Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

A signalling pathway for transcriptional regulation of sleep amount in mice

Abstract

In mice and humans, sleep quantity is governed by genetic factors and exhibits age-dependent variation1,2,3. However, the core molecular pathways and effector mechanisms that regulate sleep duration in mammals remain unclear. Here, we characterize a major signalling pathway for the transcriptional regulation of sleep in mice using adeno-associated virus-mediated somatic genetics analysis4. Chimeric knockout of LKB1 kinase—an activator of AMPK-related protein kinase SIK35,6,7—in adult mouse brain markedly reduces the amount and delta power—a measure of sleep depth—of non-rapid eye movement sleep (NREMS). Downstream of the LKB1–SIK3 pathway, gain or loss-of-function of the histone deacetylases HDAC4 and HDAC5 in adult brain neurons causes bidirectional changes of NREMS amount and delta power. Moreover, phosphorylation of HDAC4 and HDAC5 is associated with increased sleep need, and HDAC4 specifically regulates NREMS amount in posterior hypothalamus. Genetic and transcriptomic studies reveal that HDAC4 cooperates with CREB in both transcriptional and sleep regulation. These findings introduce the concept of signalling pathways targeting transcription modulators to regulate daily sleep amount and demonstrate the power of somatic genetics in mouse sleep research.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: LKB1–SIK3 signalling inhibits HDAC4/5 to regulate NREMS.
Fig. 2: ABC knockout of HDAC4 and/or HDAC5 increases NREMS amount and delta power.
Fig. 3: ABC expression HDAC4/5(CN) reduces NREMS amount and delta power.
Fig. 4: HDAC4 regulates NREMS amount in posterior hypothalamus.
Fig. 5: HDAC4 and CREB coordinate to regulate sleep downstream of SIK3 kinase.
Fig. 6: HDAC4 and CREB function cooperatively in transcriptional regulation.

Similar content being viewed by others

Data availability

All reagents and raw data generated in this study are available from the corresponding author upon reasonable request. The raw and processed RNA-seq data have been deposited at NCBI Gene Expression Omnibus under accession GSE182907. The raw and processed ChIP–seq data have been deposited at NCBI Gene Expression Omnibus under accession GSE182805. Because of their large size, the source data for mouse sleep phenotype analysis will be provided upon reasonable request. Source data are provided with this paper.

Code availability

The automated SleepV software for video-based analysis of sleep–wake behaviours4 has been deposited at GitHub (https://github.com/wochiguodong/SleepV.git.)4. The custom-designed C++ language-based automated sleep staging software for EEG/EMG data has been deposited at Zenodo (https://doi.org/10.5281/zenodo.7102778).

References

  1. Sehgal, A. & Mignot, E. Genetics of sleep and sleep disorders. Cell 146, 194–207 (2011).

    Article  CAS  Google Scholar 

  2. Webb, J. M. & Fu, Y.-H. Recent advances in sleep genetics. Curr. Opin. Neurobiol. 69, 19–24 (2021).

    Article  CAS  Google Scholar 

  3. Zheng, L. & Zhang, L. The molecular mechanism of natural short sleep: a path towards understanding why we need to sleep. Brain Sci. Adv. 8, 165–172 (2022).

    Article  Google Scholar 

  4. Wang, G. et al. Somatic genetics analysis of sleep in adult mice. J. Neurosci. 42, 5617–5640 (2022).

    Article  CAS  Google Scholar 

  5. Lizcano, J. M. et al. LKB1 is a master kinase that activates 13 kinases of the AMPK subfamily, including MARK/PAR-1. EMBO J. 23, 833–843 (2004).

    Article  CAS  Google Scholar 

  6. Walkinshaw, D. R. et al. The tumor suppressor kinase LKB1 activates the downstream kinases SIK2 and SIK3 to stimulate nuclear export of class IIa histone deacetylases. J. Biol. Chem. 288, 9345–9362 (2013).

    Article  CAS  Google Scholar 

  7. Funato, H. et al. Forward-genetics analysis of sleep in randomly mutagenized mice. Nature 539, 378–383 (2016).

    Article  ADS  CAS  Google Scholar 

  8. Honda, T. et al. A single phosphorylation site of SIK3 regulates daily sleep amounts and sleep need in mice. Proc. Natl Acad. Sci. USA 115, 10458–10463 (2018).

    Article  ADS  CAS  Google Scholar 

  9. Li, X., Song, S., Liu, Y., Ko, S.-H. & Kao, H.-Y. Phosphorylation of the histone deacetylase 7 modulates its stability and association with 14-3-3 proteins. J. Biol. Chem. 279, 34201–34208 (2004).

    Article  CAS  Google Scholar 

  10. Wang, A. H. et al. Regulation of histone deacetylase 4 by binding of 14-3-3 proteins. Mol. Cell. Biol. 20, 6904–6912 (2000).

    Article  CAS  Google Scholar 

  11. Chen, Z. et al. The role of histone deacetylase 4 during chondrocyte hypertrophy and endochondral bone development. Bone Joint Res. 9, 82–89 (2020).

    Article  Google Scholar 

  12. Haberland, M., Montgomery, R. L. & Olson, E. N. The many roles of histone deacetylases in development and physiology: implications for disease and therapy. Nat. Rev. Genet. 10, 32–42 (2009).

    Article  CAS  Google Scholar 

  13. Grubbs, J. J., Lopes, L. E., van der Linden, A. M. & Raizen, D. M. A salt-induced kinase is required for the metabolic regulation of sleep. PLoS Biol. 18, e3000220 (2020).

    Article  CAS  Google Scholar 

  14. Banks, G. T. et al. Forward genetics identifies a novel sleep mutant with sleep state inertia and REM sleep deficits. Sci. Adv. 6, eabb3567 (2020).

    Article  ADS  CAS  Google Scholar 

  15. Sunagawa, G. A. et al. Mammalian reverse genetics without crossing reveals Nr3a as a short-sleeper gene. Cell Rep. 14, 662–677 (2016).

    Article  CAS  Google Scholar 

  16. Funato, H. & Yanagisawa, M. Hunt for mammalian sleep-regulating genes. Brain Sci. Adv. 7, 75–96 (2021).

    Google Scholar 

  17. Chan, K. Y. et al. Engineered AAVs for efficient noninvasive gene delivery to the central and peripheral nervous systems. Nat. Neurosci. 20, 1172–1179 (2017).

    Article  CAS  Google Scholar 

  18. Xu, J. et al. Regulation of sleep quantity and intensity by long and short isoforms of SLEEPY kinase. Sleep https://doi.org/10.1093/sleep/zsac198 (2022).

  19. Gruffat, H., Manet, E. & Sergeant, A. MEF2-mediated recruitment of class II HDAC at the EBV immediate early gene BZLF1 links latency and chromatin remodeling. EMBO Rep. 3, 141–146 (2002).

    Article  CAS  Google Scholar 

  20. Zhu, Y. et al. Class IIa HDACs regulate learning and memory through dynamic experience-dependent repression of transcription. Nat. Commun. 10, 3469 (2019).

    Article  ADS  Google Scholar 

  21. Mihaylova, M. M. et al. Class IIa histone deacetylases are hormone-activated regulators of FOXO and mammalian glucose homeostasis. Cell 145, 607–621 (2011).

    Article  CAS  Google Scholar 

  22. Chen, B. & Cepko, C. L. HDAC4 regulates neuronal survival in normal and diseased retinas. Science 323, 256–259 (2009).

    Article  CAS  Google Scholar 

  23. Guo, X. et al. A short N-terminal domain of HDAC4 preserves photoreceptors and restores visual function in retinitis pigmentosa. Nat. Commun. 6, 8005 (2015).

    Article  ADS  CAS  Google Scholar 

  24. LeGates, T. A., Fernandez, D. C. & Hattar, S. Light as a central modulator of circadian rhythms, sleep and affect. Nat. Rev. Neurosci. 15, 443–454 (2014).

    Article  CAS  Google Scholar 

  25. Price, V., Wang, L. & D’Mello, S. R. Conditional deletion of histone deacetylase-4 in the central nervous system has no major effect on brain architecture or neuronal viability. J. Neurosci. Res. 91, 407–415 (2013).

    Article  CAS  Google Scholar 

  26. Fischle, W. et al. Enzymatic activity associated with class II HDACs is dependent on a multiprotein complex containing HDAC3 and SMRT/N-CoR. Mol. Cell 9, 45–57 (2002).

    Article  CAS  Google Scholar 

  27. Gvilia, I., Xu, F., McGinty, D. & Szymusiak, R. Homeostatic regulation of sleep: a role for preoptic area neurons. J. Neurosci. 26, 9426–9433 (2006).

    Article  CAS  Google Scholar 

  28. Kroeger, D. et al. Galanin neurons in the ventrolateral preoptic area promote sleep and heat loss in mice. Nat. Commun. 9, 4129 (2018).

    Article  ADS  Google Scholar 

  29. Yoo, S. H. et al. PERIOD2::LUCIFERASE real-time reporting of circadian dynamics reveals persistent circadian oscillations in mouse peripheral tissues. Proc. Natl Acad. Sci. USA 101, 5339–5346 (2004).

    Article  ADS  CAS  Google Scholar 

  30. Zhang, E. E. et al. A genome-wide RNAi screen for modifiers of the circadian clock in human cells. Cell 139, 199–210 (2009).

    Article  CAS  Google Scholar 

  31. Okamoto, S. et al. Dominant-interfering forms of MEF2 generated by caspase cleavage contribute to NMDA-induced neuronal apoptosis. Proc. Natl Acad. Sci. USA 99, 3974–3979 (2002).

    Article  ADS  CAS  Google Scholar 

  32. Kozhemyakina, E., Cohen, T., Yao, T.-P. & Lassar, A. B. Parathyroid hormone-related peptide represses chondrocyte hypertrophy through a protein phosphatase 2A/histone deacetylase 4/MEF2 pathway. Mol. Cell. Biol. 29, 5751 (2009).

    Article  CAS  Google Scholar 

  33. Lundell, L. S., Massart, J., Altintas, A., Krook, A. & Zierath, J. R. Regulation of glucose uptake and inflammation markers by FOXO1 and FOXO3 in skeletal muscle. Mol. Metab. 20, 79–88 (2019).

    Article  CAS  Google Scholar 

  34. Ahn, S. et al. A dominant-negative inhibitor of CREB reveals that it is a general mediator of stimulus-dependent transcription of c-fos. Mol. Cell Biol. 18, 967–977 (1998).

    Article  CAS  Google Scholar 

  35. Graves, L. A. et al. Genetic evidence for a role of CREB in sustained cortical arousal. J. Neurophysiol. 90, 1152–1159 (2003).

    Article  CAS  Google Scholar 

  36. Hendricks, J. C. et al. A non-circadian role for cAMP signaling and CREB activity in Drosophila rest homeostasis. Nat. Neurosci. 4, 1108–1115 (2001).

    Article  CAS  Google Scholar 

  37. Wimmer, M. E., Cui, R., Blackwell, J. M. & Abel, T. Cyclic AMP response element-binding protein is required in excitatory neurons in the forebrain to sustain wakefulness. Sleep 44, zsaa267 (2020).

    Article  Google Scholar 

  38. Lamph, W. W., Dwarki, V. J., Ofir, R., Montminy, M. & Verma, I. M. Negative and positive regulation by transcription factor cAMP response element-binding protein is modulated by phosphorylation. Proc. Natl Acad. Sci. USA 87, 4320–4324 (1990).

    Article  ADS  CAS  Google Scholar 

  39. Li, J. et al. Nuclear accumulation of HDAC4 in ATM deficiency promotes neurodegeneration in ataxia telangiectasia. Nat. Med. 18, 783–790 (2012).

    Article  Google Scholar 

  40. Sen, T. & Sen, N. Isoflurane-induced inactivation of CREB through histone deacetylase 4 is responsible for cognitive impairment in developing brain. Neurobiol. Dis. 96, 12–21 (2016).

    Article  CAS  Google Scholar 

  41. Zada, D., Bronshtein, I., Lerer-Goldshtein, T., Garini, Y. & Appelbaum, L. Sleep increases chromosome dynamics to enable reduction of accumulating DNA damage in single neurons. Nat. Commun. 10, 895 (2019).

    Article  ADS  CAS  Google Scholar 

  42. Zada, D. et al. Parp1 promotes sleep, which enhances DNA repair in neurons. Mol. Cell 81, 4979–4993.e4977 (2021).

    Article  CAS  Google Scholar 

  43. Brüning, F. et al. Sleep-wake cycles drive daily dynamics of synaptic phosphorylation. Science 366, eaav3617 (2019).

    Article  Google Scholar 

  44. Wang, Z. et al. Quantitative phosphoproteomic analysis of the molecular substrates of sleep need. Nature 558, 435–439 (2018).

    Article  ADS  CAS  Google Scholar 

  45. Benito, E., Valor, L. M., Jimenez-Minchan, M., Huber, W. & Barco, A. cAMP response element-binding protein is a primary hub of activity-driven neuronal gene expression. J. Neurosci. 31, 18237–18250 (2011).

    Article  CAS  Google Scholar 

  46. Sando, R. 3rd et al. HDAC4 governs a transcriptional program essential for synaptic plasticity and memory. Cell 151, 821–834 (2012).

    Article  CAS  Google Scholar 

  47. Noya, S. B. et al. The forebrain synaptic transcriptome is organized by clocks but its proteome is driven by sleep. Science 366, eaav2642 (2019).

    Article  CAS  Google Scholar 

  48. Kim, S. J. et al. Kinase signalling in excitatory neurons regulates sleep quantityand depth. Nature https://doi.org/10.1038/s41586-022-05450-1 (2022).

  49. Lu, J., McKinsey, T. A., Zhang, C.-L. & Olson, E. N. Regulation of skeletal myogenesis by association of the MEF2 transcription factor with class II histone deacetylases. Mol. Cell 6, 233–244 (2000).

    Article  CAS  Google Scholar 

  50. Sanjana, N. E., Shalem, O. & Zhang, F. Improved vectors and genome-wide libraries for CRISPR screening. Nat. Methods 11, 783–784 (2014).

    Article  CAS  Google Scholar 

  51. Yardeni, T., Eckhaus, M., Morris, H. D., Huizing, M. & Hoogstraten-Miller, S. Retro-orbital injections in mice. Lab. Anim. 40, 155–160 (2011).

    Article  Google Scholar 

  52. Zhao, H. et al. Absence of stress response in dorsal raphe nucleus in modulator of apoptosis 1-deficient mice. Mol. Neurobiol. 56, 2185–2201 (2019).

    Article  CAS  Google Scholar 

  53. Christakis, D. A., Ramirez, J. S. & Ramirez, J. M. Overstimulation of newborn mice leads to behavioral differences and deficits in cognitive performance. Sci. Rep. 2, 546 (2012).

    Article  ADS  CAS  Google Scholar 

  54. Horsch, M. et al. Requirement of the RNA-editing enzyme ADAR2 for normal physiology in mice. J. Biol. Chem. 286, 18614–18622 (2011).

    Article  CAS  Google Scholar 

  55. Liu, Q., Xu, Y., Wan, W. & Ma, Z. An unexpected improvement in spatial learning and memory ability in alpha-synuclein A53T transgenic mice. J. Neural Transm. 125, 203–210 (2018).

    Article  CAS  Google Scholar 

  56. Peters, J., Dieppa-Perea, L. M., Melendez, L. M. & Quirk, G. J. Induction of fear extinction with hippocampal-infralimbic BDNF. Science 328, 1288–1290 (2010).

    Article  ADS  CAS  Google Scholar 

  57. Soria-Gomez, E. et al. Habenular CB1 receptors control the expression of aversive memories. Neuron 88, 306–313 (2015).

    Article  CAS  Google Scholar 

  58. Zhang, J. et al. Presynaptic Excitation via GABAB receptors in habenula cholinergic neurons regulates fear memory expression. Cell 166, 716–728 (2016).

    Article  CAS  Google Scholar 

  59. Patti, C. L. et al. Effects of sleep deprivation on memory in mice: role of state-dependent learning. Sleep 33, 1669–1679 (2010).

    Article  Google Scholar 

  60. Savelyev, S. A., Larsson, K. C., Johansson, A. S. & Lundkvist, G. B. Slice preparation, organotypic tissue culturing and luciferase recording of clock gene activity in the suprachiasmatic nucleus. J. Vis. Exp. 48, 2439 (2011).

    Google Scholar 

  61. Mei, L. et al. Long-term in vivo recording of circadian rhythms in brains of freely moving mice. Proc. Natl Acad. Sci. USA 115, 4276–4281 (2018).

    Article  ADS  CAS  Google Scholar 

  62. Yang, J. et al. A quick protocol for the preparation of mouse retinal cryosections for immunohistochemistry. Open Biol. 11, 210076 (2021).

    Article  Google Scholar 

  63. Castel, M., Belenky, M., Cohen, S., Wagner, S. & Schwartz, W. J. Light-induced c-Fos expression in the mouse suprachiasmatic nucleus: immunoelectron microscopy reveals co-localization in multiple cell types. Eur. J. Neurosci. 9, 1950–1960 (1997).

    Article  CAS  Google Scholar 

  64. Xu, P. et al. NPAS4 regulates the transcriptional response of the suprachiasmatic nucleus to light and circadian behavior. Neuron 109, 3268–3282 (2021).

    Article  CAS  Google Scholar 

  65. Ju, D. et al. Chemical perturbations reveal that RUVBL2 regulates the circadian phase in mammals. Sci. Transl. Med. 12, eaba0769 (2020).

    Article  CAS  Google Scholar 

  66. Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357 (2012).

    Article  CAS  Google Scholar 

  67. Zhang, Y. et al. Model-based analysis of ChIP–seq (MACS). Genome Biol. 9, R137 (2008).

    Article  Google Scholar 

  68. Ramírez, F., Dündar, F., Diehl, S., Grüning, B. A. & Manke, T. deepTools: a flexible platform for exploring deep-sequencing data. Nucleic Acids Res. 42, W187–W191 (2014).

    Article  Google Scholar 

  69. Yu, G., Wang, L. G. & He, Q. Y. ChIPseeker: an R/Bioconductor package for ChIP peak annotation, comparison and visualization. Bioinformatics 31, 2382–2383 (2015).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank M. Luo and F. Shao for sharing reagents; Y. Zhou, M. Zhao, Z. Li, W. Min, Y. Zhuang, Y. Sun, H. Huang, J. Wang and Y. Yin for technical assistance; and X. Wang, F. Shao, T. Chen, T. Han and M. Xu for discussion and comments on the manuscript. This work was supported by the National Major Project of China Science and Technology Innovation 2030 for Brain Science and Brain-Inspired Technology (2021ZD0203400 to Q.Liu), the innovation grant (Z181100001318004 to Q.Liu), and institution grants from the Beijing Municipal Commission of Science and Technology Commission and Chinese Ministry of Science and Technology.

Author information

Authors and Affiliations

Authors

Contributions

R.Z., G.W., Q.Li and J.X. executed most of the experiments with input from Q.Liu, E.E.Z., H.F. and M.Y. F.W. helped to generate multiple mouse strains. X.G., S.Z., Z.C., L.C. and B.S. performed EEG/EMG recording and analysis. R.Z. and D.J. performed RNA-seq and ChIP–seq analyses with help from B.L., R.G. and T.C. D.S. performed ex vivo recording of SCN of Per2::Luc mice. Q.S., Y.G. and C.W. performed AAV packaging and purification. F.M. and C.L. performed stereotactic injection of AAVs in the hypothalamus and thalamus. H.W., X.W. and H.L. helped with mouse husbandry. R.Z., G.W. and Q.Liu prepared the figures and Q.Liu wrote the manuscript.

Corresponding author

Correspondence to Qinghua Liu.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature thanks the anonymous reviewer(s) for their contribution to the peer review of this work.

Additional information

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

Extended data figures and tables

Extended Data Fig. 1 LKB1-SIK3 signalling regulates NREMS by suppressing HDAC4/5.

a, Representative images showing co-immunostaining of LKB1 (green) and mCherry or Cre (HA, red) neurons in the cortex, hippocampus, thalamus, midbrain and medulla of AAV-hSyn-mCherry or AAV-hSyn-cre injected Lkb1flox/flox mice. Scale bars, 50 μm. b and c, REMS (b) or Wake (c) amount of AAV-hSyn-mCherry (control (Ctrl), n = 11) or AAV-hSyn-cre (Lkb1KO, n = 12) injected Lkb1flox/flox mice. d-g, Hourly NREMS time (d), EEG power spectra analysis (e), and REMS (f) or Wake (g) amount of Lkb1flox/flox mice co-injected with AAV-hSyn-cre and AAV-hSyn-mCherry (Lkb1KO/mCherry, n = 7), AAV-hSyn-SIK3-ST221E (Lkb1KO/SIK3-S(T221E), n = 7), or AAV-hSyn-SLP-ST221E (Lkb1KO/SLP-S(T221E), n = 8). h, Schematic of H4N/H5N-VP16 interfering with transcriptional activities of endogenous HDAC4/5 proteins. i, Representative images (two independent experiments with similar results) showing the effects of H4N-VP16 and H5N-VP16 on expression of co-transfected 3x MEF2 response element (MRE)-copGFP reporter gene in AAVpro 293T cells. Scale bars, 200 μm. j-m, Hourly NREMS time (j), EEG power spectra analysis (k), and REMS (l) or Wake (m) amount of Lkb1flox/flox mice injected with AAV-hSyn-cre and AAV-hSyn-mCherry (Lkb1KO/mCherry, n = 7) or AAV-hSyn-H4N/H5N-VP16 (Lkb1KO/H4N/H5N-VP16, n = 6). Data are mean ± s.e.m. Two-way ANOVA with Sidak’s test (b, c, l and m); mixed-effects model (d and j); one-way ANOVA with Dunnett’s test (e); two-way ANOVA with Dunnett’s test (f and g); and two-tailed unpaired t-test (k). n refers to the number of biological replicates.

Source data

Extended Data Fig. 2 ABC knockout of HDAC4/5 increases NREMS amount and delta power.

a-d, Hourly NREMS time (a), EEG power spectra analysis (b), and REMS (c) or Wake (d) amount of C57BL/6J mice injected with AAV-hSyn-mCherry (n = 15), AAV-hSyn-H4N-VP16 (n = 7) or AAV-hSyn-H4N/H5N-VP16 (n = 7). e, Schematic of ABC-KO of Hdac4 or Hdac5 by triple-target CRISPR. f and g, Immunoblotting (f) and quantification (g) of HDAC4 and HDAC5 proteins in whole brain lysates of Cas9 mice (n = 3 per group) injected with AAV-sgRNANT (ABC-NT), AAV-sgRNAHD4 (ABC-Hdac4KO), AAV-sgRNAHD5 (ABC-Hdac5KO), or AAV-sgRNAHD4 and AAV-sgRNAHD5 (ABC-Hdac4/5DKO). h-m, NREMS (h), REMS (i) or Wake (j) amount, hourly NREMS time (k), delta power density (l) and EEG power spectra analysis (m) of ABC-NT, ABC-Hdac4KO, ABC-Hdac5KO, or ABC- Hdac4/5DKO mice (n = 8 per group). n-q, Hourly NREMS time (n), EEG power spectra analysis (o), and REMS (p) or Wake (q) amount of AAV-hSyn-mCherry (ABC-Hdac4WT, n = 10) or AAV-hSyn-cre (ABC-Hdac4KO, n = 9) injected Hdac4flox/flox mice and AAV-hSyn-mCherry (ABC-Hdac5WT, n = 11) or AAV-hSyn-cre (ABC-Hdac5KO, n = 8) injected Hdac5flox/flox mice. Data are mean ± s.e.m. Mixed-effects model (a, k, l and n); one-way ANOVA with Dunnett’s test (b, g and m); two-way ANOVA with Sidak’s test (c-d, hj and p-q); and two-tailed unpaired t-test (o). See Supplementary Fig. 1 for gel source data. n refers to the number of biological replicates.

Source data

Extended Data Fig. 3 ABC-Hdac4KO mice display normal vision and behaviors.

a-d, Hourly NREMS time (a), EEG power spectra analysis (b), and REMS (c) or Wake (d) amount of AAV-hSyn-rTTA and AAV-TRE-cre co-injected Hdac4flox/flox (iHdac4KO) mice without (-Dox, n = 6) or with (+Dox, n = 8) doxycycline treatment. e, Representative images (two independent experiments with similar results) showing co-immunostaining of S-Opsin (green) and Rhodopsin (red) in the retina of AAV-hSyn-mCherry (ABC-Ctrl) or AAV-hSyn-cre (ABC-Hdac4KO) injected Hdac4flox/flox mice. Scale bars, 20 μm. f, Representative images showing immunostaining of c-Fos (green) in the SCN of ABC-Ctrl or ABC-Hdac4KO mice after 15 min of light exposure during the dark phase. Scale bars, 50 μm. g, Quantification of light-induced c-Fos expression in (f) (n = 3). h-l, Fear conditioning (h), open field (i, j), sucrose preference (k), and rotarod (l) behavioral tests of ABC-Ctrl (n = 7) and ABC-Hdac4KO (n = 8) mice. Data are mean ± s.e.m. Mixed-effects model (a); two-way ANOVA with Sidak’s test (c-d, i, and k-l); and two-tailed unpaired t-test, two tailed (b, g-h and j). n refers to the number of biological replicates.

Source data

Extended Data Fig. 4 ABC expression HDAC4(CN) or HDAC5(CN) reduces NREMS amount and delta power.

a-d, Hourly NREMS time (a), EEG power spectra analysis (b), and REMS (c) or Wake (d) amount of AAV-hSyn-mCherry (n = 9), AAV-hSyn-HDAC4WT (n = 10) or AAV-hSyn-HDAC4CN (n = 8) injected C57BL/6J mice. e-h, Hourly NREMS time (e), EEG power spectra analysis (f), and REMS (g) or Wake (h) amount of AAV-hSyn-mCherry (n = 18), AAV-hSyn-HDAC5WT (n = 10) or AAV-hSyn-HDAC5CN (n = 9) injected C57BL/6J mice. i-l, Hourly NREMS time (i), EEG power spectra analysis (j), and REMS (k) or Wake (l) amount of AAV-hSyn-mCherry (n = 8), AAV-hSyn-HDAC7WT (n = 8) or AAV-hSyn-HDAC7CN (n = 9) injected C57BL/6J mice. m-p, Hourly NREMS time (m), EEG power spectra analysis (n), and REMS (o) or Wake (p) amount of AAV-hSyn-mCherry (n = 8), AAV-hSyn-HDAC4CN (n = 9), AAV-hSyn-HDAC4CN/D934N (n = 9) or AAV-hSyn-HDAC4CN/D840N (n = 6) injected C57BL/6J mice. Data are mean ± s.e.m. Mixed-effects model (a, e and m); one-way ANOVA with Dunnett’s test (b, f, j and n); and two-way ANOVA with Dunnett’s test (c-d, g-h, k-l and o-p). n refers to the number of biological replicates.

Source data

Extended Data Fig. 5 Regulation of HDAC4/5 by sleep-wake cycle and homeostatic sleep response of ABC-Hdac4KO and ABC-Hdac4CN mice.

a and b, Immunoblotting (a) and quantification (b) of S245 phosphorylation of HDAC4/5 in total lysates of cortex from C57BL/6J mice at ZT2 and ZT14 (n = 3). c and d, Immunoblotting (c) and quantification (d) of nuclear:cytoplasmic ratio of HDAC4/5 in cytoplasmic and nuclear extracts of cortex from C57BL/6J mice at ZT2 and ZT14 (n = 3). e and i, Immunoblotting of total lysates of cortex (e) or hypothalamus (i) from 6-h slept (S6) and sleep-deprived (SD6) C57BL/6J mice. f and j, Quantification of S245 phosphorylation of HDAC4/5 in (e) and (i) (n = 3). g and k, Immunoblotting of cytoplasmic and nuclear extracts of cortex (g) or hypothalamus (k) from S6 and SD6 mice. h and l, Quantification of nuclear:cytoplasmic ratio of HDAC4/5 in (g) and (k) (n = 3). m-p, Hourly NREMS time (m) or delta power density (o) and accumulated NREMS time (n) or delta power (p) change of AAV-hSyn-GFP (Ctrl, n = 8) or AAV-hSyn-cre (Hdac4KO, n = 7) injected Hdac4flox/flox mice before (BS) and after sleep deprivation (SD). q-t, Hourly NREMS time (q) or delta power density (s) and accumulated NREMS time (r) or delta power change (t) of AAV-hSyn-mCherry (n = 9), AAV-hSyn-HDAC4WT (n = 10) or AAV-hSyn-HDAC4CN (n = 7) injected C57BL/6J mice before (BS) and after sleep deprivation (SD). Data are mean ± s.e.m. Two-tailed unpaired t-test (b, f, j and p); two-way ANOVA with Sidak’s test (d, h and l); mixed-effects model (m and q); and one-way ANOVA with Dunnett’s test (t). See Supplementary Fig. 1 for gel source data. n refers to the number of biological replicates.

Source data

Extended Data Fig. 6 HDAC4/5 regulates NREMS amount in posterior hypothalamus without altering circadian clock.

ad, Hourly NREMS time (a), EEG power spectra analysis (b), and REMS (c) or Wake (d) amount of C57BL/6J mice injected with AAV-hSyn-mCherry (n = 6), AAV-hSyn-HDAC4CN (n = 7) or AAV-hSyn-H4N/H5N-VP16 (n = 7) in posterior hypothalamus. e and f, Schematic and representative images showing HDAC4(CN) expression in SCN following stereotactic injection of AAV-hSyn-HDAC4CN in POA (n = 6) or posterior hypothalamus (n = 7). Scale bars, 1 mm. g, Representative ex vivo SCN slice bioluminescence recording of AAV-hSyn-mCherry, AAV-hSyn-HDAC4CN, or AAV-hSyn-H4N/H5N-VP16VP16 injected Per2::Luc mice (n = 3), and the short period control Cry1−/−;Per2::Luc mice (n = 1). Circadian time 0 was set as 24 h post medium change. h, Quantitative analysis of circadian periods in (g). Solid line represents the mean of SCN bioluminescence recording of AAV-hSyn-mCherry injected Per2::Luc mice, whereas dashed lines represent cut-offs of abnormality (≥ three standard deviations). i, RNAi knockdown of all four genes caused no clock phenotype in Bmal1:dLuc U2OS cells. Data were re-analyzed from a previous study30. siGL2 serves as the negative and siCRY2 as the positive controls for period changes. Circadian time 0 was set as 24 h post-medium change. j, Quantitative analyses of each knockdown result in (i) using Multicycle software. The solid line represents the mean of negative control, and dash lines indicate the cut-offs of abnormality (≥ three standard deviations). Each dot represents one experiment and duplicate experiments are applied for each gene. Data are mean ± s.e.m. Mixed-effects model (a); two-way ANOVA with Dunnett’s test (c and d); and one-way ANOVA with Dunnett’s test (h and j). n refers to the number of biological replicates.

Source data

Extended Data Fig. 7 HDAC4(CN) regulates NREMS amount in a CREB-dependent manner.

a, Schematics of wild-type and mutant MEF2A, MEF2C,MEF2D, FOXO1 and FOXO3. b-e, Hourly NREMS time (b), EEG power spectra analysis (c), and REMS (d) or Wake (e) amount of AAV-hSyn-eGFP (n = 21), AAV-hSyn-CREB (n = 14), or AAV-hSyn-A-CREB (n = 21) injected C57BL/6J mice. f, Representative images showing co-immunostaining of A-CREB-T7 (purple), HDAC4(CN)-HA (red) and NeuN (green) of cortical and thalamic sections from C57BL/6J mice co-injected with AAV-hSyn-HDAC4CN and AAV-hSyn-A-CREB. Scale bars, 20 μm. g, Quantification of the percentage of HDAC4(CN)+ neurons that co-express A-CREB and vice versa in the cortex and thalamus in (f) (n = 6). h-k, Hourly NREMS time (h), EEG power spectra analysis (i), and REMS (j) or Wake (k) amount of C57BL/6J mice injected with AAV-hSyn-mCherry (n = 6), AAV-hSyn-HDAC4CN and AAV-hSyn-mCherry (HD4(CN)+mCherry, n = 7), or AAV-hSyn-HDAC4CN and AAV-hSyn-A-CREB (HD4(CN)+A-CREB, n = 6). Data are mean ± s.e.m. Mixed-effects model (b and h); and two-way ANOVA with Dunnett’s test (d-e and j-k). n refers to the number of biological replicates.

Source data

Extended Data Fig. 8 HDAC4 and CREB coordinately regulate sleep downstream of SIK3 kinase.

a, Immunoblotting of brain lysates from AAV-hSyn-eGFP injected wild-type (WT+eGFP) mice and AAV-hSyn-eGFP (SLP+eGFP), AAV-hSyn-CREB (SLP+CREB), or AAV-hSyn-HDAC4CN (SLP+HD4(CN)) injected Sik3E13∆/+ mice. b, Daily sleep time in WT+eGFP (n = 10), SLP+eGFP (n = 9), SLP+CREB (n = 9), or SLP+HD4(CN) (n = 8) mice. c-f, Hourly NREMS time (c), EEG power spectra analysis (d), and REMS (e) or Wake (f) amount of WT+eGFP (n = 10), SLP+eGFP (n = 9), SLP+CREB (n = 6), or SLP+HD4(CN) (n = 8) mice. g, Immunoblotting of brain lysates from AAV-hSyn-eGFP (SLP+eGFP), AAV-hSyn-HDAC4WT (SLP+HD4(WT)), or AAV-hSyn-HDAC4S245A (SLP+HD4(S245A)) injected Sik3E13∆/+ mice (two independent experiments with similar results). h-m, Hourly NREMS time (h) and delta power density (i), EEG power spectra analysis (j), and NREMS (k), REMS (l), or Wake (m) amount of SLP+eGFP (n = 12), SLP+HD4(WT) (n = 8), or SLP+HD4(S245A) (n = 8) mice. n, Immunoblotting of brain lysates from AAV-hSyn-eGFP (SLP+eGFP), AAV-hSyn-HDAC5WT (SLP+HD5(WT)), or AAV-hSyn-HDAC5CN (SLP+HD5(CN)) injected Sik3E13∆/+ mice (two independent experiments with similar results). o-t, Hourly NREMS time (o) and delta power density (p), EEG power spectra analysis (q), and NREMS (r), REMS (s) or Wake (t) amount of SLP+eGFP (n = 8), SLP+HD5(WT) (n = 8), or SLP+HD5(CN) (n = 9) mice. Data are mean ± s.e.m. One-way ANOVA with Dunnett’s test (b, d and j); mixed-effects model (c, h, i and o); and two-way ANOVA with Dunnett’s test (e-f, k-m and r-t). See Supplementary Fig. 1 for gel source data. n refers to the number of biological replicates.

Source data

Extended Data Fig. 9 HDAC4 and CREB specifically rescue transcriptomic changes in the cortex and hypothalamus of Sik3E13∆/+ mice.

a, Venn diagram showing the common and unique sets of DEGs among the cortex (CTX), thalamus (THA) and hypothalamus (HPT) of SLP+eGFP mice. b, Venn diagrams showing the numbers of unique and overlapped DEGs among the SLP+eGFP, SLP+CREB and SLP+HDAC4(CN) mouse cortex, thalamus and hypothalamus, respectively. To identify DEGs, we compare the transcriptomes of WT+eGFP vs SLP+eGFP samples as well as the transcriptomes of SLP+eGFP vs SLP+CREB or SLP+HDAC4(CN) samples (n = 4 per group). c, A table listing overlapped DEGs in the cortex, thalamus and hypothalamus, respectively, among the SLP+eGFP (normalized to WT+eGFP) and SLP+CREB and SLP+HDAC4(CN) (normalized to SLP+eGFP) brain samples.

Extended Data Fig. 10 HDAC4 and CREB function cooperatively in transcriptional regulation.

a, Volcano plots of RNA-seq data from ABC-HDAC4(CN) or ABC-CREB mouse brain tissues (n = 4 for thalamus or hypothalamus, n = 3 for cortex) by binomial test via edgeR. The horizontal dashed line shows the two-sided unadjusted P = 0.05 as cutoff. The vertical dashed lines show +/-0.38-log2 fold changes as cutoff. Orange and blue dots indicate up- or down-regulated DEGs. b, RNA-seq heatmaps showing DEGs in HDAC4(CN) brain samples that are documented as target genes of MEF2, RUNX2, FOXO, or SRF, respectively. c, Percentage of CREB and/or HDAC4(CN)-bound genes (by ChIP-seq) in the DEGs of ABC-HDAC4(CN), ABC-CREB or both brain samples (by RNA-seq). d, Venn diagrams showing the numbers of unique and overlapped DEGs by comparing the RNA-seq and ChIP-seq results. e, Circadian oscillation of the overlapped DEGs (by RNA-seq) or HDAC4(CN)/CREB-bound genes (by ChIP-seq) based on previous circadian transcriptome analysis of the synaptosome or forebrain47. f, A model for transcriptional regulation of sleep amount in mice.

Supplementary information

Supplementary Fig. 1

The unprocessed Immunoblots associated with the data presented in main figures and extended data figures.

Reporting Summary

Supplementary Table 1

The DEGs of ABC-HDAC4CN or ABC-CREB mouse brain tissues (relative to the corresponding ABC-mCherry mouse brain tissues). Linked to Fig. 6.

Supplementary Table 2

The overlapping DEGs in the cortex (CTX), hypothalamus (HPT) and thalamus (THA) of ABC-HDAC4CN and ABC-CREB mouse brains (relative to the corresponding ABC-mCherry mouse brain tissues). Linked to Fig. 6.

Source data

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhou, R., Wang, G., Li, Q. et al. A signalling pathway for transcriptional regulation of sleep amount in mice. Nature 612, 519–527 (2022). https://doi.org/10.1038/s41586-022-05510-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-022-05510-6

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing