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:

Molecular interrogation of hypothalamic organization reveals distinct dopamine neuronal subtypes

Abstract

The hypothalamus contains the highest diversity of neurons in the brain. Many of these neurons can co-release neurotransmitters and neuropeptides in a use-dependent manner. Investigators have hitherto relied on candidate protein-based tools to correlate behavioral, endocrine and gender traits with hypothalamic neuron identity. Here we map neuronal identities in the hypothalamus by single-cell RNA sequencing. We distinguished 62 neuronal subtypes producing glutamatergic, dopaminergic or GABAergic markers for synaptic neurotransmission and harboring the ability to engage in task-dependent neurotransmitter switching. We identified dopamine neurons that uniquely coexpress the Onecut3 and Nmur2 genes, and placed these in the periventricular nucleus with many synaptic afferents arising from neuromedin S+ neurons of the suprachiasmatic nucleus. These neuroendocrine dopamine cells may contribute to the dopaminergic inhibition of prolactin secretion diurnally, as their neuromedin S+ inputs originate from neurons expressing Per2 and Per3 and their tyrosine hydroxylase phosphorylation is regulated in a circadian fashion. Overall, our catalog of neuronal subclasses provides new understanding of hypothalamic organization and function.

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

Figure 1: Cell-type diversity in the mouse hypothalamus.
Figure 2: Hierarchical clustering of hypothalamic neuron subtypes.
Figure 3: Neurotransmitter phenotypes in hypothalamic neurons.
Figure 4: Neuropeptide associations to individual hypothalamic neuronal subtypes.
Figure 5: Molecular interrogation of dopamine neurons defines a onecut-3-expressing periventricular subtype.
Figure 6: Efferent projections of periventricular onecut-3+ dopamine neurons.
Figure 7: Suprachiasmatic origin for neuromedin S inputs to onecut-3+ Nmur2+ A14 dopamine neurons.
Figure 8: Periventricular onecut-3+ dopamine neurons respond to neuromedin S produced during light periods.

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

References

  1. Du Vigneaud, V. Hormones of the posterior pituitary gland: oxytocin and vasopressin. Harvey Lect. 50, 1–26 (1954-1955).

    PubMed  Google Scholar 

  2. Clarke, I.J. Hypothalamus as an endocrine organ. Compr. Physiol. 5, 217–253 (2015).

    CAS  PubMed  Google Scholar 

  3. Lantos, T.A., Görcs, T.J. & Palkovits, M. Immunohistochemical mapping of neuropeptides in the premamillary region of the hypothalamus in rats. Brain Res. Brain Res. Rev. 20, 209–249 (1995).

    CAS  PubMed  Google Scholar 

  4. Swanson, L.W. & Kuypers, H.G. The paraventricular nucleus of the hypothalamus: cytoarchitectonic subdivisions and organization of projections to the pituitary, dorsal vagal complex, and spinal cord as demonstrated by retrograde fluorescence double-labeling methods. J. Comp. Neurol. 194, 555–570 (1980).

    CAS  PubMed  Google Scholar 

  5. Dahlström, A. & Fuxe, K. Evidence for existence of monoamine-containing neurons in central nervous system. I. Demonstration of monoamines in cell bodies of brain stem neurons. Acta Physiol. Scand. Suppl. 62 (Suppl. 232): 1–55 (1964).

    Google Scholar 

  6. Björklund, A., Moore, R.Y., Nobin, A. & Stenevi, U. The organization of tubero-hypophyseal and reticulo-infundibular catecholamine neuron systems in the rat brain. Brain Res. 51, 171–191 (1973).

    PubMed  Google Scholar 

  7. Horvath, T.L. Suprachiasmatic efferents avoid phenestrated capillaries but innervate neuroendocrine cells, including those producing dopamine. Endocrinology 138, 1312–1320 (1997).

    CAS  PubMed  Google Scholar 

  8. Abizaid, A., Horvath, B., Keefe, D.L., Leranth, C. & Horvath, T.L. Direct visual and circadian pathways target neuroendocrine cells in primates. Eur. J. Neurosci. 20, 2767–2776 (2004).

    PubMed  Google Scholar 

  9. Zeisel, A. et al. Brain structure. Cell types in the mouse cortex and hippocampus revealed by single-cell RNA-seq. Science 347, 1138–1142 (2015).

    CAS  PubMed  Google Scholar 

  10. Scott, N., Prigge, M., Yizhar, O. & Kimchi, T. A sexually dimorphic hypothalamic circuit controls maternal care and oxytocin secretion. Nature 525, 519–522 (2015).

    CAS  PubMed  Google Scholar 

  11. Yang, C.F. et al. Sexually dimorphic neurons in the ventromedial hypothalamus govern mating in both sexes and aggression in males. Cell 153, 896–909 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Xu, X. et al. Modular genetic control of sexually dimorphic behaviors. Cell 148, 596–607 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Yang, X. et al. Tissue-specific expression and regulation of sexually dimorphic genes in mice. Genome Res. 16, 995–1004 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Dulcis, D., Jamshidi, P., Leutgeb, S. & Spitzer, N.C. Neurotransmitter switching in the adult brain regulates behavior. Science 340, 449–453 (2013).

    CAS  PubMed  Google Scholar 

  15. Meister, B., Cortés, R., Villar, M.J., Schalling, M. & Hökfelt, T. Peptides and transmitter enzymes in hypothalamic magnocellular neurons after administration of hyperosmotic stimuli: comparison between messenger RNA and peptide/protein levels. Cell Tissue Res. 260, 279–297 (1990).

    CAS  PubMed  Google Scholar 

  16. Pinto, S. et al. Rapid rewiring of arcuate nucleus feeding circuits by leptin. Science 304, 110–115 (2004).

    CAS  PubMed  Google Scholar 

  17. Cristino, L. et al. Obesity-driven synaptic remodeling affects endocannabinoid control of orexinergic neurons. Proc. Natl. Acad. Sci. USA 110, E2229–E2238 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Romanov, R.A. et al. A secretagogin locus of the mammalian hypothalamus controls stress hormone release. EMBO J. 34, 36–54 (2015).

    CAS  PubMed  Google Scholar 

  19. Henry, F.E., Sugino, K., Tozer, A., Branco, T. & Sternson, S.M. Cell type-specific transcriptomics of hypothalamic energy-sensing neuron responses to weight-loss. Elife 4, 09800 (2015).

    Google Scholar 

  20. Krashes, M.J. et al. An excitatory paraventricular nucleus to AgRP neuron circuit that drives hunger. Nature 507, 238–242 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Everitt, B.J., Hökfelt, T., Wu, J.Y. & Goldstein, M. Coexistence of tyrosine hydroxylase-like and gamma-aminobutyric acid-like immunoreactivities in neurons of the arcuate nucleus. Neuroendocrinology 39, 189–191 (1984).

    CAS  PubMed  Google Scholar 

  22. Macosko, E.Z. et al. Highly parallel genome-wide expression profiling of individual cells using nanoliter droplets. Cell 161, 1202–1214 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Lee, S.Y., Földy, C., Szabadics, J. & Soltesz, I. Cell-type-specific CCK2 receptor signaling underlies the cholecystokinin-mediated selective excitation of hippocampal parvalbumin-positive fast-spiking basket cells. J. Neurosci. 31, 10993–11002 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Moriya, R. et al. RFamide peptide QRFP43 causes obesity with hyperphagia and reduced thermogenesis in mice. Endocrinology 147, 2916–2922 (2006).

    CAS  PubMed  Google Scholar 

  25. Zagorácz, O. et al. Effects of direct QRFP-26 administration into the medial hypothalamic area on food intake in rats. Brain Res. Bull. 118, 58–64 (2015).

    PubMed  Google Scholar 

  26. Tekin, I., Roskoski, R. Jr., Carkaci-Salli, N. & Vrana, K.E. Complex molecular regulation of tyrosine hydroxylase. J. Neural Transm. 121, 1451–1481 (2014).

    CAS  PubMed  Google Scholar 

  27. Hökfelt, T., Martensson, R., Bjorklund, A., Kleinau, S. & Goldstein, M. Distribution maps of tyrosine-hydroxylase-immunoreactive neurons in the rat brain. in Handbook of Chemical Neuroanatomy (eds. Bjorklund, A. & Hökfelt, T.) 277–379 (Elsevier, 1984).

  28. Meister, B. & Elde, R. Dopamine transporter mRNA in neurons of the rat hypothalamus. Neuroendocrinology 58, 388–395 (1993).

    CAS  PubMed  Google Scholar 

  29. Ben-Jonathan, N. & Hnasko, R. Dopamine as a prolactin (PRL) inhibitor. Endocr. Rev. 22, 724–763 (2001).

    CAS  PubMed  Google Scholar 

  30. Stagkourakis, S., Kim, H., Lyons, D.J. & Broberger, C. Dopamine autoreceptor regulation of a hypothalamic dopaminergic network. Cell Rep. 15, 735–747 (2016]).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Goudreau, J.L., Lindley, S.E., Lookingland, K.J. & Moore, K.E. Evidence that hypothalamic periventricular dopamine neurons innervate the intermediate lobe of the rat pituitary. Neuroendocrinology 56, 100–105 (1992).

    CAS  PubMed  Google Scholar 

  32. Mori, K. et al. Identification of neuromedin S and its possible role in the mammalian circadian oscillator system. EMBO J. 24, 325–335 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Zhang, L. et al. A PERIOD3 variant causes a circadian phenotype and is associated with a seasonal mood trait. Proc. Natl. Acad. Sci. USA 113, E1536–E1544 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Vitaterna, M.H. et al. Differential regulation of mammalian period genes and circadian rhythmicity by cryptochromes 1 and 2. Proc. Natl. Acad. Sci. USA 96, 12114–12119 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Hölttä, M. et al. An integrated workflow for multiplex CSF proteomics and peptidomics-identification of candidate cerebrospinal fluid biomarkers of Alzheimer's disease. J. Proteome Res. 14, 654–663 (2015).

    PubMed  Google Scholar 

  36. Irani, D.N. Properties and composition of normal cerebrospinal fluid. in Cerebrospinal Fluid in Clinical Practice Ch. 10, 69–89 (W.B. Saunders, Philadelphia, 2009).

  37. Bro, S. & Haycock, J.W. Visual cortex (translation of Ramón y Cajal, S.) translated from L'Encorce Cérébrale Suivant les Régions L'Ecorce Visuelle, Ch. 25 in Histologie du Système Nerveux de l'Homme et des Vertébrés, 1911. Behav. Biol. 21, 508–528 (1977).

    CAS  PubMed  Google Scholar 

  38. Ascoli, G.A. et al. Petilla terminology: nomenclature of features of GABAergic interneurons of the cerebral cortex. Nat. Rev. Neurosci. 9, 557–568 (2008).

    CAS  PubMed  Google Scholar 

  39. Zivraj, K.H. et al. Subcellular profiling reveals distinct and developmentally regulated repertoire of growth cone mRNAs. J. Neurosci. 30, 15464–15478 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Yulyaningsih, E. et al. Pancreatic polypeptide controls energy homeostasis via Npy6r signaling in the suprachiasmatic nucleus in mice. Cell Metab. 19, 58–72 (2014).

    CAS  PubMed  Google Scholar 

  41. Friedman, J.M. & Halaas, J.L. Leptin and the regulation of body weight in mammals. Nature 395, 763–770 (1998).

    CAS  PubMed  Google Scholar 

  42. Scott, M.M. et al. Leptin targets in the mouse brain. J. Comp. Neurol. 514, 518–532 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Boulland, J.L. et al. Vesicular glutamate and GABA transporters sort to distinct sets of vesicles in a population of presynaptic terminals. Cereb. Cortex 19, 241–248 (2009).

    PubMed  Google Scholar 

  44. Dawson, T.M. & Snyder, S.H. Gases as biological messengers: nitric oxide and carbon monoxide in the brain. J. Neurosci. 14, 5147–5159 (1994).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Jonas, P., Bischofberger, J. & Sandkühler, J. Corelease of two fast neurotransmitters at a central synapse. Science 281, 419–424 (1998).

    CAS  PubMed  Google Scholar 

  46. Ajika, K. & Hökfelt, T. Ultrastructural identification of catecholamine neurones in the hypothalamic periventricular-arcuate nucleus-median eminence complex with special reference to quantitative aspects. Brain Res. 57, 97–117 (1973).

    CAS  PubMed  Google Scholar 

  47. Horvath, T.L. An alternate pathway for visual signal integration into the hypothalamo-pituitary axis: retinorecipient intergeniculate neurons project to various regions of the hypothalamus and innervate neuroendocrine cells including those producing dopamine. J. Neurosci. 18, 1546–1558 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. DeMaria, J.E., Lerant, A.A. & Freeman, M.E. Prolactin activates all three populations of hypothalamic neuroendocrine dopaminergic neurons in ovariectomized rats. Brain Res. 837, 236–241 (1999).

    CAS  PubMed  Google Scholar 

  49. Lee, I.T. et al. Neuromedin s-producing neurons act as essential pacemakers in the suprachiasmatic nucleus to couple clock neurons and dictate circadian rhythms. Neuron 85, 1086–1102 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Sellix, M.T. et al. Anatomical and functional characterization of clock gene expression in neuroendocrine dopaminergic neurons. Am. J. Physiol. Regul. Integr. Comp. Physiol. 290, R1309–R1323 (2006).

    CAS  PubMed  Google Scholar 

  51. Tamamaki, N. et al. Green fluorescent protein expression and colocalization with calretinin, parvalbumin, and somatostatin in the GAD67-GFP knock-in mouse. J. Comp. Neurol. 467, 60–79 (2003).

    CAS  PubMed  Google Scholar 

  52. Máté, Z. et al. Spatiotemporal expression pattern of DsRedT3/CCK gene construct during postnatal development of myenteric plexus in transgenic mice. Cell Tissue Res. 352, 199–206 (2013).

    PubMed  Google Scholar 

  53. Pierreux, C.E., Vanhorenbeeck, V., Jacquemin, P., Lemaigre, F.P. & Rousseau, G.G. The transcription factor hepatocyte nuclear factor-6/Onecut-1 controls the expression of its paralog Onecut-3 in developing mouse endoderm. J. Biol. Chem. 279, 51298–51304 (2004).

    CAS  PubMed  Google Scholar 

  54. Wong, H.C., Sternini, C., Lloyd, K., De Giorgio, R. & Walsh, J.H. Monoclonal antibody to VIP: production, characterization, immunoneutralizing activity, and usefulness in cytochemical staining. Hybridoma 15, 133–139 (1996).

    CAS  PubMed  Google Scholar 

  55. Miyazaki, T., Fukaya, M., Shimizu, H. & Watanabe, M. Subtype switching of vesicular glutamate transporters at parallel fibre-Purkinje cell synapses in developing mouse cerebellum. The European journal of neuroscience 17, 2563–2572 (2003).

    PubMed  Google Scholar 

  56. Mulder, J. et al. Secretagogin is a Ca2+-binding protein identifying prospective extended amygdala neurons in the developing mammalian telencephalon. Eur. J. Neurosci. 31, 2166–2177 (2010).

    PubMed  PubMed Central  Google Scholar 

  57. Weiss, M.L. & Cobbett, P. Intravenous injection of Evans blue labels magnocellular neuroendocrine cells of the rat supraoptic nucleus in situ and after dissociation. Neuroscience 48, 383–395 (1992).

    CAS  PubMed  Google Scholar 

  58. Murray, A.J. et al. Parvalbumin-positive CA1 interneurons are required for spatial working but not for reference memory. Nat. Neurosci. 14, 297–299 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Pollak Dorocic, I. et al. A whole-brain atlas of inputs to serotonergic neurons of the dorsal and median raphe nuclei. Neuron 83, 663–678 (2014).

    CAS  PubMed  Google Scholar 

  60. Susaki, E.A. et al. Whole-brain imaging with single-cell resolution using chemical cocktails and computational analysis. Cell 157, 726–739 (2014).

    CAS  PubMed  Google Scholar 

  61. Paxinos, G. & Franklin, K.B.J. The Mouse Brain in Stereotaxic Coordinates (Academic, San Diego, 2001).

  62. Tomer, R., Ye, L., Hsueh, B. & Deisseroth, K. Advanced CLARITY for rapid and high-resolution imaging of intact tissues. Nat. Protoc. 9, 1682–1697 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Schwanhäusser, B. et al. Global quantification of mammalian gene expression control. Nature 473, 337–342 (2011).

    PubMed  Google Scholar 

  64. Islam, S. et al. Quantitative single-cell RNA-seq with unique molecular identifiers. Nat. Methods 11, 163–166 (2014).

    CAS  PubMed  Google Scholar 

  65. Tsafrir, D. et al. Sorting points into neighborhoods (SPIN): data analysis and visualization by ordering distance matrices. Bioinformatics 21, 2301–2308 (2005).

    CAS  PubMed  Google Scholar 

  66. Marques, S. et al. Oligodendrocyte heterogeneity in the mouse juvenile and adult central nervous system. Science 352, 1326–1329 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Fan, J. et al. Characterizing transcriptional heterogeneity through pathway and gene set overdispersion analysis. Nat. Methods 13, 241–244 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Tasic, B. et al. Adult mouse cortical cell taxonomy revealed by single cell transcriptomics. Nat. Neurosci. 19, 335–346 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Fuzik, J. et al. Integration of electrophysiological recordings with single-cell RNA-seq data identifies neuronal subtypes. Nat. Biotechnol. 34, 175–183 (2016).

    CAS  PubMed  Google Scholar 

  70. Sprong, H. et al. UDP-galactose:ceramide galactosyltransferase is a class I integral membrane protein of the endoplasmic reticulum. J. Biol. Chem. 273, 25880–25888 (1998).

    CAS  PubMed  Google Scholar 

  71. Jahn, O., Tenzer, S. & Werner, H.B. Myelin proteomics: molecular anatomy of an insulating sheath. Mol. Neurobiol. 40, 55–72 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  72. Ebrahimi, M. et al. Astrocyte-expressed FABP7 regulates dendritic morphology and excitatory synaptic function of cortical neurons. Glia 64, 48–62 (2016).

    PubMed  Google Scholar 

  73. Roy, S. The motile cilium in development and disease: emerging new insights. Bioessays 31, 694–699 (2009).

    CAS  PubMed  Google Scholar 

  74. Liu, G., Ma, H., Jiang, L. & Zhao, Y. Allograft inflammatory factor-1 and its immune regulation. Autoimmunity 40, 95–102 (2007).

    CAS  PubMed  Google Scholar 

  75. Ose, A. et al. Functional characterization of mouse organic anion transporting peptide 1a4 in the uptake and efflux of drugs across the blood-brain barrier. Drug Metab. Dispos. 38, 168–176 (2010).

    CAS  PubMed  Google Scholar 

  76. Li, L., Miano, J.M., Cserjesi, P. & Olson, E.N. SM22 alpha, a marker of adult smooth muscle, is expressed in multiple myogenic lineages during embryogenesis. Circ. Res. 78, 188–195 (1996).

    CAS  PubMed  Google Scholar 

  77. Tortoriello, G. et al. Miswiring the brain: Δ9-tetrahydrocannabinol disrupts cortical development by inducing an SCG10/stathmin-2 degradation pathway. EMBO J. 33, 668–685 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  78. Benjamini, Y. & Hochberg, Y. Controlling the false discovery rate – a practical and powerful approach to multiple testing. J. R. Stat. Soc. Series B Stat. Methodol. 57, 289–300 (1995).

    Google Scholar 

  79. van der Maaten, L. & Hinton, G. Visualizing data using t-SNE. J. Mach. Learn. Res. 9, 2579–2605 (2008).

    Google Scholar 

  80. Dobin, A. et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21 (2013).

    CAS  PubMed  Google Scholar 

  81. Schnell, S.A., Staines, W.A. & Wessendorf, M.W. Reduction of lipofuscin-like autofluorescence in fluorescently labeled tissue. J. Histochem. Cytochem. 47, 719–730 (1999).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors thank N.-G. Larsson and L. Olson for providing Dat1-Cre mice for the generation of reporter mice, H. Wong and M. Watanabe for antibodies and K. Meletis for his supervision of viral injections in Dat1-Cre mice. This work was supported by the Swedish Research Council (T. Harkany, T. Hökfelt, S.L., C. Broberger), Hjärnfonden (T. Harkany), the Petrus and Augusta Hedlunds Foundation (T. Harkany), the Novo Nordisk Foundation (T. Harkany, T. Hökfelt, C. Broberger), the National Brain Research Program of Hungary (MTA-SE NAP B, KTIA_NAP_13-2014-0013; A.A.), the European Commission (PAINCAGE grant, T. Harkany, T. Hökfelt), the European Research Council (BRAINCELL; S.L., ENDOSWITCH; C. Broberger and SECRET-CELLS; T. Harkany), intramural funds of the Medical University of Vienna (T. Harkany) and an NIH grant AG051459 (T.L.H.). R.A.R. is an EMBO long-term research fellow (ALTF 596-2014) cofunded by the European Commission FP7 (Marie Curie Actions, EMBOCOFUND2012, GA-2012-600394). A.Z. received support from the Human Frontier Science Program. F.C. is a Research Associate of the Fonds de la Recherche Scientifique-FNRS, Belgium. The single-cell sequencing infrastructure at CeMM was supported by a New Frontiers Research Infrastructure grant from the Austrian Academy of Sciences.

Author information

Authors and Affiliations

Authors

Contributions

T. Harkany and R.A.R. conceived the general framework of this study. T. Harkany, T.L.H., S.L., R.A.R., A.Z., T. Hökfelt, C. Broberger, K.D. designed experiments, T. Harkany, T.L.H., S.L., T. Hökfelt, C. Broberger, K.D., A.A., J.M. and C. Bock senior authors, sponsored research. R.A.R., A.Z., A.H., J.B., F.G., A.A., E.K., R.T., B.H., A.K.C., D.C., M.-D.Z., A.R. and M.F. performed research and analyzed data. H.M., C.S., D.C., Z.M., G.S., F.C., Y.Y., M.U., J.S.B. and P.W. provided unique reagents. R.A.R., A.Z., T.L.H. and T. Harkany wrote the paper. All authors reviewed the manuscript and approved its submission.

Corresponding authors

Correspondence to Tamas L Horvath or Tibor Harkany.

Ethics declarations

Competing interests

T. Harkany declares support from GW Pharmaceuticals on projects unrelated to the focus of this report.

Integrated supplementary information

Supplementary Figure 1 Methodological considerations for and quality control of single-cell RNA-seq

(a) Neuron numbers per cluster in our analysis (blue boxes), and comparison of actual numbers in repeat experiments with statistical probing of random distribution (solid red circles). Note that this comparison excluded sampling or processing-related bias due either to false positive observations or undersampling. (b) Bar plots show the total number of genes detected in individual neuronal subtypes. (c) Likewise, the total number of mRNA molecules passing our filtering criteria (see Online Methods ) were plotted. Grey circles and error bars represent means ± s.e.m. per group in (b,c).

Supplementary Figure 2 Sex and acute stress do not bias neuronal clustering

(a) Numbers of cells from female animals in each cluster (‘observed’) with respect to the range expected by random sampling (‘expected by random ± s.d.’). Clusters on the ordinate follow their listing in Figure 2. The overall frequency of cells of female origin was 30% in the neuronal dataset. The expected means ± s.d. were calculated from the binomial distribution (Bin(N,p) where p = 0.3 and N = the number of cells in each cluster). Clusters show no significant bias for sampling. Clusters #6 and #55 lean towards female and male dominance, respectively. Note that cluster #7 only contains neurons from males. Upper panel: p values calculated using binomial distribution for enrichment of males/females in each cluster. (b) Cluster distribution of cells isolated from animals 6h after acute formalin stress. Solid red circles denote neurons from stress-exposed animals (‘observed’). Blue bars represent the binomial distribution as calculated if distribution was random (‘expected by random ± s.d.’). Upper panel: p values calculated using binomial distribution for enrichment of cells from stress-exposed animals in each cluster. None of the clusters showed stress-related bias.

Supplementary Figure 3 Visualization of hypothalamic neuron subtypes on two-dimensional maps using tSNE

1,194 genes, perplexity = 5, 10 or 20 with 200 principle components. Neurons were color-coded by highest expression of well-known, cluster-defining hypothalamic markers. (a) Distribution of neurons expressing select neuropeptide and neurotransmission-related genes. (b) Distribution of 62 neuronal clusters determined by the BackSpinV2 algorithm. For abbreviations we refer to Figures 1 and 2 of the manuscript.

Supplementary Figure 4 Distribution of 62 neuronal clusters determined by the BackSpinV2 algorithm

Data for each cluster is shown separately on a cumulative tSNE background. Please note that most of the 62 groups are clustered in terms of tSNE plot coordinates while some show lower stringency: 46 of 62 clusters as relatively well separated in the tSNE plot with forming visual clusters with or without outliers; 10 neuronal groups as “satisfactory” clustered with more than one separation core or by forming segregated groups with a relatively large distance between individual neurons. Several clusters do not form segregated groups in tSNE plots: Vglut2 1, Vglut2 3, Vglut2 16, Hmit+/−, GABA 4, GABA 5 because of several possible reasons: a deeper inner heterogeneity of the clusters Vglut2 (all), Hmit+/− and GABA 5, their mixed phenotypes and/or a low number of genes segregating those cells. One needs to remember fundamental algorithmic differences between BackSpin and tSNE: BackSpin can “ignore” genes (which are enriched in other clusters) when splitting the current group. In contrast, tSNE always considers all genes. Thus, tSNE will be more sensitive to carryover of mRNA and contamination by doublets than BackSpin. Individual data points correspond to single cells.

Supplementary Figure 5 Heterogeneity of corticotropin-releasing hormone systems in the mouse hypothalamus

(a,a1) Genetic tracing reveals the distribution of Crh+ neurons concentrated in the bed nucleus of stria terminalis (BST, a)1-3 and paraventricular hypothalamic nucleus (PVH, a1), as well as shows scattered neurons with a history of Crh expression4,5. (b) Taxonomy of Crh+ neurons in the mouse hypothalamus. Note that dual GABA/glutamate phenotypes exist. Cluster numbers are as per Figure 2. (c) Crh, CRH receptors 1,2 (encoded by Crhr1/2 genes) and CRH-binding protein (Crhbp gene) mRNA expression in hypothalamic neuronal subtypes (vertical axes). Expression levels (horizontal axis) were plotted as means of log ± s.e.m. Red and green colors identify GABAergic and glutamatergic clusters (#44 and #45), which express Crh mRNA at levels exceeding 2x s.e.m. Note that significant levels of gene expression in clusters were found only for Crhr2 and Crhbp but not for Crhr1 (*q < 0.05). Crhr1 and Crhr2 were mostly present at low copy numbers in sparse hypothalamic neurons amongst different subtypes. (d) Heat-map representation of genes differentially expressed between GABAergic and glutamatergic Crh+ neurons. Increasing color intensity towards red is proportionate to higher mRNA content. Only p values (Wilcoxon rank-sum test) are shown yet all q values were also < 0.05; n = 10 (GABA) and n = 11 (glutamate) neurons in discrete branches of taxonomy. In GABA+/Crh+ neurons, we observed a predominance of Prkacb (encoding c-AMP-dependent protein kinase subunit B), Amd2 (coding for S-adenosylmethionine decarboxylase 2), Psma7 (encoding proteasome subunit α7), Syt1 (encoding synaptotagmin 1), Crim1 (producing cysteine rich transmembrane BMP regulator 1), Chn1 (coding for chimerin 1), Rgs17 (encoding regulator of G protein signaling 17), Syn2 (producing synapsin 2), Celf2 (coding for Elav-like family member 2) and Sec61a2 (encoding translocon α2 subunit). Glutamatergic neurons were found to express higher levels of Tmem50b (that is, transmembrane protein 50B protein), Tmem176b (encoding transmembrane protein 176B), Pdcl3 (coding for phosducin-like protein 3), Usp31 (encoding ubiquitin-specific peptidase 31), Frg1 (encoding FSHD region geme 1) and Cyb5r1 (producing cytochrome b5 reductase 1). Scale bars = 250 μm (a,a1). Abbreviation: 3V, third ventricle.

References:

1. Hrabovszky, E., Wittmann, G., Turi, G.F., Liposits, Z. & Fekete, C. Hypophysiotropic thyrotropin-releasing hormone and corticotropin-releasing hormone neurons of the rat contain vesicular glutamate transporter-2. Endocrinology 146, 341-347 (2005).

2. Lee, Y. & Davis, M. Role of the hippocampus, the bed nucleus of the stria terminalis, and the amygdala in the excitatory effect of corticotropin-releasing hormone on the acoustic startle reflex. J Neurosci. 17, 6434-6446 (1997).

3. McNally, G.P. & Akil, H. Role of corticotropin-releasing hormone in the amygdala and bed nucleus of the stria terminalis in the behavioral, pain modulatory, and endocrine consequences of opiate withdrawal. Neuroscience 112, 605-617 (2002).

4. Palkovits, M. Stress-induced expression of co-localized neuropeptides in hypothalamic and amygdaloid neurons. European journal of pharmacology 405, 161-166 (2000).

5. Watts, A.G., Sanchez-Watts, G. & Kelly, A.B. Distinct patterns of neuropeptide gene expression in the lateral hypothalamic area and arcuate nucleus are associated with dehydration-induced anorexia. J Neurosci. 19, 6111-6121 (1999).

Supplementary Figure 6 Representative markers for hypothalamic neuronal subtypes and their localization

(a) For each neuronal cluster, the most specific markers were calculated (gene names are on top). To identify the most unique markers for each neuronal cluster, we used power = 0 analysis to identify topmost-expressed unique genes. To force uniqueness, we excluded genes that appear in the list of top 5 markers in other clusters. Since we often observed clusters that are characterized by gene combinations rather than unique global markers, some of the top 5 markers showed low specificity. All genes were found to be statistically significant by the Wilcoxon rank-sum test (q < 0.05) with the exception of L3hypdh (p = 0.04), Prkd1 (p = 0.007) and Ing2 (p = 0.01). The color scale to the right presents values after log transform, which were centered and normalized to mean = 0 and s.d. = 1 for each gene. Saturated colors represent the upper and lower 1% (range 1-99%). (b-c1) Novel neuropeptide identities in the hypothalamus. Hypocretin (Hcrt, b) and galanin (Gal, c)-containing neuronal clusters (#35 and #37, respectively) uniquely co-express mRNAs for pyroglutamylated RFamide peptide (Qrfp; b1) and neuropeptide VF precursor (Npvf, c1). Note that Hcrt+ cluster #36 lacks Qfrp expression. In situ hybridization identifies Qrfp+ or Nvpf+ neurons in the arcuate nucleus (Arc)-lateral hypothalamic area (LHA) and dorsomedial hypothalamus (DMH), respectively. Histochemical data are from the Allen Brain Atlas (www.brain-map.org). Scale bars = 150 μm. mRNA copy numbers were expressed as means ± s.e.m. (log2(mRNA copies + 1); power = 1). *q < 0.05 (Wilcoxon rank-sum test corrected for multiple testing).

Supplementary Figure 7 Histochemical analysis of novel neuronal markers and A14 neurons in the mouse hypothalamus

(a) Novel markers for hierarchical junctions in the hypothalamic diagram (Figure 2). From left to right: oxytocin and Arg-vasopressin (Avp) in the magnocellular paraventricular nucleus of the hypothalamus (PVN), ubiquitin-specific peptidase 48 (USP48), ADP-ribosylation factor guanine nucleotide-exchange factor 1 (AARFGEF1), kinesin family member 5A (KIF5A), as well as dopamine transporter (DAT) expression at the median eminence. (b) Morphology of A14 periventricular dopamine neurons across the mouse hypothalamus. Phosphorylated-TH and onecut-3 co-existence was taken as positive cell identification (arrowheads). Numbers denote anterior-posterior coordinates relative to Bregma. (c) Quantitative immunofluorescence microscopy reveals an inverse relationship between the intensities of GFP and phospho-Ser40-TH (p-Ser40-TH) immunoreactivities for periventricular dopamine neurons in Th-GFP reporter mice. ***p < 0.001 between the groups indicated. Bracketed numbers denote group sizes. Data in box plots represent medians and 10th, 25th, 70th and 90th percentiles. (d,d1) Representative confocal micrographs of coronal single optical sections from the periventricular region of Dat1-tdTomato mice at select anterior-posterior subdivisions stained for onecut-3 and TH. Endogenous tdTomato signal was not amplified. Venn diagrams show the average number of Dat1-tdTomato (red), onecut-3 (green) and TH (blue) immunoreactive (ir) neurons ± s.e.m. per optical slice (n = 6 animals). The relative number of immunoreactive somata compared to the total number of cells is denoted as percentages. Overlap represents co-localization. Note the high degree of co-localization for the tdTomato signal with TH and onecut-3 immunoreactivities. Also note a cluster of tdTomato cells at the retrochiasmatic region that are onecut-3+ but likely lack appreciable TH expression. All encircled cells in the photomicrographs were color-coded according to the cell’s fluorescence labeling. Bregma levels were indicated at the bottom-left. (e) Single-plane views of CLARITY-reconstructed mouse hypothalami stained for TH and focusing on the A14 cell group (semi-transparent overlay). Images were taken at a semi-horizontal plane, with the lower focusing on A14 neurons (see also Supplementary Movie 3). Abbreviations: 3V, third ventricle; A13, zona incerta; Arc, arcuate nucleus; PeVN, periventricular nucleus; SCN, suprachiasmatic nucleus. Scale bars = 150 μm (a [junctions 61/49, 23 inset, 28, 27]), 250 μm (a [junction 23, 42],e), 50 μm (d,d1), 45 μm (a [junction 42 inset],b).

Supplementary Figure 8 Neuronal heterogeneity in the suprachiasmatic nucleus

(a) In situ hybridization histochemistry showing the expression of vasoactive intestinal polypeptide (Vip) mRNA in the suprachiasmatic nucleus (SCN). (b) Likewise, gastrin-releasing peptide (Grp) mRNA was selectively detected in cluster #40 and histochemically localized to the SCN. (c) VIP/GRP co-existence in boutons within the SCN and terminals in the periventricular nucleus (PeVN). (c1) GABAergic neuronal components (green), including a subset of neurons and axonal pathways co-expressing (arrowheads) a DsRed construct under the control of the Cck promoter (blue) are shown in a dual-reporter mouse, highlighting the abundance of GABA neurons in the SCN. (d) Top: Neurotransmitter heterogeneity in the SCN. Overlapping parts of the Venn diagram denote dual GABA/glutamate neurons. ND: non-defined. Bottom: Molecular heterogeneity of neuromedin S-containing neurons. Note the abundance of clock genes, Vip and Cck. (e) Differential Per3 mRNA expression in neuronal subclasses of the hypothalamus. Note highest levels of Per3 mRNA assignment to clusters #40 and #41. Red color denotes expression levels of > 2x s.e.m. from zero. Clusters were ordered according to Figure 2. mRNA copy numbers were expressed as means ± s.e.m. (log2(mRNA copies + 1)), power = 0. (f) Period gene 2 (Per2) mRNA localization by in situ hybridization in the SCN. (g) Neuromedin S (NMS) immunoreactivity around the third ventricle (3V) including synaptic boutons co-stained for the presynaptic protein VAMP2. The existence of particular neuronal subclasses was confirmed by in situ hybridization data from the open source Allen Brain Atlas database (www.brain-map.org). Abbreviations: 3V, third ventricle. Scale bars = 150 μm (a,b,c left,c1), 70 μm (f left), 20 μm (c right,f right).

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–8 and Supplementary Table 1 (PDF 2300 kb)

Supplementary Methods Checklist (PDF 434 kb)

Supplementary Table 2

Gene expression in neurons in hypothalamic clusters #1-#62 (XLSX 1165 kb)

Supplementary Table 3: Marker sets used to define junction points during dendrogram construction

This table shows the top markers that separate each junction of the dendrogram in Figure 2a. For each junction, we searched for genes that best separate the two sides of the junction. The average of the log2(x+1) expression (left) and the fraction of positive cells in each group (expression > 0; right) were calculated, and genes were ranked by their difference. The table shows the top 50 genes found to be specific for each side of particular junctions (left and right) as mentioned above each sub-table. In addition to the score, the p value was calculated using binomial distribution against the null hypothesis that the positive cells are distributed randomly between the groups. q values correspond to p values corrected for multiple testing since each gene was tested for all 61 junctions. (XLSX 727 kb)

Supplementary Table 4: Expression of neuropeptide-coding genes in hypothalamic clusters #1-#62

Increasing color depth from white toward red was used to visualize genes expressed by individual clusters at distinct levels of statistical significance (q values are shown). (XLSX 18 kb)

Supplementary Table 5: P values for neuron-specific genes (Wilcoxon rank-sum test) expressed by hypothalamic neuronal clusters #1-#62

Increasing color depth from white toward red was used to visualize genes at distinct levels of statistical significance. (XLSX 5722 kb)

Supplementary Table 6: Q values for neuron-specific genes (Wilcoxon rank-sum test corrected for multiple testing using horizontal correction) expressed by hypothalamic neuronal clusters #1-#62

Increasing color depth from white toward red was used to visualize genes at distinct levels of statistical significance. (XLSX 3151 kb)

Three-dimensional reconstruction of the suprachiasmatic nucleus-paraventricular hypothalamic nucleus region by light-sheet microscopy.

Red and green colors correspond to phospho-Ser40-TH and onecut-3 immunosignals respectively. Data in rendered form are shown in Fig. 5f. Imaging was performed on a Zeiss Lightsheet Z.1 microscope at 5x primary magnification. (AVI 61385 kb)

Three-dimensional reconstruction of the retrochiasmatic-arcuate nucleus rostral-caudal extent by light-sheet microscopy.

Red and green colors correspond to phospho-Ser40-TH and onecut-3 immunosignals respectively. Data in rendered form are shown in Fig. 5f. Imaging was performed on a Zeiss Lightsheet Z.1 microscope at 5x primary magnification. (AVI 15287 kb)

Three-dimensional reconstruction of TH+ neurons of the hypothalamus in the intact adult mouse forebrain by CLARITY.

TH+ cells were visualized using TH immunostaining (see Expanded Methods for details). The size of the bounding box in the movie (i.e. zoomed in volume) is 4.624 mm × 1.910 mm. (WMV 32860 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Romanov, R., Zeisel, A., Bakker, J. et al. Molecular interrogation of hypothalamic organization reveals distinct dopamine neuronal subtypes. Nat Neurosci 20, 176–188 (2017). https://doi.org/10.1038/nn.4462

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.4462

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