Skip to main content

Crucial role of TFAP2B in the nervous system for regulating NREM sleep

Abstract

The AP-2 transcription factors are crucial for regulating sleep in both vertebrate and invertebrate animals. In mice, loss of function of the transcription factor AP-2β (TFAP2B) reduces non-rapid eye movement (NREM) sleep. When and where TFAP2B functions, however, is unclear. Here, we used the Cre-loxP system to generate mice in which Tfap2b was specifically deleted in the nervous system during development and mice in which neuronal Tfap2b was specifically deleted postnatally. Both types of mice exhibited reduced NREM sleep, but the nervous system-specific deletion of Tfap2b resulted in more severe sleep phenotypes accompanied by defective light entrainment of the circadian clock and stereotypic jumping behavior. These findings indicate that TFAP2B in postnatal neurons functions at least partly in sleep regulation and imply that TFAP2B also functions either at earlier stages or in additional cell types within the nervous system.

Introduction

Sleep is an evolutionarily conserved state that is crucial for maintaining health in humans [1]. The exact functions and mechanisms of sleep, however, remain elusive. In humans, there are reports of natural short sleepers whose reduced sleep does not lead to daytime fatigue [2]. In cases with an inherited natural short sleep trait, identification and analyses of the causal gene are expected to provide clues to the molecular mechanisms of sleep.

In the present study, we focused on the mouse transcription factor AP-2β (TFAP2B). In humans, partial mutations in TFAP2B cause Char syndrome, which is characterized by morphologic abnormalities of the face, limbs, and heart [3, 4]. Two human kindreds with Char syndrome self-reported abnormalities in sleep, namely parasomnia (sleep-walking) and short sleep, respectively [5]. In nematodes, the orthologous transcription factor APTF-1 is critical for specification of the sleep-promoting RIS neuron, and aptf-1 mutants exhibit severely reduced sleep [6,7,8]. In fruit flies, knockdown of neuronal TFAP2 results in reduced nighttime sleep [9]. In mammals, homozygous Tfap2b-knock out (KO) mice die shortly after birth [10, 11] whereas heterozygous knockout or partial mutations in mouse Tfap2b lead to reduced or fragmented non-rapid eye movement (NREM) sleep [12, 13]. This phenotype is partly explained by the roles of TFAP2B in GABAergic neurons [14], but considering the rather mild effects on NREM sleep of the conditional knockdown (cKD) of Tfap2b in GABAergic neurons, the entire picture of the mechanisms by which mouse TFAP2B functions to regulate NREM sleep are unclear. During embryogenesis, Tfap2b is expressed in neural crest cells, which generate various cell types, and in the midbrain and hindbrain, where it contributes to the specification of various neurons, including noradrenergic neurons in the locus coeruleus (LC) [15, 16]. Postnatally, Tfpa2b is expressed in the paraventricular hypothalamic nucleus (PVH), superior colliculus (SC), parabrachial nucleus (Pb), cerebellum, LC, and nucleus of the solitary tract (NTS), some of which are regions involved in sleep regulation [12]. It is unknown whether TFAP2B acts during development or postnatally in sleep regulation. Here, using the Cre-loxP system, we tested the effect of homozygous deletion of Tfap2b in either the nervous system during development or in neurons postnatally.

Results

Nervous system-specific Tfap2b deletion induces decreased NREM sleep and increased wakefulness

To elucidate whether TFAP2B functions in the nervous system or other tissues in sleep regulation, we used Nes-Cre mice to delete Tfap2b in a nervous system-specific manner. In Nes-Cre mice, Cre recombinase is expressed in neuronal and glial cell precursors of the central and peripheral nervous systems [17]. Nes-Cre mice were crossed with Tfap2bflox/flox mice, in which the Tfap2b exon 3 is flanked by loxP. We confirmed that Tfap2b exon 3 is deleted in the genomic DNA of the brain in a Nes-Cre-dependent manner (Additional file 1: Fig. S1A). Homozygous Tfap2b KO mice die shortly after birth [10, 11], whereas the nervous system-specific homozygous Tfap2b-conditionally knocked down (cKD) (Nes-Cre; Tfap2bflox/flox) mice lived to adulthood. However, when we crossed Nes-Cre; Tfap2bflox/+ mice with Tfap2bflox/flox mice to obtain the cKD mice, only 19 out of 176 pups were cKD, which is lower than the expected ratio of 1/4, implying that some of the cKD mice died either prenatally or immediately after birth. The Nes-Cre; Tfap2bflox/flox male mice had a lower body weight than control male mice (Tfap2bflox/+, Tfap2bflox/flox, or Nes-Cre; Tfap2bflox/+ mice; Additional file 1: Fig. S1B). We compared the sleep architecture between Nes-Cre; Tfap2bflox/flox and control male mice (Fig. 1). Nes-Cre; Tfap2bflox/flox mice exhibited increased wakefulness and decreased NREM sleep compared with control mice (Fig. 1B). Decreased NREM sleep could be attributed to a decrease in the episode number (Fig. 1C), whereas the episode duration seemed unaffected (Fig. 1D). In contrast, increased wakefulness could be attributed to a longer episode duration (Fig. 1D), whereas the episode number was rather decreased (Fig. 1C), suggesting that episodes of wakefulness are highly consolidated. When light and dark phases were separately analyzed in Nes-Cre; Tfap2bflox/flox mice, NREM sleep was largely decreased in the light phase with the amount being comparable between the light and dark phases (Fig. 1B). A similar trend was observed for wakefulness, implying that Nes-Cre; Tfap2bflox/flox mice lack typical daily fluctuations in the amount of sleep/wake time (Fig. 1B). This trend was even more obvious in rapid eye movement (REM) sleep; while the total amount of REM sleep was unaffected, the amount of REM sleep in the light phase was decreased, whereas that in the dark phase was increased (Fig. 1B). Analyses of the bi-hourly changes in the sleep–wake cycle (Fig. 1A) also revealed a similar trend; the amount of sleep or wakefulness seemed to lack daily oscillations, and when compared with control mice, the amounts of NREM and REM sleep were decreased at some time points in the light phase and increased at some time points in the dark phase, and vice versa for wakefulness (Fig. 1A). To further address whether Nes-Cre; Tfap2bflox/flox mice exhibit circadian rhythm deficits, we analyzed locomotor activity detected with an infrared sensor. Each individual control mouse exhibited high locomotor activity in the dark phase (Additional file 1: Fig. S2A, B). By contrast, while individual Nes-Cre; Tfap2bflox/flox mice seemed to have a daily locomotor activity rhythm, the start of the active phase varied among individuals and seemed random (Additional file 1: Fig. S2C), suggesting that the circadian clock is not entrained by external light–dark cycles in these mice. When we defined the active phase and inactive phase based on locomotor activity (see Materials and methods for details), Nes-Cre; Tfap2bflox/flox mice appeared to have reduced NREM sleep during both the active and inactive periods, whereas REM sleep seemed unaffected (Additional file 1: Fig. S3), providing further support that the sleep deficits in these mice result from a combination of reduced NREM sleep, increased wakefulness, and abnormal light entrainment of the circadian clock. In addition, when the behavior was monitored with a video camera, Nes-Cre; Tfap2bflox/flox mice frequently displayed stereotypic behaviors such as repeated jumping or climbing (Additional file 2: Video 1). In the electrogram (EEG) power spectrum, Nes-Cre; Tfap2bflox/flox mice exhibited an increase in the EEG power at approximately 5.5–17 Hz during wakefulness, whereas no obvious defects were detected during sleep (Additional file 1: Fig. S4).

Fig. 1
figure 1

Comparison of the sleep architecture in nervous system-specific Tfap2b cKD and control mice. A Bi-hourly amount of wakefulness, NREM sleep, and REM sleep across 24 h in male mice. B Total amount of wakefulness, NREM sleep, and REM sleep during 24 h, light phase, and dark phase in male mice. N = 9 Tfap2bflox/flox mice (grey), N = 9 Nes-Cre; Tfap2bflox/+ mice (blue), N = 6 Nes-Cre; Tfap2bflox/flox mice (pink). C, D Episode numbers (C) and durations (D) of wakefulness, NREM sleep, and REM sleep during 24 h, light phase, and dark phase. N = 8 Tfap2bflox/flox mice (grey), N = 9 Nes-Cre; Tfap2bflox/+ mice (blue), N = 6 Nes-Cre; Tfap2bflox/flox mice (pink). # and indicate a significant main effect of genotype and a significant interaction between genotype and time, respectively, in 2-way repeated-measures ANOVA (#, p < 0.05, ###, ‡‡‡p < 0.001, ####, ‡‡‡‡p < 0.0001). indicates significance in the 1-way ANOVA (p < 0.05, ††p < 0.01, ††††p < 0.0001). * indicates significance in the post-hoc Bonferroni multiple comparison test (*p < 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001). Data are presented as mean ± SEM

Postnatal neuron-specific Tfap2b deletion induces decreased NREM sleep and increased wakefulness

To test whether TFAP2B in neurons acts postnatally in sleep regulation, we crossed Tfap2bflox/flox mice with Syn1CreERT2 mice, which express CreERT2 specifically in neurons [18]. To delete Tfap2b in neurons postnatally, we administered tamoxifen to Syn1CreERT2; Tfap2bflox/flox mice at postnatal day (P) 14, 17, and 21. We confirmed that Tfap2b exon 3 is deleted in the genomic DNA of the brain in a Syn1CreERT2-dependent manner (Additional file 1: Fig. S5A). These Syn1CreERT2; Tfap2bflox/flox mice survived to adulthood and exhibited no significant difference in body weight (Additional file 1: Fig. S5B). Repetitive jumping or climbing behaviors were not observed (Additional file 3: Video 2). We compared the sleep architecture between Syn1CreERT2; Tfap2bflox/flox and Tfap2bflox/flox male mice administered with tamoxifen (at P14, 17, and 21) (Fig. 2). Syn1CreERT2; Tfap2bflox/flox mice showed increased wakefulness and decreased NREM sleep (Fig. 2A, B), similar to Nes-Cre; Tfap2bflox/flox mice, although the phenotype seemed milder. Perhaps due to the mild phenotype, episode numbers and durations were not significantly affected (Fig. 2C, D). When light and dark phases were analyzed separately, the amount of wakefulness and NREM sleep seemed to be affected throughout both phases (Fig. 2B). Unlike Nes-Cre; Tfap2bflox/flox mice, the circadian rhythm seemed normal, with a high amount of wakefulness during the dark phase and a low amount during the light phase, and vice versa for NREM and REM sleep (Fig. 2A, B). EEG power spectra during wakefulness, NREM sleep, or REM sleep appeared unaffected (Additional file 1: Fig. S6). We also tested whether sleep homeostasis is normal in these mice. Mice were subjected to 5 h of sleep deprivation (SD) at the beginning of the light phase. Both Syn1CreERT2; Tfap2bflox/flox mice and Tfap2bflox/flox mice showed an increase in NREM sleep during the following 7 h of the light phase after SD (Additional file 1: Fig. S7A, B). Sleep architecture and EEG power spectra were also similar (Additional file 1: Fig. S7C-E), suggesting that the Syn1CreERT2; Tfap2bflox/flox mice exhibit a normal response to SD. Importantly, a previous study showed that the Syn1CreERT2 allele does not affect the amount of wakefulness, NREM sleep, and REM sleep [18]. However, we cannot rule out the possibility that it might have had some effect on the sleep architecture or the EEG power spectra, although these factors seemed not to differ between Syn1CreERT2; Tfap2bflox/flox and Tfap2bflox/flox mice.

Fig. 2
figure 2

Comparison of the sleep architecture in postnatal neuron-specific Tfap2b cKD and control mice. A Bi-hourly amount of wakefulness, NREM sleep, and REM sleep across 24 h in tamoxifen-injected (at P14, 17, and 21) male mice. B Total amount of wakefulness, NREM sleep, and REM sleep during 24 h, light phase, and dark phase in tamoxifen-injected (at P14, 17, and 21) male mice. C, D Episode numbers (C) and durations (D) of wakefulness, NREM sleep, and REM sleep during 24 h, light phase, and dark phase in tamoxifen-injected (at P14, 17, and 21) male mice. N = 6 Tfap2bflox/flox mice (grey), N = 8 Syn1CreERT2; Tfap2bflox/flox mice (orange). # and indicate a significant main effect of genotype and significant interaction between genotype and time, respectively, in 2-way repeated-measures ANOVA (##,‡‡p < 0.01, ###p < 0.001). ^ indicates significance in the Welch’s test (^^ p < 0.01). Data are presented as mean ± SEM

Differential patterns of recombination in Nes-Cre and Syn1 CreERT2 mice.

To reveal the Cre-mediated recombination patterns in Nes-Cre and Syn1-CreERT2 mice, we crossed these mice with Rosa26LSL-L10-GFP Cre-reporter mice. In the mature brain, TFAP2B is expressed in the PVH, SC, Pb, cerebellum, LC, and NTS [12]. Within these areas, Nes-Cre; Rosa26LSL-L10-GFP/+ mice had dense GFP signals in the SC, Pb, cerebellum, and NTS, whereas GFP signals were sparse in the PVH and almost undetectable in the LC (Fig. 3A–E’). Syn1CreERT2; Rosa26LSL-L10-GFP/+ mice injected with tamoxifen (at P14, 17, and 21) had dense GFP signals in the PVH, SC, Pb, LC, and NTS, whereas GFP signals were sparse in the cerebellum (Fig. 3F–J’). GFP signals in the LC mostly overlapped with tyrosine hydroxylase, suggesting that the recombination occurred in noradrenergic neurons (Additional file 1: Fig. S8).

Fig. 3
figure 3

Comparison of the Cre-mediated recombination patterns in Nes-Cre and Syn1CreERT2 mice. (AJ’) Images of areas where GFP signals were detected in Nes-Cre; Rosa26LSL−L10−GFP/+ mice (A–E’) and tamoxifen-injected (at P14, 17, and 21) Syn1CreERT2; Rosa26LSL−L10−GFP/+ mice (F–J’). PVH, paraventricular hypothalamic nucleus (A’, F’); SC, superior colliculus (B’, G’); Pb, parabrachial nucleus (C’, H’); cerebellum (D’, I’), LC, locus coeruleus (D’’, I’’); NTS, solitary tract nucleus (E’, J’), scp, superior cerebellar peduncle. Scale bars in (AJ), 1 mm; scale bars in (A’–J’), 200 µm. Representative images from a replicate of N = 2 Nes-Cre; Rosa26LSL−L10−GFP/+ mice, N = 4 Syn1CreERT2; Rosa26LSL−L10−GFP/+ mice are shown

Discussion

Heterozygous deletion of Tfap2b leads to reduced NREM sleep [12]. The present study showed that homozygous deletion of Tfap2b in the nervous system during development leads to a large reduction in NREM sleep. Moreover, homozygous deletion of Tfap2b in neurons postnatally reduced NREM sleep, although to a milder extent. These findings indicate that TFAP2B in postnatal neurons at least partly functions in sleep regulation and imply that TFAP2B also functions either at earlier stages or in additional cell types.

Consistent with previous studies [19], in Nes-Cre mice, Cre recombination was not detected in various brain regions. Within areas that postnatally express Tfap2b, Cre recombination was detected in the SC in both Nes-Cre and Syn1CreERT2 mice. GABAergic neurons in the SC regulate sleep by inhibiting dopaminergic neurons in the ventral tegmental area [20], and loss of function in these neurons might be partly responsible for the reduced NREM sleep in both Syn1CreERT2; Tfap2bflox/flox and Nes-Cre; Tfap2bflox/flox mice. This is consistent with a recent study demonstrating that deletion of Tfap2b specifically in GABAergic neurons leads to reduced NREM sleep [14]. The effect of deleting Tfap2b in GABAergic neurons is limited, however, suggesting that TFAP2B in other neuronal types also functions in sleep regulation. Among areas that postnatally express Tfap2b, Cre recombination was also detected in the NTS. Glutamatergic neurons in the NTS promote NREM sleep [21], and loss of function in these neurons might also contribute to reduce NREM sleep. By contrast, among areas that postnatally express Tfap2b, Cre recombination was almost undetectable in the LC in Nes-Cre mice. TFAP2B is required for the synthesis of noradrenaline in the LC [22], which in turn is crucial for maintaining wakefulness [23]. Thus, the milder reduction of NREM sleep in Syn1CreERT2; Tfap2bflox/flox mice compared with Nes-Cre; Tfap2bflox/flox mice might be partly due to the additional functional loss of LC noradrenergic neurons in the Syn1CreERT2; Tfap2bflox/flox mice, which might have counteracted the NREM sleep-reducing effect of deleting Tfap2b. Another possibility might be that TFAP2B also functions in glial cells or neural progenitor cells to regulate sleep and that deletion of Tfap2b in these cells are responsible for the more severe sleep phenotypes of Nes-Cre; Tfap2bflox/flox mice compared with Syn1CreERT2; Tfap2bflox/flox mice. Indeed, homozygous Tfap2b KO affects expression of genes involved in glial cell differentiation [14]. These possibilities should be tested in the future using mice that express Cre specifically in either neurons or glial cells.

Nes-Cre; Tfap2bflox/flox mice had prolonged wake episodes, during which they exhibited stereotypic behaviors, including repeated jumping and climbing. Similar stereotypic behaviors are observed in mice with knockout of dopamine transporter 1 (Dat1) and those with knockout of SH3 and multiple ankyrin repeat domains 2 (Shank2) [24, 25] and are thought to share properties with human stereotypic behaviors observed in various psychiatric or developmental disorders [26]. This phenotype is likely due to loss of Tfap2b either during the fetal period or in glial cells as such abnormal behavior was not observed in Syn1CreERT2; Tfap2bflox/flox mice.

In addition, Nes-Cre; Tfap2bflox/flox mice exhibited a deficit in light entrainment of the circadian rhythm. Intrinsically photosensitive retinal ganglion cells (ipRGCs) function as photosensors for this light entrainment [27]. Within the retina, TFAP2B is crucial for the development of horizontal and amacrine cells [28, 29], and considering the deficits in light entrainment, TFAP2B might also have a role in the development of ipRGCs.

Whereas Syn1CreERT2; Tfap2bflox/flox mice showed reduced basal NREM sleep, they responded normally to SD with increased amounts of NREM sleep in the subsequent hours. Thus, TFAP2B might not be required in postnatal neurons for the homeostatic regulation of sleep, but TFAP2B in other cell types may still play a role. We could not conduct a similar SD experiment with Nes-Cre; Tfap2bflox/flox mice because the beginning of the inactive phase varied among individual mice.

This study showed that postnatal neurons require TFAP2B to regulate sleep. To further narrow down the brain regions and neuronal subtypes in which Tfap2b is involved in regulating sleep, additional studies using region-specific Cre-driver lines or local microinjection of adeno-associated viral vectors carrying Cre are required. Once the responsible populations of neurons are identified, further studies using single-cell RNA sequencing may identify the target genes of TFAP2B that are required for regulating NREM sleep and contribute to a deeper understanding of the molecular mechanisms of sleep.

Materials and methods

Mice

All animal experiments were approved by the institutional animal care and use committee of the University of Tsukuba. All animals were maintained according to the institutional guidelines of the animal facilities of the Laboratory of Animal Resource Center, University of Tsukuba. Mice were maintained and bred within the International Institute for Integrative Sleep Medicine under a 12-h light/dark cycle and with free access to food and water. Male Nes-Cre; Tfap2bflox/flox, Syn1CreERT2; Tfap2bflox/flox, and corresponding control mice aged 10–13 weeks were used for sleep analyses. To obtain Tfap2b cKD mice, Tfap2btm1a(EUCOMM)Wtsi/+ mice were crossed with mice carrying the CAG-flp allele [30] to remove a genomic DNA fragment containing IRES:lacZ and a neomycin-resistant cassette flanked by FRT. Homozygous floxed Tfap2b exon 3 (Tfap2bflox/flox) mice were mated to mice carrying the Nes-Cre transgene [17] or the Syn1CreERT2 allele [18] to induce Cre-mediated recombination in a nervous system-specific manner or a neuron-specific manner, respectively.

 To confirm the recombination sites in the Nes-Cre and Syn1CreERT2 mice, Nes-Cre and Syn1CreERT2 mice were mated with the GFP reporter strain B6.129S4-Gt (ROSA)26Sortm1(CAG-EGFP/Rpl10a-birA) Wtp/J (Rosa26LSL-L10-GFP, Jackson Laboratory ; #022367) [31].

To induce CreERT2-mediated recombination, 500 μg of tamoxifen (Toronto Research Chemicals Inc.; T006000) dissolved in corn oil was administered intraperitoneally at P14, P17, and P21.

Genotyping

Genotyping of mice carrying Tfap2bflox, Nes-Cre, Syn1CreERT2, or Rosa26LSL-L10-GFP was performed using the following primers. Tfap2bflox: 5′-AAGGCGCATAACGATACCAC-3′, 5′-CCGCCTACTGCGACTATAGAGA-3′, 5′-GACATCCTACAATGCACAGCT-3′, and 5′- TTGCTGTGAGCTAAGAGCTTC-3′ (Tfap2bflox: 218 bp, Tfap2b+: 381 bp), Nes-Cre: 5′-GACGATGCAACGAGTGATGA-3′ and 5′-AGCATTGCTGTCACTTGGTC-3′ (Nes-Cre: 300 bp), Syn1CreERT2: 5′-TGCCTCCACCTTGTCTCTCT-3′, 5′-AACAAAGGCATGGAGCATCT-3′, and 5′-GATCTGGAGGTGACCAGGAA-3′ (Syn1CreERT2: 443 bp, Syn1+: 387 bp), Rosa26LSL-L10-GFP: 5′-AAGGGAGCTGCAGTGGAGTA-3′, 5′-CCGAAAATCTGTGGGAAGTC-3′, 5′-AAGATCCGCCACAACATCG-3′, and 5′-TTCTCGTTGGGGTCTTTGCT-3′ (Rosa26LSL-L10-GFP: 146 bp, Rosa26+: 297 bp).

PCR using genomic DNA extracted from brain

Each brain was harvested and divided in half at the midline and frozen with liquid nitrogen. Genomic DNA was extracted using ISOSPIN Tissue DNA Kit (NIPPON GENE; 316-08891). PCR was performed using the following primers. Tfap2b: 5′-AAGGCGCATAACGATACCAC-3′ and 5′-ACTGATGGCGAGCTCAGACC-3′ (Tfap2b: 174 bp, Tfap2bflox: 1024 bp), Chn1: 5′-AGGGCTTTCCTTGCTGTGTC-3′ and 5′-TAGGTCCCTTCTCATGAACC-3′ (Chn1+: 120 bp).

EEG/ electromyogram (EMG) and locomotor recording

EEG and EMG signals were recorded from freely moving mice at the age of 10–13 weeks according to the method described in a previous study [32], with some modifications. First, 8–10 weeks-old mice were implanted with EEG/EMG electrodes. Stainless steel EEG electrodes were implanted epidurally over the cerebellum and parietal cortex, and EMG electrodes were embedded into the trapezius muscles bilaterally. Mice were allowed to recover from the surgery in their home cage for at least 4 days. Then, mice were placed in a sleep recording chamber and habituated for at least 5 days. Subsequently, EEG/EMG signals were recorded from the onset of the light phase. EEG/EMG signals were filtered (bandpass 0.5–250 Hz), collected, and digitized at a 512 Hz sampling rate using VitalRecorder (Kissei Comtec). Video and infrared signals were also recorded from the start of the light phase. Locomotor activity detected by infrared signals was used to determine the onset of the active phase, which was automatically done using a template-matching algorithm in the ClockLab software. Briefly, the onset of the active phase was defined as the start point of 5 consecutive hours of high locomotor activity (above the 20% percentile of average locomotor activity) that followed 5 consecutive hours of low locomotor activity. 12 consecutive hours from the onset of the active phase were categorized as the active phase, and the remaining 12 consecutive hours were categorized as the inactive phase.

EEG/EMG analysis

EEG/EMG data were divided into 4-s epochs (time window), and EEG data were further subjected to fast Fourier transform analyses using SleepSign (Kissei Comtec). The vigilance state in each epoch was manually classified as REM sleep, NREM sleep, or wake based on absolute delta (0.5–4 Hz) power, theta (6–10 Hz) power to delta power ratio, and the integral of EMG signals. If a single epoch contained multiple states, the state with the highest occupancy was assigned. The EEG power spectrum of each state was calculated and normalized by EEG power at 16–30 Hz averaged across 24 h. Epochs that contained multiple stages or presumable large movement-derived artifacts in the EEG data were included in the stage analysis but excluded from the EEG power spectrum analysis.

Sleep deprivation (SD)

SD was performed for 5 h during ZT0-ZT5. We monitored EEG/EMG to detect sleep. The cage was changed at the start of SD. Whenever sleep was detected, Kimwipes or marbles were placed in or removed from the cage, or the cage was gently tapped.

Histology

Deeply anesthetized mice were killed by injecting a lethal dose of anesthetic and transcardially perfused with 0.1 M phosphate-buffered saline (PBS) followed by 4% paraformaldehyde (w/v) in 0.1 M PBS or 10% formalin neutral buffer solution. The brains were postfixed in the same fixative for 1 day and subsequently equilibrated in 30% sucrose (w/v) in PBS. The brains were sectioned at 40 μm using a sliding microtome (Yamato Kohki) or a cryostat (Leica).

Immunohistochemistry

Sections were washed 3 times in tris-buffered saline with Tween20 (TBST; 50 mM Tris-HCl, pH 7.6, 150 mM NaCl, 0.05% Tween20), once in 0.3% H2O2/TBST, and 3 times again in TBST. The sections were then blocked for 30 min in blocking buffer [tris-buffered saline with 0.5% Blocking Reagent (Perkin Elmer; FP1012)]. The sections were incubated with a primary antibody [1/2000 mouse anti-tyrosine hydroxylase (Sigma; T2928)]. After washing 3 times in TBST, the sections were incubated with a secondary antibody [1/500 horseradish peroxidase-conjugated donkey anti-mouse IgG (Abcam; ab7061)] for 2 h. After washing 3 times in TBST, the sections were incubated with Tyramide Signal Amplification (TSA) plus Cyanine 5 reagent (Perkin Elmer; NEL745001KT) for 30 min. Finally, after washing 4 times in TBST, the sections were mounted on a slide glass with mounting medium (ThermoFisher; TA-006-FM). Images were captured with a digital slide scanner (Carl Zeiss; AxioScan.Z1) or a confocal microscope (Carl Zeiss; LMS800).

Quantification and statistical analysis

The experimenter was blinded to the genotype during sleep scoring. All statistical analyses were performed using Prism9 (GraphPad), and statistical significance was set at p < 0.05. Where applicable, all statistical tests were 2-tailed.

Availability of data and materials

All data are available upon reasonable request to the corresponding author.

Abbreviations

REM sleep:

Rapid eye movement sleep

NREM sleep:

Non-rapid eye movement sleep

EEG:

Electroencephalogram

EMG:

Electromyogram

GFP:

Green fluorescent protein

TFAP2B:

Transcription factor AP-2β

KO:

Knocked out

cKD:

Conditional knockdown

WT:

Wild-type

NES:

Nestin

SYN1:

Synapsin 1

CHN1:

Chimerin 1

PVH:

Paraventricular hypothalamic nucleus

SC:

Superior colliculus

Pb:

Parabrachial nucleus

LC:

Locus coeruleus

NTS:

Nucleus of the solitary tract

SD:

Sleep deprivation

TH:

Tyrosine hydroxylase

ipRGCs:

Intrinsically photosensitive retinal ganglion cells

SEM:

Standard error of the mean

References

  1. Miyazaki S, Liu C-Y, Hayashi Y. Sleep in vertebrate and invertebrate animals, and insights into the function and evolution of sleep. Neurosci Res. 2017;118:3–12.

    Article  PubMed  Google Scholar 

  2. Shi G, Wu D, Ptáček LJ, Fu YH. Human genetics and sleep behavior. Curr Opin Neurobiol. 2017; 43–9.

  3. Davidson HR. A large family with patent ductus arteriosus and unusual face. J Med Genet. 1993;30:503–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Satoda M, Pierpont MEM, Diaz GA, Bornemeier RA, Gelb BD. Char syndrome, an inherited disorder with patent ductus arteriosus, maps to chromosome 6p12-p21. Circulation. 1999;99:3036–42.

    Article  CAS  PubMed  Google Scholar 

  5. Mani A, Radhakrishnan J, Farhi A, Carew KS, Warnes CA, Nelson-Williams C, et al. Syndromic patent ductus arteriosus: evidence for haploinsufficient TFAP2B mutations and identification of a linked sleep disorder. Proc Natl Acad Sci. 2005;102:2975–9.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  6. Konietzka J, Fritz M, Spiri S, McWhirter R, Leha A, Palumbos S, et al. Epidermal growth factor signaling promotes sleep through a combined series and parallel neural circuit. Curr Biol. 2020;30:1-16.e13.

    Article  CAS  PubMed  Google Scholar 

  7. Turek M, Lewandrowski I, Bringmann H. An AP2 transcription factor is required for a sleep-active neuron to induce sleep-like quiescence in C. elegans. Curr Biol. 2013;23:2215–23.

    Article  CAS  PubMed  Google Scholar 

  8. Turek M, Besseling J, Spies J-P, König S, Bringmann H. Sleep-active neuron specification and sleep induction require FLP-11 neuropeptides to systemically induce sleep. Elife. 2016;5:1–18.

    Article  Google Scholar 

  9. Kucherenko MM, Ilangovan V, Herzig B, Shcherbata HR, Bringmann H. TfAP-2 is required for night sleep in Drosophila. BMC Neurosci. 2016;17:72.

    Article  PubMed  PubMed Central  Google Scholar 

  10. Moser M, Rüschoff J, Buettner R. Comparative analysis of AP-2α and AP-2β gene expression during murine embryogenesis. Dev Dyn. 1997;208:115–24.

    Article  CAS  PubMed  Google Scholar 

  11. Moser M, Dahmen S, Kluge R, Gröne H, Dahmen J, Kunz D, et al. Terminal renal failure in mice lacking Transcription factor AP-2β. Lab Investig. 2003;83:571–8.

    Article  CAS  PubMed  Google Scholar 

  12. Nakai A, Fujiyama T, Nagata N, Kashiwagi M, Ikkyu A, Takagi M, et al. Sleep architecture in mice is shaped by the Transcription factor AP-2β. Genetics. 2020;216:753–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Hu Y, Korovaichuk A, Astiz M, Schroeder H, Islam R, Barrenetxea J, et al. Functional divergence of mammalian TFAP2a and TFAP2b transcription factors for bidirectional sleep control. Genetics. 2020;216:735–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Hu Y, Bringmann H. Tfap2b acts in GABAergic neurons to control sleep in mice. Sci Rep. 2023;13:8026.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  15. Hong SJ, Huh YH, Leung A, Choi HJ, Ding Y, Kang UJ, et al. Transcription factor AP-2β regulates the neurotransmitter phenotype and maturation of chromaffin cells. Mol Cell Neurosci. 2011;46:245–51.

    Article  CAS  PubMed  Google Scholar 

  16. Zainolabidin N, Kamath SP, Thanawalla AR, Chen AI. Distinct activities of Tfap2A and Tfap2B in the specification of GABAergic interneurons in the developing cerebellum. Front Mol Neurosci. 2017;10:1–14.

    Article  Google Scholar 

  17. Isaka F, Ishibashi M, Taki W, Hashimoto N, Nakanishi S, Kageyama R. Ectopic expression of the bHLH gene Math1 disturbs neural development. Eur J Neurosci. 1999;11:2582–8.

    Article  CAS  PubMed  Google Scholar 

  18. Iwasaki K, Fujiyama T, Nakata S, Park M, Miyoshi C, Hotta-Hirashima N, et al. Induction of mutant Sik3 sleepy allele in neurons in late infancy increases sleep need. J Neurosci. 2021;41:2733–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Declercq J, Brouwers B, Pruniau VPEG, Stijnen P, de Faudeur G, Tuand K, et al. Metabolic and behavioural phenotypes in Nestin-Cre mice are caused by hypothalamic expression of human growth hormone. PLoS ONE. 2015;10:e0135502.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Zhang Z, Liu W-Y, Diao Y-P, Xu W, Zhong Y-H, Zhang J-Y, et al. Superior colliculus GABAergic neurons are essential for acute dark induction of wakefulness in mice. Curr Biol. 2019;29:637-644.e3.

    Article  CAS  PubMed  Google Scholar 

  21. Yao Y, Barger Z, Saffari Doost M, Tso CF, Darmohray D, Silverman D, et al. Cardiovascular baroreflex circuit moonlights in sleep control. Neuron. 2022;110:3986-3999.e6.

    Article  CAS  PubMed  Google Scholar 

  22. Hong SJ, Lardaro T, Oh MS, Huh Y, Ding Y, Kang UJ, et al. Regulation of the noradrenaline neurotransmitter phenotype by the Transcription Factor AP-2β. J Biol Chem. 2008;283:16860–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Yamaguchi H, Hopf FW, Li S-B, de Lecea L. In vivo cell type-specific CRISPR knockdown of dopamine beta hydroxylase reduces locus coeruleus evoked wakefulness. Nat Commun. 2018;9:5211.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  24. Pogorelov VM, Rodriguiz RM, Insco ML, Caron MG, Wetsel WC. Novelty seeking and stereotypic activation of behavior in mice with disruption of the Dat1 gene. Neuropsychopharmacology. 2005;30:1818–31.

    Article  CAS  PubMed  Google Scholar 

  25. Won H, Lee H-R, Gee HY, Mah W, Kim J-I, Lee J, et al. Autistic-like social behaviour in Shank2-mutant mice improved by restoring NMDA receptor function. Nature. 2012;486:261–5.

    Article  ADS  CAS  PubMed  Google Scholar 

  26. Silverman JL, Yang M, Lord C, Crawley JN. Behavioural phenotyping assays for mouse models of autism. Nat Rev Neurosci. 2010;11:490–502.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Berson DM, Dunn FA, Takao M. Phototransduction by retinal ganglion cells that set the circadian clock. Science. 2002;295:1070–3.

    Article  ADS  CAS  PubMed  Google Scholar 

  28. Bassett EA, Korol A, Deschamps PA, Buettner R, Wallace VA, Williams T, et al. Overlapping expression patterns and redundant roles for AP-2 transcription factors in the developing mammalian retina. Dev Dyn. 2012;241:814–29.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Jin K, Jiang H, Xiao D, Zou M, Zhu J, Xiang M. Tfap2a and 2b act downstream of Ptf1a to promote amacrine cell differentiation during retinogenesis. Mol Brain. 2015;8:28.

    Article  PubMed  PubMed Central  Google Scholar 

  30. Kanki H, Suzuki H, Itohara S. High-efficiency CAG-FLPe deleter mice in C57BL/6J background. Exp Anim. 2006;55:137–41.

    Article  CAS  PubMed  Google Scholar 

  31. Liu J, Krautzberger AM, Sui SH, Hofmann OM, Chen Y, Baetscher M, et al. Cell-specific translational profiling in acute kidney injury. J Clin Invest. 2014;124:1242–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Hayashi Y, Kashiwagi M, Yasuda K, Ando R, Kanuka M, Sakai K, et al. Cells of a common developmental origin regulate REM/non-REM sleep and wakefulness in mice. Science. 2015;350:957–61.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank all the members of the Hayashi laboratory for their support, especially Chung-Kuan Chen, Kosho Asaga, and Shinichi Miyazaki for advice on behavioral analysis. We also thank Ruth Li for advice on locomotor activity analysis.

Funding

This work was supported by JST SPRING under Grant Number JPMJSP2124 (to AN); AMED under Grant Numbers JP19gm1110008, JP21wm0425018, and JP21zf0175005; JSPS KAKENHI under Grant Numbers JP22K19354, JP23H04668, and JP23H04210; the Asahi Glass Foundation; and the Kao Foundation for Research on Health Science (to YH).

Author information

Authors and Affiliations

Authors

Contributions

AN, MK, TF, KI, AH, HF, MY, TS, and YH designed and performed the experiments. AN analyzed the data. AN and YH wrote the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Yu Hayashi.

Ethics declarations

Ethics approval and consent to participate

All required animal ethics approvals have been obtained for this research.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1. Fig. S1.

 Confirmation of Tfap2b deletion and comparison of bodyweight in nervous system-specific Tfap2b cKD and control mice. Fig. S2. Locomotor activity of nervous system-specific Tfap2b cKD mice. Fig. S3. Comparison of the amount of sleep/wake in nervous system-specific Tfap2b cKD and control mice during the active and inactive phase determined by locomotor activity.  Fig. S4. Comparison of EEG power spectra in nervous system-specific Tfap2b cKD and control mice. Fig. S5. Confirmation of Tfap2b deletion and comparison of bodyweight in postnatal neuron-specific Tfap2b cKD and control mice. Fig. S6. Comparison of EEG power spectra in postnatal neuron-specific Tfap2b cKD and control mice. Fig. S7. Comparison of the sleep architecture after sleep deprivation (SD) in postnatal neuron-specific Tfap2b cKD and control mice. Fig. S8. Analyses of the Cre-mediated recombination pattern in Syn1CreERT2 mice at the LC.

Additional file 2. Video 1. Representative video of nervous system-specific Tfap2b cKD and control mice. (A-D) Video recordings of an individual Nes-Cre; Tfap2bflox/+ (A), Tfap2bflox/flox (B), Nes-Cre; Tfap2bflox/+ (C), or Nes-Cre; Tfap2bflox/flox (D) male mouse during the light phase.

Additional file 3: Video 2. Representative video of postnatal neuron-specific Tfap2b cKD and control mice. (A, B) Video recordings of an individual Syn1CreERT2; Tfap2bflox/flox (A) or Tfap2bflox/flox (B) male mouse injected with tamoxifen (at P14, 17, and 21) during the light phase.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Nakai, A., Kashiwagi, M., Fujiyama, T. et al. Crucial role of TFAP2B in the nervous system for regulating NREM sleep. Mol Brain 17, 13 (2024). https://doi.org/10.1186/s13041-024-01084-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13041-024-01084-8

Keywords