucleus injection
is therefore a good marker for neuronal populations activated by 2 2

Flag
sleep loss but not restricted to a single structure. We thus generated Chrom. PABP IRES eGFP

BAC-based transgenic mice by replacing the first five exons of 1 1 3 3

Homer1 (corresponding to activity-induced Homer1a transcript;


Fig. 3A) by a FLAG-tagged poly(A) binding protein 1 (PABP) C 700
+ - + - + - Tg mice
D Tg mice Control
500
followed by internal ribosome entry site (IRES)-eGFP (Fig. 3B). 300
Sleep Deprived
142
- + - +

Because PABP binds poly(A) tails of mRNAs, affinity-purification Loading control


FLAG PABP eGFP
of FLAG-tagged PABP proteins from whole-brain lysates is ex-
pected to coprecipitate all mRNAs from neurons expressing 98
Flag
Homer1a. Seven transgenic lines were obtained, and analysis indi- 64

cated that at least three lines expressed the transgene at very similar 36
eGFP
amounts, at both mRNA and protein levels (Fig. 3 C and D). Also, 22

the induction of the transgene by 6-h sleep deprivation was very


Fig. 3. A transgenic mouse model to analyze neuron-specific gene expres-
similar to that of endogenous Homer1a, indicating that our BAC sion. (A) Schematic representation of the genomic structure of Homer1. The

NEUROSCIENCE
construction contains the regulatory elements for the correct putative transcriptional initiation site is depicted by a bent arrow at the
functional expression of Homer1a (Fig. 3D). beginning of exon 1; the translational stops for short activity-induced
The expression of this construct was also verified by in situ Homer1a (H1a) and Ania-3, as well as for long constitutively expressed
hybridization (Fig. 4A). Although eGFP inserted after the IRES did Homer1b/c (H1b/c) are indicated by black circles. Intron 5 is here divided into
not result in reliable signal under epifluorescence, riboprobes four segments (4.4 kb of Homer1a 3⬘ UTR, 5.7 kb up to Ania-3, 1.4 kb of
against Homer1a and eGFP clearly indicated that both are similarly Ania-3-specific sequence, and 18.8 kb to exon 6). The Homer1a-specific exon
coexpressed in the same brain regions (Fig. 4A). Double fluores- 5⬘ extends exon 5 by the intron 5 sequence. The Ania-3-specific exon 6⬘ sits
within intron 5 (adapted from ref. 25). (B) General strategy for generating
cent in situ hybridization (FISH) with riboprobes against Homer1a
Homer1a-PABP transgenic mice. The relative position of the Homer1a gene in
and eGFP revealed that, at least in two transgenic lines, eGFP was the fully sequenced 227,644-bp RP23–262I3 BAC clone (GenBank accession no.
almost exclusively expressed in Homer1a-expressing neurons (Fig. AC120347), and the different steps used to introduce a PCR-amplified con-
4B), which represented ⬇40% of the neurons in the cingulate cortex struction containing PABP cDNA, followed by IRES-eGFP and the zeocine (ZEO)
and 30% in the dorsal striatum (not counted in the hippocampus). selection cassette flanked by FRT sites (hashed boxes) by BAC recombination
All mRNA immunoprecipitations and microarray data presented in the EL250 bacterial stain (see SI Materials and Methods). (C) Transgenic mice
below are from a single transgenic line (line 36). Sleep recordings (Tg) were identified by RT-PCR with the primer pairs depicted in B for the
in Homer1-PABP transgenic mice indicated that they react to sleep presence of the FLAG (primers 1/1), PABP (primers 2/2), and eGFP (primers 3/3).
deprivation similar to their wild-type littermates (SI Fig. 5). The (D) Western blot verification of transgene expression in transgenic mice line
36 (Tg) indicated the presence of the FLAG and eGFP and the up-regulation of
specificity of this mRNA pull-down method was also verified by
the FLAG after sleep deprivation. The loading control is a nonspecific band
quantitative RT-PCR with probes specific for eGFP, FLAG, generated by the GFP antibody.
Homer1a, and Hcrt (a gene expressed only in the lateral hypothal-
amus, where Homer1a expression is very low). The results indicated
that eGFP, FLAG, and the endogenous Homer1a were enriched were identified (SI Table 7), among which the most significant
and induced by sleep deprivation, whereas only trace amounts of ones—Homer1a, Egr2 (NGFI-B), and Fosl2 (Fos-like antigen 2)—
Hcrt could be detected (SI Fig. 6). were similarly induced after sleep loss in the pull-down and the
Immunoprecipitated mRNAs were prepared from Homer1- whole-brain extracts, suggesting that gene expression changes in
PABP transgenic mice with (n ⫽ 6) or without (n ⫽ 6) a 6-h sleep the pull-down samples recapitulate the most significant changes at
deprivation at light onset for gene expression profiling. For com- the whole-brain level. In addition, several unique immediate-early
parison, RNA extractions were also made from supernatants (n ⫽ genes were specifically identified in pull-down samples that might
4) after immunoprecipitation and from transgenic whole brains be coinduced with Homer1a, namely prostaglandin–endoperoxide
(n ⫽ 8). To test for transcriptional changes after sleep deprivation synthase 2 (Ptgs2), junctophilin 3 (Jph3), and neuronal pentraxin 2
in Homer1-expressing cells, we proceeded in two steps: (i) we (Nptx2). Interestingly, both Jph3 and Nptx2 are activity-induced
identified probe sets enriched in the pull-down extracts and (ii) through either ryanodine receptor-mediated intracellular calcium
among those probe sets, we compared sleep deprivation with mobilization (Jph3) or activity-induced AMPA receptor synaptic
control condition in both pull-down (6 vs. 6-chip comparison) and clustering (Nptx2). Among the very few down-regulated transcripts
whole-brain (4 vs. 4-chip comparison) extracts. We found that 4,728 (SI Table 7), we have identified another activity-induced gene,
probe sets were significantly enriched at 5% false discovery rate 4-nitrophenylphosphatase domain and nonneuronal SNAP25-like
when pull-downs were compared with both supernatant and whole- protein homolog 1 (Nipsnap1), suggesting that plasticity genes can
brain extracts (SI Materials and Methods). Again, very few genes be up- or down-regulated by sleep deprivation.

Maret et al. PNAS 兩 December 11, 2007 兩 vol. 104 兩 no. 50 兩 20093
Nipsnap1, one of the proposed plasticity genes (24), is actually
A Homer1a eGFP B Antisense Sense
down-regulated after sleep deprivation in our mRNA tagging
experiment.

DAPI
Cg

Three different genes in mammals encode Homer proteins.


Antisense

Homer1 encodes constitutively expressed long-form proteins,


CPu whereas short-form Homer1a is activity-induced (25). Homer1
long-form proteins dimerize and interact with metabotropic gluta-

Homer1a
Pir mate receptors and increase calcium from intracellular stores.
Short-form proteins, which lack the dimerization domain, function
as natural activity-dependent dominant negative forms that regu-
late the scaffolding and signaling capabilities of the long forms and

GFP
reduce glutamate-induced intracellular calcium release (26, 27). We
have recorded sleep, and the response to a 6-h sleep deprivation, in
Sense

Homer1 (all forms) mutant mice but found a very similar pattern
compared with their wild-type littermates (data not shown). This
M e rged

finding could be expected due to the fact that, because Homer1a


functions as a dominant negative form of the long forms, consti-
5µm
tutive loss of all isoforms might not result in any specific sleep
phenotype.
Fig. 4. Colocalization of Homer 1a and FLAG-tagged PABP eGFP transcripts.
Conceptually, spontaneous or enforced wakefulness repre-
(A) In situ hybridization with Homer1a and eGFP antisense riboprobes indi-
cated that both are expressed in similar brain structures. Cg, cingulate cortex;
sents a stressor activating a series of stress–response mecha-
CPu, caudate putamen; Pir, piriform cortex. (B) Confocal images of FISH nisms of the organism, which, at the transcriptional level, could
analysis. FISH experiments were performed with both Homer1a (red) and eGFP be translated into the induction of genes such as heat shock
(green) riboprobes at the same time and revealed that almost all positive proteins in most tissues. However, unlike in other organs,
neurons were double-labeled, indicating the colocalization of the endoge- brain-specific stress–response pathways are primarily triggered
nous Homer1a and the transgene. Negative control conditions were obtained by glutamate. Glutamate is the major excitatory neurotrans-
with both sense riboprobes. mitter in the central nervous system and acts through either
ionotropic or metabotropic receptors (mGluRs). Long Hom-
er1 tetramers bind group I mGluRs and inositol 1,4,5-
Discussion triphosphate receptors, thus enabling efficient calcium release
The results presented here demonstrate that 6 h of sleep depriva- from intracellular stores, whereas monomeric Homer1a com-
tion, which importantly impacts sleep physiology and behavior, petitively disrupts synaptic glutamatergic signaling complexes
results in only minimal changes in brain transcriptional adaptation. to reduce glutamate-induced intracellular calcium release (26,
As reported for changes in delta power (7), we also showed here 27). Homer1 also activates ryanodine receptors and L-type
that sleep loss-induced transcriptional changes are largely affected calcium channels (28, 29). Interestingly, Jph3, which was
by genetic background. As opposed to other studies that did not identified by our mRNA-tagging strategy as being up-
take genetic background into account and did not contrast their regulated by sleep deprivation, has been shown to play a major
findings to peripheral tissue (12, 15, 20), our results indicate that role in ryanodine receptor-mediated, calcium-induced open-
only a few genes reliably change expression after sleep deprivation. ing of small-conductance, calcium-activated potassium (SK)
The surprising finding that sleep loss induced a larger number of channels (28, 30). SK channels are responsible for the gener-
changes in liver than in brain suggests either that sleep deprivation ation of slow afterhyperpolarizations in neurons of the nucleus
might have a specific impact on the liver or that the brain might be reticularis thalami and thus contribute to the EEG slow waves
protected against major transcriptional changes. characteristic of NREM sleep (31).
Another important aspect of the present study is the interaction According to Tononi and Cirelli (3), plastic processes occurring
between the homeostatic and circadian processes. Although we did during wakefulness result in increased synaptic strength, whereas
not sleep-deprive the animals under constant conditions, and the role of sleep is to downscale synaptic strength to a basal level.
therefore the direct and indirect effects of light, for instance, on Homer1a transcription is rapidly up-regulated in neurons in re-
gene expression could not be accounted for, we have shown that a sponse to synaptic activity induced by long-term potentiation,
large majority (⬎80%) of changes in gene expression were driven seizure, inflammation, stimulant drugs, or even selectively in the
by the prior sleep–wake history. hippocampus of rodents by exploratory behavior (32, 33). In this
We also adopted, and further developed, a reliable mRNA context, Homer1a, by buffering intracellular calcium and disassem-
tagging technique to investigate gene expression changes in neu- bling synaptic glutamatergic signaling complexes, could play a
rons. This technique can be used to evaluate different neuronal pivotal role in synaptic downscaling. Because Homer1a and the
subpopulations without the burden of sampling several structures other genes identified here (Egr2, Fosl2, Ptgs2, Jph3, and Nptx2) are
or using labor-intensive laser microdissection to isolate neurons. all induced by stressful conditions such as seizure, stroke and
The results of this technique confirmed that sleep loss-induced hypoxia, and inflammation, an alternative, complementary view
transcriptional changes occur for very few genes, among which could be that they play a primary brain-protective or recovery role.
Homer1a remains the most specific. This view is also of relevance for the etiology of neuropsychiatric
In addition to Homer1a, we identified overexpression of other disorders because it is increasingly recognized that stress is impli-
genes involved in synaptic plasticity, but only Egr2 and Homer1a cated in many of such disorders (34). Both environmental and
were found to consistently change across experiments. Others pharmacological stressors up-regulate Homer1a mRNA in key
reported overexpression for a number of plasticity-related genes, structures involved in higher brain functions (35): the same struc-
and these observations are commonly used in support of a func- tures in which Homer1a is up-regulated after sleep deprivation. It
tional role for sleep in plasticity. Because the expression of most has also been shown that overexpression of Homer1a after inflam-
plasticity-related genes was not reliably changed, our findings do not mation, seizure, and psychostimulant or antipsychotic drug use
support such a general conclusion and instead suggest that the plays a major role in neuroprotection (28, 35). It is tempting to
molecular mechanisms might not be identical for most plasticity- relate the dramatic improvement in depression in humans after
regulated genes. In this context it is important to note that sleep-deprivation (36) to the sleep deprivation-induced up-

20094 兩 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0710131104 Maret et al.


regulation of Homer1a. In conclusion, converging evidence strongly lated by using the bicinchoninic acid assay (Pierce) with BSA as a
implicates Homer1a as a brain-coping marker against stressors, and standard. Eighty micrograms of each fraction were analyzed by
our findings suggest that Homer1a might represent the molecular SDS/PAGE, followed by Western blotting using antibodies as
link between sleep, cognition, and neuropsychiatric disorders. follows: mouse anti-tubulin 1/1,000 (Santa Cruz), goat anti-
Homer1a 1/200 (Santa Cruz), mouse anti-Flag M2 1/300 (affinity-
Materials and Methods purified; Sigma), and rabbit anti-GFP 1/2,500 (AbCam). Secondary
Animal Handling. All experiments were performed in accor- antibodies were all coupled with HRP, except for the anti-goat
dance with the protocols approved by the Ethical Committee antibody, which was IRDye800-conjugated for Lycor analysis.
of the State of Vaud Veterinary Office, Switzerland. Sleep- In situ hybridizations with coronal cryosections of 12 ␮m were
deprivation and sleep-recording procedures are described in SI performed according to Allen Brain Atlas protocols (enzymatic
Materials and Methods. BCIP/NBT revelation) (37). All reagents and solutions were pur-
chased and prepared based on Eurexpress II in situ hybridization
cRNA Preparation, cDNA Microarray Hybridization, and Real-Time consortium instructions. GFP and Homer1a riboprobes were syn-
RT-PCR. For the first experiment, we isolated total RNA from whole thesized by in vitro transcription on a linearized pGEM-Easy vector
brain and liver by using the RNAXEL kit (Eurobio), treated the (Promega) containing the corresponding sequences. The cDNA
RNA with DNase, and cleaned it using RNeasy columns (Qiagen). insert of this plasmid was generated by RT-PCR from mouse brain
Equal quantities of total RNA from three individual mice of each RNA, using the following primers: Homer1a forward, 5⬘-
strain were pooled in triplicate (nine mice of each strain in each GCTGTCAGAAGCTTAGGATGTG-3⬘; Homer1a reverse,
condition). A hybridization mixture containing 15 ␮g of biotin- 5⬘-AAAGTGCAGAAAGTCCAGCAGC-3⬘; GFP forward, 5⬘-
ylated cRNA was hybridized to GeneChip Mouse Expression Set GAGCTGGACGGCGACGTAAACG-3⬘; and GFP reverse, 5⬘-
430. Chips were washed, scanned, and analyzed with Affymetrix AGGACCATGTGATCGCGCTTCTC-3⬘.
GeneSpring software. FISH was performed as described in ref. 38, using anti-DIG-
For the around-the-clock microarray experiment, RNA from POD 1/600 (Roche), anti-FLU-AP 1/100 (Roche), and SA-Alexa
whole brain was isolated and purified with the RNeasy Lipid Tissue 488 (Molecular Probes) and counterstained with DAPI (Sigma).
Midi kit (Quiagen) and DNase-treated. All RNA quantities were
assessed with a NanoDrop ND-1000 spectrophotometer, and the Transgenic and mRNA Tagging. Transgenic mice were generated as
quality of RNA was controlled on Agilent 2100 bioanalyzer chips. described in Fig. 3. See SI Materials and Methods for details.
Equal amounts of total RNA were pooled from three mice within
each of the 24 experimental groups (three strains, two conditions, We thank K. Harshman, A. Paillusson, and M. Bueno for assistance in
4 ZT ⫽ 24; in triplicate: 24 䡠 3 ⫽ 72 chips). Three micrograms of microarray and real-time RT-PCR analyses at the Lausanne DNA Array
each of these pools were used to perform the chip array experiment, Facility; P. Descombes, M. Docquier, D. Chollet, and C. Delucinge for
according to the Affymetrix Gene Expression procedure. Twelve assistance in microarray and real-time RT-PCR analyses at the Geneva

NEUROSCIENCE
micrograms of biotinylated cRNA from each sample were frag- Genomics Platform, National Center for Competence in Research Frontiers
in Genetics; S. Excoffier for help in the transgenic construction; P. Seeburg,
mented and hybridized to GeneChip Mouse 430 2.0 arrays, accord-
M. Schwarz (Max Planck Institute, Heidelberg, Germany), and P. Worley
ing to standard procedures. Microarray analyses and qPCR verifi- (Johns Hopkins School of Medicine, Baltimore, MD) for providing Homer1
cations were performed as reported in SI Materials and Methods. mutant mice; and A. Vassali for constructive discussions. This work was
supported by the Swiss National Science Foundation and the State of Vaud
Immunoblot and in Situ Hybridization. Total protein extract was (M.T.) and in part by National Institutes of Mental Health Grant MH67752
prepared with RIPA lysis buffer. Protein concentration was calcu- (to P.F.).

1. Dijk DJ, Franken P (2005) in Principles and Practice of Sleep Medicine, eds 21. Roy PJ, Stuart JM, Lund J, Kim SK (2002) Nature 418:975–979.
Kryeger TH, Roth TH, Dement W (Saunders, Philadelphia), pp 418–434. 22. Kunitomo H, Uesugi H, Kohara Y, Iino Y (2005) Genome Biol 6:R17.
2. Krueger JM, Obal F (1993) J Sleep Res 2:63–69. 23. Yang Z, Edenberg HJ, Davis RL (2005) Nucleic Acids Res 33:e148.
3. Tononi G, Cirelli C (2006) Sleep Med Rev 10:49–62. 24. Satoh K, Takeuchi M, Oda Y, Deguchi-Tawarada M, Sakamoto Y, Matsubara
4. Huber R, Ghilardi MF, Massimini M, Ferrarelli F, Riedner BA, Peterson MJ, K, Nagasu T, Takai Y (2002) Genes Cells 7:187–197.
Tononi G (2006) Nat Neurosci 9:1169–1176. 25. Bottai D, Guzowski JF, Schwarz MK, Kang SH, Xiao B, Lanahan A, Worley
5. Marshall L, Helgadottir H, Molle M, Born J (2006) Nature 444:610–613. PF, Seeburg PH (2002) J Neurosci 22:167–175.
6. Molle M, Marshall L, Gais S, Born J (2004) Proc Natl Acad Sci USA 26. Worley PF, Zeng W, Huang G, Kim JY, Shin DM, Kim MS, Yuan JP, Kiselyov
101:13963–13968. K, Muallem S (2007) Cell Calcium 42:363–371.
7. Franken P, Chollet D, Tafti M (2001) J Neurosci 21:2610–2621. 27. Xiao B, Tu JC, Worley PF (2000) Curr Opin Neurobiol 10:370–374.
8. Van Dongen HP, Baynard MD, Maislin G, Dinges DF (2004) Sleep 27:423–433. 28. Kakizawa S, Kishimoto Y, Hashimoto K, Miyazaki T, Furutani K, Shimizu H,
9. Tucker AM, Dinges DF, Van Dongen HP (2007) J Sleep Res 16:170–180. Fukaya M, Nishi M, Sakagami H, Ikeda A, et al. (2007) EMBO J 26:1924–1933.
10. Shifman S, Bell JT, Copley RR, Taylor MS, Williams RW, Mott R, Flint J 29. Yamamoto K, Sakagami Y, Sugiura S, Inokuchi K, Shimohama S, Kato N
(2006) PLoS Biol 4:e395. (2005) Eur J Neurosci 22:1338–1348.
11. Cirelli C, Gutierrez CM, Tononi G (2004) Neuron 41:35–43. 30. Moriguchi S, Nishi M, Komazaki S, Sakagami H, Miyazaki T, Masumiya H,
12. Mackiewicz M, Shockley KR, Romer MA, Galante RJ, Zimmerman JE, Saito SY, Watanabe M, Kondo H, Yawo H, et al. (2006) Proc Natl Acad Sci USA
Naidoo N, Baldwin DA, Jensen ST, Churchill GA, Pack A (2007) Physiol 103:10811–10816.
Genomics 31:441–457. 31. Blethyn KL, Hughes SW, Toth TI, Cope DW, Crunelli V (2006) J Neurosci
13. Nelson SE, Duricka DL, Campbell K, Churchill L, Krueger JM (2004) Neurosci 26:2474–2486.
Lett 367:105–108. 32. Tappe A, Klugmann M, Luo C, Hirlinger D, Agarwal N, Benrath J, Ehren-
14. Huber R, Tononi G, Cirelli C (2007) Sleep 30:129–139. gruber MU, During MJ, Kuner R (2006) Nat Med 12:677–681.
15. Terao A, Steininger TL, Hyder K, Apte-Deshpande A, Ding J, Rishipathak D, 33. Kato A, Ozawa F, Saitoh Y, Fukazawa Y, Sugiyama H, Inokuchi K (1998) J Biol
Davis RW, Heller HC, Kilduff TS (2003) Neuroscience 116:187–200. Chem 273:23969–23975.
16. Wisor JP, Morairty SR, Huynh NT, Steininger TL, Kilduff TS (2006) Neuroscience 34. Southwick SM, Vythilingam M, Charney DS (2005) Annu Rev Clin Psychol
141:371–378. 1:255–291.
17. Franken P, Thomason R, Heller HC, O’Hara BF (2007) BMC Neurosci 8:87. 35. Szumlinski KK, Kalivas PW, Worley PF (2006) Curr Opin Neurobiol 16:251–
18. Wisor JP, O’Hara BF, Terao A, Selby CP, Kilduff TS, Sancar A, Edgar DM, 257.
Franken P (2002) BMC Neurosci 3:20. 36. Giedke H, Schwarzler F (2002) Sleep Med Rev 6:361–377.
19. Cirelli C, Faraguna U, Tononi G (2006) J Neurochem 98:1632–1645. 37. Lein ES, Hawrylycz MJ, Ao N, Ayres M, Bensinger A, Bernard A, Boe AF,
20. Terao A, Greco MA, Davis RW, Heller HC, Kilduff TS (2003) Neuroscience Boguski MS, Brockway KS, Byrnes EJ, et al. (2007) Nature 445:168–176.
120:1115–1124. 38. Ishii T, Omura M, Mombaerts P (2004) J Neurocytol 33:657–669.

Maret et al. PNAS 兩 December 11, 2007 兩 vol. 104 兩 no. 50 兩 20095

You might also like