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:

Loss of function of NCOR1 and NCOR2 impairs memory through a novel GABAergic hypothalamus–CA3 projection

An Author Correction to this article was published on 20 June 2019

This article has been updated

Abstract

Nuclear receptor corepressor 1 (NCOR1) and NCOR2 (also known as SMRT) regulate gene expression by activating histone deacetylase 3 through their deacetylase activation domain (DAD). We show that mice with DAD knock-in mutations have memory deficits, reduced anxiety levels, and reduced social interactions. Mice with NCOR1 and NORC2 depletion specifically in GABAergic neurons (NS-V mice) recapitulated the memory deficits and had reduced GABAA receptor subunit α2 (GABRA2) expression in lateral hypothalamus GABAergic (LHGABA) neurons. This was associated with LHGABA neuron hyperexcitability and impaired hippocampal long-term potentiation, through a monosynaptic LHGABA to CA3GABA projection. Optogenetic activation of this projection caused memory deficits, whereas targeted manipulation of LHGABA or CA3GABA neuron activity reversed memory deficits in NS-V mice. We describe de novo variants in NCOR1, NCOR2 or HDAC3 in patients with intellectual disability or neurodevelopmental disorders. These findings identify a hypothalamus–hippocampus projection that may link endocrine signals with synaptic plasticity through NCOR-mediated regulation of GABA signaling.

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

Access options

Buy this article

Purchase on Springer Link

Instant access to full article PDF

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

Fig. 1: NS-DADm mice display cognitive dysfunction, altered anxiety, and social avoidance.
Fig. 2: NCORs regulate GABAA receptor gene expression in the hypothalamus.
Fig. 3: Depletion of NCORs specifically in GABAergic neurons causes memory deficits.
Fig. 4: NCORs control the excitation–inhibition balance of LH GABAergic neurons.
Fig. 5: The LHGABA to CA3GABA projection regulates memory and learning behaviors.
Fig. 6: The LHGABA to CA3GABA projection regulates hippocampal synaptic plasticity in NS-V mice.
Fig. 7: Genetic variants involving NCORs and HDAC3 in human with neurocognitive disorders.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available from the corresponding author upon request. RNA-seq and ChIP–seq data are available in GEO (GSE92452).

Change history

  • 20 June 2019

    In the version of this article initially published, the Acknowledgements erroneously included a grant number that did not directly support the work in the article. The last sentence of the Acknowledgments should have read, “The authors’ laboratories were supported by National Natural Science Foundation of China grants 31671222 and 31571556 (G.D.), a Taishan Scholarship (X.H.), the American Diabetes Association (ADA1–17-PDF-138) (Y.H.), the US Department of Agriculture (USDA) Cris6250-51000-059-04S (Y.X.), National Institutes of Health grants R01DK101379, R01DK117281, P01DK113954, R01DK115761 (Y.X.), the American Heart Association grant AHA30970064 (Z.S.), and grants R21CA215591 and R01ES027544 (Z.S.).” The error has been corrected in the HTML and PDF versions of the article.

References

  1. Rudenko, A. & Tsai, L.-H. Epigenetic regulation in memory and cognitive disorders. Neuroscience 264, 51–63 (2014).

    Article  CAS  Google Scholar 

  2. Lonard, D. M. & O’Malley, B. W. Nuclear receptor coregulators: modulators of pathology and therapeutic targets. Nat. Rev. Endocrinol. 8, 598–604 (2012).

    Article  CAS  Google Scholar 

  3. Mottis, A., Mouchiroud, L. & Auwerx, J. Emerging roles of the corepressors NCoR1 and SMRT in homeostasis. Genes Dev. 27, 819–835 (2013).

    Article  CAS  Google Scholar 

  4. Perissi, V., Jepsen, K., Glass, C. K. & Rosenfeld, M. G. Deconstructing repression: evolving models of co-repressor action. Nat. Rev. Genet. 11, 109–123 (2010).

    Article  CAS  Google Scholar 

  5. Guenther, M. G. et al. A core SMRT corepressor complex containing HDAC3 and TBL1, a WD40-repeat protein linked to deafness. Genes Dev. 14, 1048–1057 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Wen, Y. D. et al. The histone deacetylase-3 complex contains nuclear receptor corepressors. Proc. Natl. Acad. Sci. USA 97, 7202–7207 (2000).

    Article  CAS  Google Scholar 

  7. Li, J. et al. Both corepressor proteins SMRT and N-CoR exist in large protein complexes containing HDAC3. EMBO J. 19, 4342–4350 (2000).

    Article  CAS  Google Scholar 

  8. Guenther, M. G., Barak, O. & Lazar, M. A. The SMRT and N-CoR corepressors are activating cofactors for histone deacetylase 3. Mol. Cell. Biol. 21, 6091–6101 (2001).

    Article  CAS  Google Scholar 

  9. Sun, Z. et al. Hepatic Hdac3 promotes gluconeogenesis by repressing lipid synthesis and sequestration. Nat. Med. 18, 934–942 (2012).

    Article  CAS  Google Scholar 

  10. Sun, Z. et al. Deacetylase-independent function of HDAC3 in transcription and metabolism requires nuclear receptor corepressor. Mol. Cell 52, 769–782 (2013).

    Article  CAS  Google Scholar 

  11. Hong, S. et al. Dissociation of muscle insulin sensitivity from exercise endurance in mice by HDAC3 depletion. Nat. Med. 23, 223–234 (2017).

    Article  CAS  Google Scholar 

  12. Kokura, K. et al. The Ski protein family is required for MeCP2-mediated transcriptional repression. J. Biol. Chem. 276, 34115–34121 (2001).

    Article  CAS  Google Scholar 

  13. Amir, R. E. et al. Rett syndrome is caused by mutations in X-linked MECP2, encoding methyl-CpG-binding protein 2. Nat. Genet. 23, 185–188 (1999).

    Article  CAS  Google Scholar 

  14. Lyst, M. J. et al. Rett syndrome mutations abolish the interaction of MeCP2 with the NCoR/SMRT co-repressor. Nat. Neurosci. 16, 898–902 (2013).

    Article  CAS  Google Scholar 

  15. Ebert, D. H. et al. Activity-dependent phosphorylation of MeCP2 threonine 308 regulates interaction with NCoR. Nature 499, 341–345 (2013).

    Article  CAS  Google Scholar 

  16. Yoon, H.-G. et al. Purification and functional characterization of the human N-CoR complex: the roles of HDAC3, TBL1 and TBLR1. EMBO J. 22, 1336–1346 (2003).

    Article  CAS  Google Scholar 

  17. Ishizuka, T. & Lazar, M. A. The N-CoR/histone deacetylase 3 complex is required for repression by thyroid hormone receptor. Mol. Cell. Biol. 23, 5122–5131 (2003).

    Article  CAS  Google Scholar 

  18. You, S.-H. et al. Nuclear receptor co-repressors are required for the histone-deacetylase activity of HDAC3 in vivo. Nat. Struct. Mol. Biol. 20, 182–187 (2013).

    Article  CAS  Google Scholar 

  19. 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 

  20. Lahm, A. et al. Unraveling the hidden catalytic activity of vertebrate class IIa histone deacetylases. Proc. Natl. Acad. Sci. USA 104, 17335–17340 (2007).

    Article  CAS  Google Scholar 

  21. McQuown, S. C. et al. HDAC3 is a critical negative regulator of long-term memory formation. J. Neurosci. 31, 764–774 (2011).

    Article  CAS  Google Scholar 

  22. Lenègre, A., Chermat, R., Avril, I., Stéru, L. & Porsolt, R. D. Specificity of piracetam’s anti-amnesic activity in three models of amnesia in the mouse. Pharmacol. Biochem. Behav. 29, 625–629 (1988).

    Article  Google Scholar 

  23. Vong, L. et al. Leptin action on GABAergic neurons prevents obesity and reduces inhibitory tone to POMC neurons. Neuron 71, 142–154 (2011).

    Article  CAS  Google Scholar 

  24. Schnütgen, F. et al. A directional strategy for monitoring Cre-mediated recombination at the cellular level in the mouse. Nat. Biotechnol. 21, 562–565 (2003).

    Article  Google Scholar 

  25. Madisen, L. et al. A robust and high-throughput Cre reporting and characterization system for the whole mouse brain. Nat. Neurosci. 13, 133–140 (2010).

    Article  CAS  Google Scholar 

  26. Louis, G. W., Leinninger, G. M., Rhodes, C. J. & Myers, M. G. Jr. Direct innervation and modulation of orexin neurons by lateral hypothalamic LepRb neurons. J. Neurosci. 30, 11278–11287 (2010).

    Article  CAS  Google Scholar 

  27. Herman, A. M. et al. A cholinergic basal forebrain feeding circuit modulates appetite suppression. Nature 538, 253–256 (2016).

    Article  CAS  Google Scholar 

  28. Roth, B. L. DREADDs for Neuroscientists. Neuron 89, 683–694 (2016).

    Article  CAS  Google Scholar 

  29. Gomez, J. L. et al. Chemogenetics revealed: DREADD occupancy and activation via converted clozapine. Science 357, 503–507 (2017).

    Article  CAS  Google Scholar 

  30. Griebel, G. et al. SL651498: an anxioselective compound with functional selectivity for α2- and α3-containing ɣ-aminobutyric acidA (GABAA) receptors. J. Pharmacol. Exp. Ther. 298, 753–768 (2001).

    CAS  PubMed  Google Scholar 

  31. Laezza, F. & Dingledine, R. Voltage-controlled plasticity at GluR2-deficient synapses onto hippocampal interneurons. J. Neurophysiol. 92, 3575–3581 (2004).

    Article  CAS  Google Scholar 

  32. Pelkey, K. A., Topolnik, L., Yuan, X.-Q., Lacaille, J.-C. & McBain, C. J. State-dependent cAMP sensitivity of presynaptic function underlies metaplasticity in a hippocampal feedforward inhibitory circuit. Neuron 60, 980–987 (2008).

    Article  CAS  Google Scholar 

  33. Galván, E. J. et al. Synapse-specific compartmentalization of signaling cascades for LTP induction in CA3 interneurons. Neuroscience 290, 332–345 (2015).

    Article  Google Scholar 

  34. Lek, M. et al. Analysis of protein-coding genetic variation in 60,706 humans. Nature 536, 285–291 (2016).

    Article  CAS  Google Scholar 

  35. Firth, H. V. et al. DECIPHER: Database of Chromosomal Imbalance and Phenotype in Humans Using Ensembl Resources. Am. J. Hum. Genet. 84, 524–533 (2009).

    Article  CAS  Google Scholar 

  36. Watson, P. J., Fairall, L., Santos, G. M. & Schwabe, J. W. R. Structure of HDAC3 bound to co-repressor and inositol tetraphosphate. Nature 481, 335–340 (2012).

    Article  CAS  Google Scholar 

  37. Yang, L. et al. Hypocretin/orexin neurons contribute to hippocampus-dependent social memory and synaptic plasticity in mice. J. Neurosci. 33, 5275–5284 (2013).

    Article  CAS  Google Scholar 

  38. Selbach, O. et al. Orexins/hypocretins control bistability of hippocampal long-term synaptic plasticity through co-activation of multiple kinases. Acta Physiol. (Oxf). 198, 277–285 (2010).

    Article  CAS  Google Scholar 

  39. Jennings, J. H., Rizzi, G., Stamatakis, A. M., Ung, R. L. & Stuber, G. D. The inhibitory circuit architecture of the lateral hypothalamus orchestrates feeding. Science 341, 1517–1521 (2013).

    Article  CAS  Google Scholar 

  40. Jepsen, K. et al. Combinatorial roles of the nuclear receptor corepressor in transcription and development. Cell 102, 753–763 (2000).

    Article  CAS  Google Scholar 

  41. Norwood, J., Franklin, J. M., Sharma, D. & D’Mello, S. R. Histone deacetylase 3 is necessary for proper brain development. J. Biol. Chem. 289, 34569–34582 (2014).

    Article  Google Scholar 

  42. Nott, A. et al. Histone deacetylase 3 associates with MeCP2 to regulate FOXO and social behavior. Nat. Neurosci. 19, 1497–1505 (2016).

    Article  CAS  Google Scholar 

  43. Guy, J., Cheval, H., Selfridge, J. & Bird, A. The role of MeCP2 in the brain. Annu. Rev. Cell. Dev. Biol. 27, 631–652 (2011).

    Article  CAS  Google Scholar 

  44. Heckman, L. D., Chahrour, M. H. & Zoghbi, H. Y. Rett-causing mutations reveal two domains critical for MeCP2 function and for toxicity in MECP2 duplication syndrome mice. eLife 3, e02676 (2014).

    Article  Google Scholar 

  45. Li, W. & Pozzo-Miller, L. Beyond Widespread Mecp2 Deletions to Model Rett Syndrome: Conditional Spatio-Temporal Knockout, Single-Point Mutations and Transgenic Rescue Mice. Autism Open Access 2012(Suppl 1), 5 (2012).

    PubMed  Google Scholar 

  46. Chao, H.-T. et al. Dysfunction in GABA signalling mediates autism-like stereotypies and Rett syndrome phenotypes. Nature 468, 263–269 (2010).

    Article  CAS  Google Scholar 

  47. Chahrour, M. et al. MeCP2, a key contributor to neurological disease, activates and represses transcription. Science 320, 1224–1229 (2008).

    Article  CAS  Google Scholar 

  48. Chen, L. et al. MeCP2 binds to non-CG methylated DNA as neurons mature, influencing transcription and the timing of onset for Rett syndrome. Proc. Natl. Acad. Sci. USA 112, 5509–5514 (2015).

    Article  CAS  Google Scholar 

  49. Dani, V. S. et al. Reduced cortical activity due to a shift in the balance between excitation and inhibition in a mouse model of Rett syndrome. Proc. Natl. Acad. Sci. USA 102, 12560–12565 (2005).

    Article  CAS  Google Scholar 

  50. Mullican, S. E. et al. Histone deacetylase 3 is an epigenomic brake in macrophage alternative activation. Genes Dev. 25, 2480–2488 (2011).

    Article  CAS  Google Scholar 

  51. Zhou, W., Kavelaars, A. & Heijnen, C. J. Metformin prevents cisplatin-induced cognitive impairment and brain damage in mice. PLoS One 11, e0151890 (2016).

    Article  Google Scholar 

  52. Polter, A. et al. Forkhead box, class O transcription factors in brain: regulation and behavioral manifestation. Biol. Psychiatry 65, 150–159 (2009).

    Article  CAS  Google Scholar 

  53. Khalil, R. & Fendt, M. Increased anxiety but normal fear and safety learning in orexin-deficient mice. Behav. Brain. Res. 320, 210–218 (2017).

    Article  CAS  Google Scholar 

  54. Vorhees, C. V. & Williams, M. T. Morris water maze: procedures for assessing spatial and related forms of learning and memory. Nat. Protoc. 1, 848–858 (2006).

    Article  Google Scholar 

  55. Zhou, W. et al. The effects of glycogen synthase kinase-3beta in serotonin neurons. PLoS One 7, e43262 (2012).

    Article  CAS  Google Scholar 

  56. Zhou, W., Chen, L., Yang, S., Li, F. & Li, X. Behavioral stress-induced activation of FoxO3a in the cerebral cortex of mice. Biol. Psychiatry 71, 583–592 (2012).

    Article  CAS  Google Scholar 

  57. Armbruster, B. N., Li, X., Pausch, M. H., Herlitze, S. & Roth, B. L. Evolving the lock to fit the key to create a family of G protein-coupled receptors potently activated by an inert ligand. Proc. Natl. Acad. Sci. USA 104, 5163–5168 (2007).

  58. Krashes, M. J. et al. Rapid, reversible activation of AgRP neurons drives feeding behavior in mice. J. Clin. Invest. 121, 1424–1428 (2011).

    Article  CAS  Google Scholar 

  59. Grippo, R. M., Purohit, A. M., Zhang, Q., Zweifel, L. S. & Güler, A. D. Direct midbrain dopamine input to the suprachiasmatic nucleus accelerates circadian entrainment. Curr. Biol. 27, 2465–2475.e3 (2017).

    Article  CAS  Google Scholar 

  60. Ren, H. et al. FoxO1 target Gpr17 activates AgRP neurons to regulate food intake. Cell 149, 1314–1326 (2012).

    Article  CAS  Google Scholar 

  61. Liu, T. et al. Fasting activation of AgRP neurons requires NMDA receptors and involves spinogenesis and increased excitatory tone. Neuron 73, 511–522 (2012).

    Article  Google Scholar 

  62. Fenselau, H. et al. A rapidly acting glutamatergic ARC→PVH satiety circuit postsynaptically regulated by α-MSH. Nat. Neurosci. 20, 42–51 (2017).

    Article  CAS  Google Scholar 

  63. Smith, K. R. et al. Cadherin-10 maintains excitatory/inhibitory ratio through interactions with synaptic proteins. J. Neurosci. 37, 11127–11139 (2017).

    Article  CAS  Google Scholar 

  64. Yeckel, M. F., Kapur, A. & Johnston, D. Multiple forms of LTP in hippocampal CA3 neurons use a common postsynaptic mechanism. Nat. Neurosci. 2, 625–633 (1999).

    Article  CAS  Google Scholar 

  65. Urban, N. N., Henze, D. A., Lewis, D. A. & Barrionuevo, G. Properties of LTP induction in the CA3 region of the primate hippocampus. Learn. Mem. 3, 86–95 (1996).

    Article  CAS  Google Scholar 

  66. Khalilov, I., Minlebaev, M., Mukhtarov, M., Juzekaeva, E. & Khazipov, R. Postsynaptic GABA(B) receptors contribute to the termination of giant depolarizing potentials in CA3 neonatal rat hippocampus. Front. Cell. Neurosci. 11, 179 (2017).

    Article  Google Scholar 

  67. Petreanu, L., Huber, D., Sobczyk, A. & Svoboda, K. Channelrhodopsin-2-assisted circuit mapping of long-range callosal projections. Nat. Neurosci. 10, 663–668 (2007).

    Article  CAS  Google Scholar 

  68. Adhikari, A. et al. Basomedial amygdala mediates top-down control of anxiety and fear. Nature 527, 179–185 (2015).

    Article  CAS  Google Scholar 

  69. Bates, D., Mächler, M., Bolker, B. & Walker, S. Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67, 1–48 (2015).

  70. Kim, D. et al. TopHat2: accurate alignment of transcriptomes in the presence of insertions, deletions and gene fusions. Genome. Biol. 14, R36 (2013).

    Article  Google Scholar 

  71. Anders, S., Pyl, P. T. & Huber, W. HTSeq–a Python framework to work with high-throughput sequencing data. Bioinformatics 31, 166–169 (2015).

    Article  CAS  Google Scholar 

  72. Anders, S. & Huber, W. Differential expression analysis for sequence count data. Genome. Biol. 11, R106 (2010).

    Article  CAS  Google Scholar 

  73. Ng, P. C. & Henikoff, S. SIFT: predicting amino acid changes that affect protein function. Nucleic Acids Res. 31, 3812–3814 (2003).

    Article  CAS  Google Scholar 

  74. Adzhubei, I. A. et al. A method and server for predicting damaging missense mutations. Nat. Methods 7, 248–249 (2010).

    Article  CAS  Google Scholar 

  75. Richards, S. et al. Standards and guidelines for the interpretation of sequence variants: a joint consensus recommendation of the American College of Medical Genetics and Genomics and the Association for Molecular Pathology. Genet. Med. 17, 405–424 (2015).

    Article  Google Scholar 

  76. Faraco, G. et al. Dietary salt promotes neurovascular and cognitive dysfunction through a gut-initiated TH17 response. Nat. Neurosci. 21, 240–249 (2018).

    Article  CAS  Google Scholar 

  77. Peixoto, R. T., Wang, W., Croney, D. M., Kozorovitskiy, Y. & Sabatini, B. L. Early hyperactivity and precocious maturation of corticostriatal circuits in Shank3B˗/˗ mice. Nat. Neurosci. 19, 716–724 (2016).

    Article  CAS  Google Scholar 

  78. Witton, J. et al. Hippocampal circuit dysfunction in the Tc1 mouse model of Down syndrome. Nat. Neurosci. 18, 1291–1298 (2015).

    Article  CAS  Google Scholar 

  79. Xu, P. et al. Estrogen receptor-α in medial amygdala neurons regulates body weight. J. Clin. Invest. 125, 2861–2876 (2015).

    Article  Google Scholar 

  80. Perusini, J. N. et al. Optogenetic stimulation of dentate gyrus engrams restores memory in Alzheimer’s disease mice. Hippocampus 27, 1110–1122 (2017).

    Article  CAS  Google Scholar 

  81. Wang, W. et al. Chemogenetic activation of prefrontal cortex rescues synaptic and behavioral deficits in a mouse model of 16p11.2 deletion syndrome. J. Neurosci. 38, 5939–5948 (2018).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank Mitchell Lazar at the University of Pennsylvania for the NS-DADm and Hdac3loxP/loxP mice, Martin Myers Jr at University of Michigan for the WGA adenovirus, Bryan Roth at the University of North Carolina for the hM4Di AAV viral vector, Ben Arenkiel at Baylor College of Medicine (BCM) for the Syn-GFP AAV viral vector, Zhiping Pang at Rutgers University and Long-Jun Wu at Mayo Clinic for critical reading of the manuscript, Tim O’Brien at the University of Pennsylvania for technical consultancy in relation to behavior tests, Pingwen Xu and Justin Qian at BCM for technical assistance. We thank the BCM Neurobehavioral Core for providing space and training for behavioral tests (U54HD083092), the BCM Genomic and RNA Profiling Core (Texas Medical Center Digestive Disease Center P30DK56338) and Texas A&M AgriLife Research Genomics and Bioinformatics Service for nucleotide sequencing, the BCM Gene Vector core (Diabetes Research Center P30DK076938) for viral vector production, and the BCM RNA In Situ Hybridization Core (U54HD083092, P30DK056338, U42OD011174) for assistance with histology studies. This study makes use of data generated by the DECIPHER community. A full list of centers that contributed to the generation of the data is available from http://decipher.sanger.ac.uk and via e-mail from decipher@sanger.ac.uk. Funding for the project was provided by the Wellcome Trust. Those who carried out the original analysis and collection of the data bear no responsibility for the further analysis or interpretation of it by the recipient or its registered users. The DDD study presents independent research commissioned by the Health Innovation Challenge Fund (grant number HICF-1009-003), a parallel funding partnership between the Wellcome Trust and the UK Department of Health, and the Wellcome Sanger Institute (grant number WT098051). The views expressed in this publication are those of the authors and not necessarily those of the Wellcome or the UK Department of Health. The study has UK Research Ethics Committee (REC) approval (10/H0305/83, granted by the Cambridge South REC, and GEN/284/12 granted by the Republic of Ireland REC). The DDD research team acknowledges the support of the National Institute for Health Research, through the Comprehensive Clinical Research Network. The authors’ laboratories were supported by National Natural Science Foundation of China grants 31671222 and 31571556 (G.D.), a Taishan Scholarship (X.H.), the American Diabetes Association (ADA1–17-PDF-138) (Y.H.), the US Department of Agriculture (USDA) Cris6250-51000-059-04S (Y.X.), National Institutes of Health grants R01DK101379, R01DK117281, P01DK113954, R01DK115761 (Y.X.), the American Heart Association grant AHA30970064 (Z.S.), and grants R21CA215591 and R01ES027544 (Z.S.).

Author information

Authors and Affiliations

Authors

Consortia

Contributions

Z.S. conceived the study. W.Z., Y.H., Y.X., and Z.S. designed the experiments. W.Z. and C.W. performed and analyzed the studies involving behavior tests, injection surgery, histology analysis, ChIP, and gene expression analysis. Y.H. performed and analyzed the electrophysiology recordings. Y.K., S.H., G.D., and Y.G. performed and analyzed western blot, molecular cloning, immunoprecipitation, and HDAC assays. J.D., E.S., M.O., J.E.V.M., and the DDD research team provided the human sequencing data resource. A.U.R. and P.L. analyzed the human sequencing data. H.K.Y., Y.W., B.P., and Z.L. analyzed the ChIP-seq and RNA-seq data. W.Z. and H.L. analyzed the electrophysiology data. W.Z., Y.H., X.H., Q.W., Q.T., P.L., Y.X., and Z.S. interpreted the data. Y.X. and Z.S. secured the funding. Z.S. wrote the manuscript with input from the other authors.

Corresponding authors

Correspondence to Pengfei Liu, Yong Xu or Zheng Sun.

Ethics declarations

Competing interests

The authors declareno competing interests.

Additional information

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

Integrated supplementary information

Supplementary Figure 1 Characterization of NS-DADm mice.

(a) HDAC1 enzyme activity assay. HDAC1 was immunoprecipitated from the whole brain lysates followed by a fluorescence HDAC assay. Data were analyzed by two-tailed unpaired t test. nWT = 6, nNS-DADm = 6 mice. Male 5 months-old mice were used. (b) Brain weight of WT and NS-DADm mice. Data were analyzed by two-tailed unpaired t test. nWT = 11, nNS-DADm = 12 mice. Male 5months-old mice were used. (c) Body weight of WT and NS-DADm mice. Data were analyzed by two-tailed unpaired t test. nWT_2months = 10, nNS-DADm_2months = 9, nWT_4months = 8, nNS-DADm_4months = 8 mice. Male 2–4 months-old mice were used. (d) Body weight for Rotarod test. Data were analyzed by two-tailed unpaired t test. nWT = 8, nNS-DADm = 8 male mice. (e) Representative images of hindlimb activity. No hindlimb clamping was observed in WT or NS-DADm mice. The experiment was repeated independently once with similar results. n = 6 mice per group. (f) Representative images of plethysmography. (g) Respiration frequency and enhanced pause (Penh) during plethysmography. Data were analyzed by two-tailed unpaired t test. nWT = 5, nNS-DADm = 5 mice. Male 4 months-old mice were used. (h) Representative actogram of freewheel running in normal light/dark cycles. Male 4 months-old mice were used. (i) Representative images of Golgi staining of coronal brain sections. The experiment were repeated independently once with similar results. n = 3 mice for each group. Male 4 months-old mice were used. (j) Exploration time during the NOR test. Data were analyzed by two-way ANOVA. nWT = 18, nNS-DADm = 16 mice. Male 4 months-old mice were used. (k) Total exploration time during the NOR test. Data were analyzed by two-tailed unpaired t test. nWT = 18, nNS-DADm = 16 mice. Male 4 months-old mice were used. Data is expressed as mean ± S.E.M. For detailed statistics results, see Supplementary Table 1. * P ≤ 0.05 is set as significance.

Supplementary Figure 2 Characterization of mice with region-specific depletion of HDAC3.

(a) Scheme of intracranial injection of AAV-Cre at the CA3 in HDAC3loxP/loxP mice. (b) Immunostaining with anti-HDAC3 antibodies to confirm region-specific depletion of HDAC3 at CA3. Male 4 months-old mice were used. (c) Quantification of the immunostaining. CA1 served as a control. Data were analyzed by two-tailed unpaired t test. nGFP = 8, nCre = 8 mice. Male 4 months-old mice were used. (d) Discrimination index in NOR test on CA3HDAC3 knockout mice and control mice. Data were analyzed by two-tailed unpaired t test. nGFP = 8, nCre = 8 mice. Male 4 months-old mice were used. (e) Representative heat map of NOR test. White spots indicate novel or old objects in a 3-chamber box. Time duration that a mouse spent in exploring each object was denoted by color from dark blue (less time) to light blue (more time). (f) Total Exploration time in NOR. Data were analyzed by two-tailed unpaired t test. nGFP = 8, nCre = 8 mice. Male 4 months-old mice were used. (g-i) Morris Water Maze test, probe test, and total swimming distance on CA3Hdac3 knockout mice and control mice. Data were analyzed by two-tailed unpaired t test or one-way repeated ANOVA. NOR: nGFP = 8, nCre = 8 mice. Male 4 months-old mice were used. MWM: nGFP = 7, nCre = 7 mice. Male 4 months-old mice were used. (j) Scheme of intracranial injection of AAV-Cre at LH in HDAC3loxP/loxP mice. (k) Immunostaining with anti-HDAC3 antibodies to confirm region-specific depletion of HDAC3 at LH. 3V: 3rd ventricle, LH: lateral hypothalamus. Red: HDAC3, Blue: DAPI. Male 4 months-old mice were used. (l) Quantification of the immunostaining. Data were analyzed by two-tailed unpaired t test. nGFP = 8, nCre = 8 mice. Male 4 months-old mice were used. (m) Discrimination index in NOR test on LHHDAC3 knockout mice and control mice. Data were analyzed by two-tailed unpaired t test. NOR: nGFP = 9, nCre = 11 mice. Male 4 months-old mice were used. (n) Representative heat map of NOR test. (o) Total Exploration time in NOR test. Data were analyzed by two-tailed unpaired t test. NOR: nGFP = 9, nCre = 11 mice. Male 4 months-old mice were used. (p-r) Morris Water Maze test, probe test, and total swimming distance on LHHDAC3 knockout mice and control mice. Data were analyzed by two-tailed unpaired t test or one-way repeated ANOVA. nGFP = 9, nCre = 11 mice. Male 4 months-old mice were used. Data is expressed as mean ± S.E.M. For detailed statistics results, see Supplementary Table 1. * P ≤ 0.05 is set as significance.

Supplementary Figure 3 Gene and protein expression analysis in NS-DADm and NS-V mice.

(a) Quantification of the immunostaining analysis of GABRA2 in LH. Data were analyzed by two-tailed unpaired t test. nWT = 25 section/4 mice, nNS-DADm = 25 sections/4 mice. Male 4 months-old mice were used. (b) RT-qPCR analysis of mRNA of hippocampus from WT and NS-DADm mice. Box plots center line, median; box limits, upper and lower quartiles; whiskers, minimal and maximum values. Data were analyzed by two-tailed unpaired t test. n = 6 mice for each group. Male 4 months-old mice were used. (c) Representative immunofluorescence microscopy images and quantification of GABRA2 in CA3. Data were analyzed by two-tailed unpaired t test. nWT = 22 section/4 mice, nNS-DADm = 18 sections/4 mice. Male 4 months-old mice were used. (d) In-situ hybridization (ISH) analysis of NCOR1 confirms the ubiquitous expression of NCOR1 in the brain. The experiment was repeated independently once with similar results. n = 3 mice for each group. Male 4 months-old mice were used. (e) NCOR1-positive or/and glutamate decarboxylase 1 (GAD1)-positive cells in hippocampal CA3 of WT and NS-V mice. The experiment was repeated independently once with similar results. n = 3 mice for each group. Male 4 months-old mice were used. (f) NCOR1- or/and GAD1- positive cells in LH of WT and NS-V mice. Green: GAD1; Red: NCOR1; Yellow: co-expression of GAD1 and NCOR1. The experiment was repeated independently once with similar results. n = 3 mice for each group. Male 4 months-old mice were used. Arrows indicate cells positive for both NCOR1 and GAD1. Data is expressed as mean ± S.E.M. For detailed statistics results, see Supplementary Table 1. * P ≤ 0.05 is set as significance.

Supplementary Figure 4 The LHGABA → CA3GABA circuit mapping.

(a) Resting membrane potentials in NS-V and control Vgat-Cre (WT) mice as measured by a perforated clamp. Data were analyzed by two-way ANOVA. CA3: nWT = 16 neurons/2 mice, nNS-V = 16 neurons/2 mice; CA1: nWT = 15 neurons/2 mice, nNS-V = 13 neurons/2 mice; DG: nWT = 13 neurons/2 mice, nNS-V = 14 neurons/2 mice; LH: nWT = 14 neurons/2 mice, nNS-V = 15 neurons/2 mice. Male 2 months-old mice were used. (b) CRACM in NS-V mice. The amplitude of eIPSC of CA3GABA neurons in NS-V mice upon blue light stimulation of LHGABA neurons. Brain slices were recorded in pure artificial CSF (aCSF), or in the presence of 4-aminopyridine (4-AP, a potassium channel blocker) plus tetrodotoxin (TTX, a sodium channel blocker), or GABAA Receptor antagonist bicuculline (Bic). Data were analyzed by two-way ANOVA. naCSF = 12 neurons/3 mice, nNS-4AP+TTX = 7 neurons/3 mice, nbic = 10 neurons/3 mice. Male 4 months-old mice were used. (c) Representative trace of eIPSCs in CA3GABA neurons of NS-V mice. (d) The latency of eIPSCs in CA3GABA neurons of NS-V mice. Data were analyzed by two-way ANOVA. naCSF = 11 neurons/3 mice, nNS-4AP+TTX = 9 neurons/3 mice. Male 4 months-old mice were used. (e) Microscopy analysis of WGA-GFP in the projected site (CA3) in WT mice. Green: WGA-GFP, Blue: DAPI, Red: Glutamate positive neurons. The experiment was repeated independently once with similar results. n = 6 mice. Male 4 months-old mice were used. (f) Quantification of GFP-positive (green), Glutamate (red), and double-positive (yellow) cells in CA3. Male 4 months-old mice were used. n = 6 slides. (g) Immunofluorescence microscopy of contralateral LH and CA3 that did not receive Ad-iN/WGA-GFP. Green: WGA-GFP; Red: tdTomato; Blue: DAPI. The experiment was repeated independently once with similar results. n = 3 mice. Male 4 months-old mice were used. (h) Microscopy analysis of GFP-tagged Synaptophysin (Syn-GFP) at the primary injection site (LH) and the projected site (CA3). The experiment was repeated independently once with similar results. n = 3 mice. Female 5 months-old mice were used. Data is expressed as mean ± S.E.M. For detailed statistics results, see Supplementary Table 1. * P ≤ 0.05 is set as significance. Independently, the electrophysiological data was re-analyzed using the linear mixed-effects models (Supplementary Table 2).

Supplementary Figure 5 The LHGABA → CA3GABA projection in cognitive behaviors.

(a) Microscopy analysis of the expression of hM3Dq-mCherry in CA3. Red: mCherry. n(WT_mCherry) = 6, n(WT_hM3Dq) = 6, n(NS-V_mCherry) = 6, n(NS-V_hM3Dq) = 6 mice. The experiment was repeated independently once with similar results. (b) Exploration time during the NOR test in hM3Dq-injected or mCherry-injected NS-V or WT (Vgat-Cre) mice with CNO treatment. Data were analyzed by two-way ANOVA. nWT_mCherry = 6, nWT_hM3Dq = 6, nNS-V_mCherry = 6, nNS-V_hM3Dq = 6 mice. Female 4 months-old mice were used. (c-d) Escape latency and swimming distance during MWM test in hM3Dq-injected or mCherry-injected NS-V or WT (Vgat-Cre) mice with CNO treatment. Data were analyzed by two-way ANOVA followed by post hoc LSD test or repeated ANOVA. nWT_mCherry = 6, nWT_hM3Dq = 6, nNS-V_mCherry = 6, nNS-V_hM3Dq = 6 mice. Female 4 months-old mice were used. (e) Total exploration time on the objects during the NOR test with optogenetic stimulation of CA3GABA-innervating LHGABA neurons. Data were analyzed by two-tailed unpaired t test. noff = 7, non = 7 mice. Male 4 months-old mice were used. (f) Travel distance during the NOR test in hM4Di-injected NS-V or WT (Vgat-Cre) mice with CNO or Saline treatment. Data were analyzed by two-way ANOVA. nWT_Saline = 11, nWT_CNO = 11, nNS-V_Saline = 7, nNS-V_CNO = 7 mice. Male 4 months-old mice. (g) Swimming distance during the MWM test in hM4Di-injected NS-V or WT (Vgat-Cre) mice with CNO or Saline treatment. Data were analyzed by two-way ANOVA. nWT_Saline = 7, nWT_CNO = 7, nNS-V_Saline = 8, nNS-V_CNO = 7 mice. Male 4 months-old mice were used. (h) Microscopy analysis of eNpHR-EYFP at the primary injection site (LH). LH: lateral hypothalamus: 3V: 3rd ventricle. The experiment was repeated independently once with similar results. (i) Exploration time of NS-V or WT (Vgat-Cre) mice with AAV-FLEX-EYFP or AAV-FLEX-eNpHR injected at LH and yellow light shed on CA3. (Output power at the probe tip is around 7 mW). Data were analyzed by two-way ANOVA. nWT_EYFP = 7, nWT_eNpHR = 6, nNS-V_EYFP = 6, nNS-V_eNpHR = 8 mice. Male 4 months-old mice were used. (j) Scheme of intracranial infusion of SL651498. (k) Confirmation of accurate cannulation through dye injection. Data were analyzed by two-way ANOVA. nWT_Vehicle = 9, nWT_SL651498 = 9, nNS-V_Vehicle = 9, nNS-V_SL651498 = 10 mice. Male 4 months-old mice were used. (l) Local injection of selective GABAA positive modulator at LH does not change total distance or exploration time in the NOR test. Data were analyzed by two-way ANOVA. nWT_Vehicle = 9, nWT_SL651498 = 9, nNS-V_Vehicle = 9, nNS-V_SL651498 = 10 mice. Male 4 months-old mice were used. Data is expressed as mean ± S.E.M. For detailed statistics results, see Supplementary Table 1. * P ≤ 0.05 is set as significance.

Supplementary Figure 6 The LHGABA → CA3GABA projection in synaptic plasticity.

(a) The ratio of AMPAR-dependent sEPSC versus NMDAR-dependent sEPSC amplitudes in LHGABA-innervated CA3GABA neurons with or without MWM training. Data were analyzed by two-way ANOVA. nWT_Vehicle = 10 neurons/2 mice, nNS-V_Naive = 13 neurons/2 mice, nWT_Train = 10 neurons/2 mice, nNS-V_Train = 14 neurons/2 mice. Male 3 months-old mice were used. (b-c) Immunostaining of GABRA2 of LH from wild-type mice with or without MWM training. Green: GABRA2; Red: tdTomato; Blue: DAPI. The experiments were repeated independently once with similar results. n(WT_Train) = 6, n(WT_Naive) = 6 mice. Male 4 months-old mice were used. (d) The expression level of GABRA2 of LHGABA neurons from wild-type mice with or without MWM training. Box plots center line, median; box limits, upper and lower quartiles; whiskers, minimal and maximum values. Data were analyzed by two-tailed unpaired t test. nNaive = 73 neurons/12 mice, nTrain = 115 neurons/12 mice. Male 4 months-old mice were used. (e-h) Neural activities of CA3-projecting LHGABA neurons after retrograde beads injection at CA3 in tdTomato/Vgat-Cre mice. Data were analyzed by two-way ANOVA. nNaive = 33–34 neurons/4 mice, nTrain = 35–37 neurons/4 mice. Male 3 months-old mice were used. (i) The ratio of AMPAR-dependent sEPSC versus NMDAR-dependent sEPSC in hM4Di-manipulated LHGABA-innervated CA3GABA neurons in NS-V mice after MWM training. Data were analyzed by two-way ANOVA. nmCherry = 17 neurons/3 mice, nTrain = 18 neurons/3 mice. Male 3 months-old mice were used. Data is expressed as mean ± S.E.M. For detailed statistics results, see Supplementary Table 1. * P ≤ 0.05 is set as significance. Independently, the electrophysiological data was re-analyzed using the linear mixed-effects models (Supplementary Table 2).

Supplementary Figure 7 Original western blots.

(a) Original western blot of proteins associated with HDAC3 in HDAC3 immunoprecipitates (IP). (b) Original western blot of HEK-293 cells transfected with plasmids expressing wild-type (WT) HDAC3 with or without mutant L266S. (c) Original western blot of HEK-293 cells transfected with plasmids expressing WT HDAC3, WT NCOR1, with or without the NCOR1 deletion mutant (Del).

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhou, W., He, Y., Rehman, A.U. et al. Loss of function of NCOR1 and NCOR2 impairs memory through a novel GABAergic hypothalamus–CA3 projection. Nat Neurosci 22, 205–217 (2019). https://doi.org/10.1038/s41593-018-0311-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-018-0311-1

This article is cited by

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