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:

Hypothalamic melanin-concentrating hormone regulates hippocampus-dorsolateral septum activity

Abstract

Hypothalamic melanin-concentrating hormone (MCH) polypeptide contributes to regulating energy homeostasis, sleep and memory, although the mechanistic bases of its effects are unknown. In this study, in mice, we uncovered the physiological mechanism underlying the functional role of MCH signaling in projections to the dorsolateral septum (dLS), a region involved in routing hippocampal firing rhythms and encoding spatial memory based on such rhythms. Firing activity within the dLS in response to dorsal CA3 (dCA3) excitation is limited by strong feed-forward inhibition (FFI). We found that MCH synchronizes dLS neuronal firing with its dCA3 inputs by enhancing GABA release, which subsequently reduces the FFI and augments dCA3 excitatory input strength, both via pre-synaptic mechanisms. At the functional level, our data reveal a role for MCH signaling in the dLS in facilitating spatial memory. These findings support a model in which peptidergic signaling within the dLS modulates dorsal hippocampal output and supports memory encoding.

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: Endogenous MCH suppresses spontaneous neural activity in the dLS.
Fig. 2: MCH enhances excitatory synaptic transmission in the dLS .
Fig. 3: MCH enhances excitatory synaptic transmission in the dHP-to-dLS pathway.
Fig. 4: Augmentation of inhibitory synaptic transmission by MCH accounts for the suppressive effect on dLS neuronal activity and excitability.
Fig. 5: MCH enables high-frequency-coded information flow in the HP-dLS-LHA circuit.
Fig. 6: MCH is required for HP-dependent spatial learning and memory consolidation.

Similar content being viewed by others

Data availability

All data are contained in the main text, extended data or supplementary information. Source data can be downloaded at https://figshare.com/s/608dcd5f1a9d5a1d60df. Source data are provided with this paper.

Code availability

No custom code was used in this study.

References

  1. van den Pol, A. N. Neuropeptide transmission in brain circuits. Neuron 76, 98–115 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  2. Bargmann, C. I. & Marder, E. From the connectome to brain function. Nat. Methods 10, 483–490 (2013).

    Article  CAS  PubMed  Google Scholar 

  3. Marder, E. Neuromodulation of neuronal circuits: back to the future. Neuron 76, 1–11 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Qu, D. et al. A role for melanin-concentrating hormone in the central regulation of feeding behaviour. Nature 380, 243–247 (1996).

    Article  CAS  PubMed  Google Scholar 

  5. Bittencourt, J. C. et al. The melanin-concentrating hormone system of the rat brain: an immuno- and hybridization histochemical characterization. J. Comp. Neurol. 319, 218–245 (1992).

    Article  CAS  PubMed  Google Scholar 

  6. Bittencourt, J. C. Anatomical organization of the melanin-concentrating hormone peptide family in the mammalian brain. Gen. Comp. Endocrinol. 172, 185–197 (2011).

    Article  CAS  PubMed  Google Scholar 

  7. Steininger, T. L., Kilduff, T. S., Behan, M., Benca, R. M. & Landry, C. F. Comparison of hypocretin/orexin and melanin-concentrating hormone neurons and axonal projections in the embryonic and postnatal rat brain. J. Chem. Neuroanat. 27, 165–181 (2004).

    Article  CAS  PubMed  Google Scholar 

  8. Jego, S. et al. Optogenetic identification of a rapid eye movement sleep modulatory circuit in the hypothalamus. Nat. Neurosci. 16, 1637–1643 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Vetrivelan, R. et al. Melanin-concentrating hormone neurons specifically promote rapid eye movement sleep in mice. Neuroscience 336, 102–113 (2016).

    Article  CAS  PubMed  Google Scholar 

  10. Adamantidis, A. & de Lecea, L. A role for melanin-concentrating hormone in learning and memory. Peptides 30, 2066–2070 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Blouin, A. M. et al. Human hypocretin and melanin-concentrating hormone levels are linked to emotion and social interaction. Nat. Commun. 4, 1547 (2013).

    Article  PubMed  Google Scholar 

  12. Whiddon, B. B. & Palmiter, R. D. Ablation of neurons expressing melanin-concentrating hormone (MCH) in adult mice improves glucose tolerance independent of MCH signaling. J. Neurosci. 33, 2009–2016 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Domingos, A. I. et al. Hypothalamic melanin concentrating hormone neurons communicate the nutrient value of sugar. eLife 2, e01462 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  14. Adams, A. C. et al. Ablation of the hypothalamic neuropeptide melanin concentrating hormone is associated with behavioral abnormalities that reflect impaired olfactory integration. Behav. Brain Res. 224, 195–200 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Kosse, C. & Burdakov, D. Natural hypothalamic circuit dynamics underlying object memorization. Nat. Commun. 10, 2505 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Brailoiu, G. C. et al. Nesfatin-1: distribution and interaction with a G protein-coupled receptor in the rat brain. Endocrinology 148, 5088–5094 (2007).

    Article  CAS  PubMed  Google Scholar 

  17. Elias, C. F. et al. Characterization of CART neurons in the rat and human hypothalamus. J. Comp. Neurol. 432, 1–19 (2001).

    Article  CAS  PubMed  Google Scholar 

  18. Chee, M. J., Arrigoni, E. & Maratos-Flier, E. Melanin-concentrating hormone neurons release glutamate for feedforward inhibition of the lateral septum. J. Neurosci. 35, 3644–3651 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Izawa, S. et al. REM sleep-active MCH neurons are involved in forgetting hippocampus-dependent memories. Science 365, 1308–1313 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Pereira-Da-Silva, M., De Souza, C. T., Gasparetti, A. L., Saad, M. J. A. & Velloso, L. A. Melanin-concentrating hormone induces insulin resistance through a mechanism independent of body weight gain. J. Endocrinol. 186, 193–201 (2005).

    Article  CAS  PubMed  Google Scholar 

  21. Jeon, J. Y. et al. MCH−/− mice are resistant to aging-associated increases in body weight and insulin resistance. Diabetes 55, 428–434 (2006).

    Article  CAS  PubMed  Google Scholar 

  22. Adamantidis, A. et al. Disrupting the melanin-concentrating hormone receptor 1 in mice leads to cognitive deficits and alterations of NMDA receptor function. Eur. J. Neurosci. 21, 2837–2844 (2005).

    Article  PubMed  Google Scholar 

  23. Pachoud, B. et al. Major impairments of glutamatergic transmission and long-term synaptic plasticity in the hippocampus of mice lacking the melanin-concentrating hormone receptor-1. J. Neurophysiol. 104, 1417–1425 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Shimada, M., Tritos, N. A., Lowell, B. B., Flier, J. S. & Maratos-Flier, E. Mice lacking melanin-concentrating hormone are hypophagic and lean. Nature 396, 670–674 (1998).

    Article  CAS  PubMed  Google Scholar 

  25. Risold, P. Y. & Swanson, L. W. Connections of the rat lateral septal complex. Brain Res. Brain Res. Rev. 24, 115–195 (1997).

    Article  CAS  PubMed  Google Scholar 

  26. Swanson, L. W. & Cowan, W. M. The connections of the septal region in the rat. J. Comp. Neurol. 186, 621–655 (1979).

    Article  CAS  PubMed  Google Scholar 

  27. Tingley, D. & Buzsaki, G. Transformation of a spatial map across the hippocampal-lateral septal circuit. Neuron 98, 1229–1242 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Zhou, T. L., Tamura, R., Kuriwaki, J. & Ono, T. Comparison of medial and lateral septal neuron activity during performance of spatial tasks in rats. Hippocampus 9, 220–234 (1999).

    Article  CAS  PubMed  Google Scholar 

  29. Gonzalez, J. A., Iordanidou, P., Strom, M., Adamantidis, A. & Burdakov, D. Awake dynamics and brain-wide direct inputs of hypothalamic MCH and orexin networks. Nat. Commun. 7, 11395 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Hassani, O. K., Lee, M. G. & Jones, B. E. Melanin-concentrating hormone neurons discharge in a reciprocal manner to orexin neurons across the sleep-wake cycle. Proc. Natl Acad. Sci. USA 106, 2418–2422 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Blanco-Centurion, C. et al. Dynamic network activation of hypothalamic MCH neurons in REM sleep and exploratory behavior. J. Neurosci. 39, 4986–4998 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Kong, D. et al. Glucose stimulation of hypothalamic MCH neurons involves KATP channels, is modulated by UCP2, and regulates peripheral glucose homeostasis. Cell Metab. 12, 545–552 (2010).

    Article  CAS  PubMed  Google Scholar 

  33. Phelan, K. D., Hasuo, H., Twery, M. J. & Gallagher, J. P. Projection neurons in the rat dorsolateral septal nucleus possess recurrent axon collaterals. Neurosci. Lett. 97, 259–265 (1989).

    Article  CAS  PubMed  Google Scholar 

  34. Arrigoni, E. & Saper, C. B. What optogenetic stimulation is telling us (and failing to tell us) about fast neurotransmitters and neuromodulators in brain circuits for wake-sleep regulation. Curr. Opin. Neurobiol. 29, 165–171 (2014).

    Article  CAS  PubMed  Google Scholar 

  35. Park, J. Y. et al. Novel neuroprotective effects of melanin-concentrating hormone in Parkinson’s disease. Mol. Neurobiol. 54, 7706–7721 (2017).

    Article  CAS  PubMed  Google Scholar 

  36. Tan, C. P. et al. Melanin-concentrating hormone receptor subtypes 1 and 2: species-specific gene expression. Genomics 79, 785–792 (2002).

    Article  CAS  PubMed  Google Scholar 

  37. Saito, Y. et al. Molecular characterization of the melanin-concentrating-hormone receptor. Nature 400, 265–269 (1999).

    Article  CAS  PubMed  Google Scholar 

  38. Chee, M. J., Pissios, P. & Maratos-Flier, E. Neurochemical characterization of neurons expressing melanin-concentrating hormone receptor 1 in the mouse hypothalamus. J. Comp. Neurol. 521, 2208–2234 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Kokkotou, E. G., Tritos, N. A., Mastaitis, J. W., Slieker, L. & Maratos-Flier, E. Melanin-concentrating hormone receptor is a target of leptin action in the mouse brain. Endocrinology 142, 680–686 (2001).

    Article  CAS  PubMed  Google Scholar 

  40. Engle, S. E. et al. A CreER mouse to study melanin concentrating hormone signaling in the developing brain. Genesis 56, e23217 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  41. David, D. J. et al. Efficacy of the MCHR1 antagonist N-[3-(1-{[4-(3,4-difluorophenoxy)phenyl]methyl}(4-piperidyl))-4-methylphenyl]-2-m ethylpropanamide (SNAP 94847) in mouse models of anxiety and depression following acute and chronic administration is independent of hippocampal neurogenesis. J. Pharmacol. Exp. Ther 321, 237–248 (2007).

    Article  CAS  PubMed  Google Scholar 

  42. Owen, S. F., Liu, M. H. & Kreitzer, A. C. Thermal constraints on in vivo optogenetic manipulations. Nat. Neurosci. 22, 1061–1065 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Margeta-Mitrovic, M., Mitrovic, I., Riley, R. C., Jan, L. Y. & Basbaum, A. I. Immunohistochemical localization of GABAB receptors in the rat central nervous system. J. Comp. Neurol. 405, 299–321 (1999).

    Article  CAS  PubMed  Google Scholar 

  44. O’Keefe, J. & Dostrovsky, J. The hippocampus as a spatial map. Preliminary evidence from unit activity in the freely-moving rat. Brain Res. 34, 171–175 (1971).

    Article  PubMed  Google Scholar 

  45. Huxter, J., Burgess, N. & O’Keefe, J. Independent rate and temporal coding in hippocampal pyramidal cells. Nature 425, 828–832 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. O’Keefe, J. & Recce, M. L. Phase relationship between hippocampal place units and the EEG theta rhythm. Hippocampus 3, 317–330 (1993).

    Article  PubMed  Google Scholar 

  47. Ranck, J. B. Jr. Studies on single neurons in dorsal hippocampal formation and septum in unrestrained rats. I. Behavioral correlates and firing repertoires. Exp. Neurol. 41, 461–531 (1973).

    Article  PubMed  Google Scholar 

  48. Kjelstrup, K. B. et al. Finite scale of spatial representation in the hippocampus. Science 321, 140–143 (2008).

    Article  CAS  PubMed  Google Scholar 

  49. Takamura, Y. et al. Spatial firing properties of lateral septal neurons. Hippocampus 16, 635–644 (2006).

    Article  PubMed  Google Scholar 

  50. Conejo, N. M., Gonzalez-Pardo, H., Gonzalez-Lima, F. & Arias, J. L. Spatial learning of the water maze: progression of brain circuits mapped with cytochrome oxidase histochemistry. Neurobiol. Learn. Mem. 93, 362–371 (2010).

    Article  CAS  PubMed  Google Scholar 

  51. Leroy, F. et al. A circuit from hippocampal CA2 to lateral septum disinhibits social aggression. Nature 564, 213–218 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Antonawich, F. J., Melton, C. S., Wu, P. & Davis, J. N. Nesting and shredding behavior as an indicator of hippocampal ischemic damage. Brain Res. 764, 249–252 (1997).

    Article  CAS  PubMed  Google Scholar 

  53. 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).

    Article  PubMed  PubMed Central  Google Scholar 

  54. Schneeberger, M. et al. Functional analysis reveals differential effects of glutamate and MCH neuropeptide in MCH neurons. Mol. Metab. 13, 83–89 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Ferrante, M., Migliore, M. & Ascoli, G. A. Feed-forward inhibition as a buffer of the neuronal input-output relation. Proc. Natl Acad. Sci. USA 106, 18004–18009 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Gao, X. B. Electrophysiological effects of MCH on neurons in the hypothalamus. Peptides 30, 2025–2030 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Saito, Y., Cheng, M., Leslie, F. M. & Civelli, O. Expression of the melanin-concentrating hormone (MCH) receptor mRNA in the rat brain. J. Comp. Neurol. 435, 26–40 (2001).

    Article  CAS  PubMed  Google Scholar 

  58. Tingley, D. & Buzsaki, G. Routing of hippocampal ripples to subcortical structures via the lateral septum. Neuron 105, 138–149 (2020).

    Article  CAS  PubMed  Google Scholar 

  59. Rawlins, J. N. & Olton, D. S. The septo-hippocampal system and cognitive mapping. Behav. Brain Res. 5, 331–358 (1982).

    Article  CAS  PubMed  Google Scholar 

  60. Manning, J. R., Jacobs, J., Fried, I. & Kahana, M. J. Broadband shifts in local field potential power spectra are correlated with single-neuron spiking in humans. J. Neurosci. 29, 13613–13620 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Liu, J. J., Bello, N. T. & Pang, Z. P. Presynaptic regulation of leptin in a defined lateral hypothalamus-ventral tegmental area neurocircuitry depends on energy state. J. Neurosci. 37, 11854–11866 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Vogel-Ciernia, A. & Wood, M. A. Examining object location and object recognition memory in mice. Curr. Protoc. Neurosci. 69, 8.31.1–17 (2014).

    Article  Google Scholar 

  63. Glass, R., Norton, S., Fox, N. & Kusnecov, A. W. Maternal immune activation with staphylococcal enterotoxin A produces unique behavioral changes in C57BL/6 mouse offspring. Brain Behav. Immun. 75, 12–25 (2019).

    Article  CAS  PubMed  Google Scholar 

  64. 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  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We would like to thank G. Buzsáki and members of the Tsien laboratory, especially X. Wang and E. McCurdy, for their critical reading and constructive suggestions. We would like to thank A. Kusnecov, C. Manzini, B. Xia, A. Chang and Y. Zhang for their generous help in the behavioral tests and analyses. We would also like to thank A. Rabson and N. Francis of Rutgers for their input on the manuscript. We especially would like to thank B. Mensh, G. Edgerton and S. Aamodt for their critical input and suggestions on the manuscript. We also thank Z.-R. Xu, V. Li and J. Salvatore for suggestions on statistical analyses. The Pang laboratory is supported by NIH R01AA023797, NIH RF1MH120144 and, in part, by the Robert Wood Johnson Foundation (grant no. 74260) to the Child Health Institute of New Jersey.

Author information

Authors and Affiliations

Authors

Contributions

J.-J.L. and Z.P.P. conceived the study. J.-J.L. performed the experiments and analyzed the data. R.W.T. provided constructive intellectual input to the overall study. All authors contributed to the final manuscript.

Corresponding author

Correspondence to Zhiping P. Pang.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Neuroscience thanks Scott Kanoski and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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

Extended data

Extended Data Fig. 1 Optogenetic stimulation of LHA MCH-dLS may induce immediate impact on spontaneous dLS neuronal firing.

a, Wiring diagram of possible dLS neurocircuitry. b, Sample traces of optogenetically-induced oEPSCs and feed-forward-inhibitory oIPSCs. c, Representative traces of dLS neuronal spontaneous action potentials (sAPs) before, during, and after 10 s optogenetic stimulation of MCH fibers. d. Raster plot of sAPs. e-h, Normalized and pooled data of dLS neuronal firing under 10 s or 80 s, 10 Hz or 20 Hz stimulation. Data are mean ± SEM; numbers of neurons/animals analyzed are indicated. Paired two-tailed Student’s t-tests were used, e, before vs. during p = 0.622758; before vs. after p = 0.109378; f, before vs. during p = 0.01251; before vs. after p = 0.575929; g, before vs. during p = 0.0.875236; before vs. after p = 0.000832; h, before vs. during p = 0.000201; before vs. after p = 1.25×10−9. * p < 0.05, ** p < 0.01, ***p < 0.001.

Source data

Extended Data Fig. 2 MCH enhances excitatory transmission in dLS-to-LHA projecting neurons via a presynaptic mechanism.

a, Quantification of mRNAs of MCHR1 or vGAT in the dLS.n = 10. b, Sample traces of sEPSCs before and after trains of optogenetic activation of MCH fibers (left panel); right panel, Pooled data of the comparisons of frequencies and amplitudes of sEPSCs. n = 15 cells/8 animals and 16/9 cells/animals, at 5 Hz and 20 Hz, respectively. Paired Student’s t-tests were used: sEPSC frequency: Control vs. Post 5 Hz p = 0.0053; Control vs. Post 20 Hz p = 0.0051; sEPSC amplitude: Control vs. Post 5 Hz p = 0.0533; Control vs. Post 20 Hz p = 0.047. c, Sample traces of sEPSCs before and after optogenetic activation of MCH fibers in the presence of SNAP94847; right panel, Pooled data of the comparisons of frequencies and amplitudes of sEPSCs. n = 7 cells/3 animals and 10/4, at 5 Hz and 20 Hz, respectively. Student’s t-tests were used: sEPSC frequency: Control vs. Post 5 Hz p = 0.3281; Control vs. Post 20 Hz p = 0.5684; sEPSC amplitude: Control vs. Post 5 Hz p = 0.6994; Control vs. Post 20 Hz p = 0.8967. d&e, In the presence of Tc-MCH7c, no changes in spontaneous (d) or evoked EPSCs (e) were observed post 20 Hz optogenetic stimulation of MCH fibers.n = 9 cells /4 animals (d), 12/4 (e). Student’s t-tests were used: d, sEPSC frequency, Control vs. Post 20 Hz p = 0.584; sEPSC amplitude, Control vs. Post 20 Hz p = 0.9283; e, eEPSC amplitude, Control vs. Post 20 Hz p = 0.2007; PPR, Control vs. Post 20 Hz p = 0.4978. Data are mean ± SEM, * p < 0.05, ** p < 0.01, ***p < 0.001.

Source data

Extended Data Fig. 3 MCH regulates excitatory transmission in dLS neurons.

a-e, Impact of exogenously applied MCH and [Ala17]-MCH on sEPSCs, showing representative traces (a) and pooled data (b-e). n = 11 cells/6 animals for b&c; and 10/3 for d&e. Paired two-tailed Student’s t-tests were used: b, frequency p = 0.0006, amplitude p = 0.936; c, frequency p = 0.0025, amplitude p = 0.021.d-f, Expression of eYFP in MCH neurons after optogenetic stimulation does not impact synaptic transmission. d, Sample traces of PSCs and sEPSCs before and after trains of optogenetic stimulation (20 Hz 1 min x 2) of MCH fibers expressing eYFP. e&f, Pooled data of the normalized frequency or amplitude of PSCs (e) and sEPSCs (f). Data are mean ± SEM; n = 6 cells/3 animals for e; n = 11/3 for f. Paired two-tailed Student’s t-tests were used, no statistical significances were found. For PSCs, Frequency p = 0.8676, Amplitude p = 0.8205; for sEPSCs, Frequency p = 0.0717, Amplitude p = 0.4319. g-l, Knockdown of MCH abolish the regulatory effects after optogenetic activation of MCH neuronal fibers in the dLS. g, Sample images of tissue following unilateral injections of shRNA AAVs shRNA-scramble or shRNA-Pmch viruses. h&i Quantification of MCH neurons and MCH immunofluorescent signal in the hypothalamic area after shRNA knockdown of pMch. n = 7, 15, 7, in contralateral, scrambled and shRNA-pMch groups, respectively. h, Welch’s AN OVA tests were used, F(2,11.45)=109.8, p < 0.0001. Contralateral vs. shRNA scramble p = 0.9845, Contralateral vs. shRNA-pMch p < 0.0001, shRNA-scramble vs. shRNA-pMch p < 0.0001. I, Welch’s ANOVA tests were used, F(2,11.27)= 18.22, p = 0.0008, Contralateral vs. shRNA scramble p = 0.3187, Contralateral vs. shRNA-pMch p = 0.0133, shRNA-scramble vs. shRNA-pMch p = 0.0016. j, Quantification of MCH immunofluorescent signal in cells with expression of C1V1-mCherry. n = 8,6, in Scrambled shRNA and shRNA-pMch groups. Unpaired Student’s t-test, p = 0.0266 k-l, Impact of MCH shRNA AAV (or control scrambled shRNAs) on the modulatory effect of post-optogenetic stimulation of sEPSCs, showing representative traces (k) and pooled data (l). n = 12, 14 for l in shRNA-scramble and shRNA-pMch groups, respectively. Paired Student’s t-tests were used, shRNA-scramble: Control vs. post 20 Hz, Frequency p = 0.003531; Amplitude p = 0.001719; shRNA-pMch; Control vs. post 20 Hz, Frequency, p = 0.0.998361, Amplitude, p = 0.95653. Data are mean ± SEM; * p < 0.05, ** p < 0.01, ***p < 0.001.

Source data

Extended Data Fig. 4 MCH facilitates hippocampal excitatory inputs to dLS neurons.

a, Sample images of retrograde labeled CA2/3 neurons after microfluorescent RetroBeads were injected into the dLS region. b, Retrograde labelled LHA neurons (green) with immunoreactivity for MCH (blue). c, Quantification of mRNA of MCHR1 and vGAT in the dorsal hippocampal CA3 region. n = 7, unpaired Student’s t-test, p < 0.0001. d, Sample oEPSCs evoked by paired-pulse protocols at different intervals before (control) and after the application of MCH. e&f, Quantification of oEPSCs originating from hippocampus in dLS neurons before and after the application of MCH. Data are mean ± SEM; n = 7 cells/3 animals for e & f. Paired two-tailed Student’s t-tests were used: e, p = 0.0031; f, p = 0.0033 ** p < 0.01. g&h, Sample images of injection sites. i&j Endogenous MCH facilitates HP-to-dLS EPSCs via MCHR1 signaling. i, Representative superimposed traces of optogenetically-induced EPSCs in the dLS, in the presence of TC-MCH7c (10 µM) or SNAP94847 (20 µM), pre- (black trace) and post- (red trace) prolonged 590 nm light activation on C1V1-positive MCH axons. j, Pooled single oEPSC amplitude data. Paired two-tailed Student’s t-tests were used but no statistical differences were found, p = 0.0907 in TC-MCH7c and p = 0.6607 in SNAP94847.

Source data

Extended Data Fig. 5 Endogenous MCH increases GABA tone in the dLS.

a, Pooled data for amplitude. n = 11 cells/9 animals and 14/9 for 5 Hz and 20 Hz, respectively. Paired two-tailed Student’s t-tests were used: 5 Hz, p = 0.2905 and 20 Hz, p = 0.132. AAV-DIO-ChR2-eYFP was injected into the lateral hypothalamic region of MCH-Cre mice, sIPSCs (isolated pharmacologically in the presence of CNQX and APV) were recorded in dLS neurons. MCH fibers were optogenetically activated using two different paradigms (5 Hz 5 second x 5 or 20 Hz 2 min). sIPSCs were recorded before optogenetic stimulation (Control) and 2–3 min after optogenetic stimulation (indicated as 5 Hz or 20 Hz). b, Sample traces showing the impact of optogenetic stimulation of MCH fibers on sIPSCs in dLS neurons, as described above, in the presence of Tc-MCH7c. c, Pooled data show that in the presence of TC-MCH7c, the facilitatory effects induced post-prolonged optogenetic stimulation of MCH fibers were abolished, suggesting the involvement of MCH signaling. n = 14/7 and 13/7 for 5 Hz and 20 Hz, respectively. Paired two-tailed Student’s t-tests were used: 5 Hz, p = 0.4489 and 20 Hz, p = 0.9195. d, Sample traces of sIPSCs in the absence (Control) or presence of MCH. f, Pooled data for e. n = 8/4 neurons/animals. Paired two-tailed Student’s t-tests were used, p = 0.0198. f-h MCH suppresses collateral transmission in the dLS. g, Diagram of experimental setup and sample images showing AAV-ChR2-eYFP infection in the dLS. Whole-cell patch clamp recordings were made from cells without ChR2-eYFP infection in the dLS; fibers and neuronal cell bodies expressing ChR2 were activated by 470 nm LED light (each pulse at 1 ms duration at 1, 5 and 10 Hz); collateral optogenetically-induced IPSCs were recorded. g, Sample traces showing the GABAAR-mediated inhibitory collateral synaptic transmission before (black traces) and after (green traces) administration of exogenous MCH (600 nM) through perfusion chamber. h, Pooled data from b. n = 12 cells/5 animals. Paired two-tailed Student’s t-tests were used, p = 0.0006. i, GABAB receptor-mediated inhibition suppresses spontaneous firing of dLS neurons. Sample traces showing that GABAB receptor-mediated inhibitory synaptic transmission is sufficient to suppress dLS spontaneous action potential firing. AAV-ChR2-eYFP was injected into the dLS. Whole-cell patch clamp recordings under current clamp mode were made from cells without ChR2-eYFP infection in the dLS; fibers and neuronal cell bodies expressing ChR2 were activated by 470 nm LED light (each pulse was 1 ms duration delivered at 1, 5 or 10 Hz). In the presence of GABAAR blocker PTX, optogenetic stimulation delivered at varying frequencies (1 Hz, 5 Hz and 10 Hz) to activate dLS neurons suppressed spontaneous firing. Higher frequencies produce prolonged hyperpolarization in dLS neurons. In the presence of CGP55845, a GABABR blocker, the suppression of firing and hyperpolarization of membrane potential were both blocked, suggesting the involvement of GABAB receptor signaling. Data are mean ± SEM * p < 0.05, ** p < 0.01, ***p < 0.001.

Source data

Extended Data Fig. 6 GABABR-mediated FFI functions as a low-pass filter.

The FFI acts as a gate to block high-frequency-encoded information generated in the hippocampus from passing through to downstream targets through the dLS neural network. a, Sample traces of dLS neuron firing in response to optogenetically-evoked oEPSCs from the dHP. b&c, Pooled data from a. d, Sample traces of neurons firing spontaneous APs interferes with 1 Hz dHP excitatory inputs induced AP generations. n = 38 cells/ 9 animals. Data are presented as mean ± SEM.

Source data

Extended Data Fig. 7 Hippocampal inputs induced postsynaptic action potentials.

a, Sample traces of neuronal oAPs induced by 1 Hz dHP excitatory inputs. b, Superimposed postsynaptic potentials (upper panel) and postsynaptic currents (lower panel) under different conditions (indicated by different color as indicted in a). c, Diagram of possible wiring of HP-dLS-LHA neurocircuitry.

Source data

Extended Data Fig. 8 Behavioral impacts of MCH signaling in the dLS.

a, Light microscope image of sample coronal brain slices illustrating the guide and internal cannula tracks. Tracks of injector tips are labeled by stars. b-e, RotaRod Motor learning task data. b&d, pooled data for the first 10 trials, in which speed accelerated from 4–40 revolutions/min in 300 s. Two-way ANOVA test was used. b, animals x treatment p = 0.9602, treatment p = 0.464. d, animals x treatment p = 0.9509, treatment p = 0.5329. c&e, Pooled data from the last 2 trials, in which speed accelerated from 4–40 revolutions/min in150s. Two-way ANOVA test was used. c, animals x treatment p = 0.1222, treatment p = 0.2944. e, animals x treatment p = 0.0776, treatment p = 0.4157. f-k, Performance on the locomotor open field test with patterned wall. n = 6,7,7 in control, MCH and Tc-MCH7c groups. One-way ANOVA test with posthoc Tukey’s multiple comparison were used for all analysis. f, Schematic of the testing arena. g, Distance traveled under each treatment condition. F(2,17)=0.1059,p = 0.9001. h, Time spent in center time under each treatment condition. Treatment F(2,17)=2.124, p = 0.1501. i,Time spent active under each treatment condition. Treatment F(2,17)=0.2081, p = 0.8142. j, Movement velocity under each treatment condition. treatment F(2,17)=0.1037, p = 0.902. k, Time spent near patterned walls under each treatment condition. Treatment F(2,17)=4.522, *p = 0.0.0266; Control vs. Ala17]MCH, p = 0.0477, [Ala17]MCH vs. TC-MCH7c, p = 0.0482. l-n Social recognition behavior with or without MCHR1 agonist. Metric=time (%) spent in the side of each chamber (Chamber 1: familiar mouse; Chamber 2: novel mouse). m, Preference for chamber 1: Two-way ANOVA test was used for statistics, F(1,11)=8.459, p = 0.0142. n, Data showing a shifted preference, n = 7 and 6 in control and MCH groups, respectively. Unpaired Two-tailed Student’s t-test was used: p = 0.04. o&p, Nesting behavioral analyses. Metric = arbitrary nesting scores. o, Two-way ANOVA test with posthoc Tukey’s multiple comparison were used for statistics, treatment x day F(6,44)=0.7998, p = 0.5754, treatment F(1.985, 43.68)=67.51, p < 0.0001; Control vs. [Ala17]MCH p = 0.0106, Control vs. TC-MCH7c p < 0.0001, [Ala17]MCH vs. TC-MCH7c, p < 0.0001. p, Average nesting score over 4 days. F(2,19)=30.36, p < 0.0001. Posthoc Tukey’s multiple comparison, Control vs. MCH: p=0.0096; Control vs. TC-MCH7c, p < 0.0001; MCH vs. TC-MCH7c, p < 0.0001. Data are presented as mean ± SEM, * p < 0.05; ** p < 0.01, *** p < 0.001.

Extended Data Fig. 9 Summary of inferred mechanism through which MCH regulates hippocampo-septal neurocircuits.

a, Diagram of hippocampo-septal circuit, including the role of MCH releasing neurons (green). b, Diagram showing presynaptic regulation of both excitatory and inhibitory synaptic inputs onto a dLS neuron. Synaptic release of GABA activates both GABAA and GABAB receptors. c, Downstream impact of MCH release on dLS neurons. Left: in the absence of MCH, strong collateral inhibition limits the hippocampal-dLS excitatory synaptic inputs induced dLS neuronal firing. Right: In the presence of MCH, dLS neurons are silenced by MCH-induced enhancement of GABAergic inhibition. This in turn reduces the collateral inhibition. However, MCH also enhances dHP-dLS excitatory synaptic inputs, thus enabling the postsynaptic dLS neuron to fire in high fidelity with HP neuronal firing. d, Plot depicts that MCH enhances dLS neuronal firing fidelity in response to HP inputs and thus increased the dynamic range of firing capacity.

Supplementary information

Source data

Source Data Fig. 1

Statistical source data

Source Data Fig. 2

Statistical source data

Source Data Fig. 3

Statistical source data

Source Data Fig. 4

Statistical source data

Source Data Fig. 5

Statistical source data

Source Data Fig. 6

Statistical source data

Source Data Extended Data Fig. 1

Statistical source data

Source Data Extended Data Fig. 2

Statistical source data

Source Data Extended Data Fig. 3

Statistical source data

Source Data Extended Data Fig. 4

Statistical source data

Source Data Extended Data Fig. 5

Statistical source data

Source Data Extended Data Fig. 6

Statistical source data

Sourec Data Extended Data Fig. 7

Statistical source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liu, JJ., Tsien, R.W. & Pang, Z.P. Hypothalamic melanin-concentrating hormone regulates hippocampus-dorsolateral septum activity. Nat Neurosci 25, 61–71 (2022). https://doi.org/10.1038/s41593-021-00984-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-021-00984-5

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