Skip to main content

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

  • Article
  • Published:

A genetically encoded sensor measures temporal oxytocin release from different neuronal compartments

Abstract

Oxytocin (OT), a peptide hormone and neuromodulator, is involved in diverse physiological and pathophysiological processes in the central nervous system and the periphery. However, the regulation and functional sequences of spatial OT release in the brain remain poorly understood. We describe a genetically encoded G-protein-coupled receptor activation-based (GRAB) OT sensor called GRABOT1.0. In contrast to previous methods, GRABOT1.0 enables imaging of OT release ex vivo and in vivo with suitable sensitivity, specificity and spatiotemporal resolution. Using this sensor, we visualize stimulation-induced OT release from specific neuronal compartments in mouse brain slices and discover that N-type calcium channels predominantly mediate axonal OT release, whereas L-type calcium channels mediate somatodendritic OT release. We identify differences in the fusion machinery of OT release for axon terminals versus somata and dendrites. Finally, we measure OT dynamics in various brain regions in mice during male courtship behavior. Thus, GRABOT1.0 provides insights into the role of compartmental OT release in physiological and behavioral functions.

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: Development of GRABOT sensors and activity-dependent OT release in brain slices.
Fig. 2: Probing the spatial and temporal dynamics of OT release in the axonal and somatodendritic compartments.
Fig. 3: N-type and L-type VGCCs support axonal and somatodendritic OT release, respectively.
Fig. 4: SNARE proteins have distinct roles in axonal and somatodendritic OT release.
Fig. 5: OT1.0 can be used to monitor OT release in vivo during male mating.
Fig. 6: Model showing the molecular basis for axonal versus somatodendritic OT release.

Similar content being viewed by others

Data availability

Plasmids for expressing OT1.0 and OT1.0mut used in this study have been deposited to Addgene (185384–185389; https://www.addgene.org/Yulong_Li/). Source data are provided with this paper.

Code availability

The custom-written ImageJ macro, MATLAB code and Arduino programs used in this study are available at https://github.com/OTsensor/2-photon_imaging_analysis, https://github.com/OTsensor/csv2xls and https://github.com/OTsensor/elec_stim_arduino.

References

  1. Jahr, C. E. & Nicoll, R. A. Dendrodendritic inhibition: demonstration with intracellular recording. Science 207, 1473–1475 (1980).

    Article  CAS  PubMed  Google Scholar 

  2. Cheramy, A., Leviel, V. & Glowinski, J. Dendritic release of dopamine in the substantia nigra. Nature 289, 537–542 (1981).

    Article  CAS  PubMed  Google Scholar 

  3. Castel, M., Morris, J. & Belenky, M. Non-synaptic and dendritic exocytosis from dense-cored vesicles in the suprachiasmatic nucleus. Neuroreport 7, 543–547 (1996).

    Article  CAS  PubMed  Google Scholar 

  4. Simmons, M. L., Terman, G. W., Gibbs, S. M. & Chavkin, C. L-type calcium channels mediate dynorphin neuropeptide release from dendrites but not axons of hippocampal granule cells. Neuron 14, 1265–1272 (1995).

    Article  CAS  PubMed  Google Scholar 

  5. Ludwig, M. et al. Intracellular calcium stores regulate activity-dependent neuropeptide release from dendrites. Nature 418, 85–89 (2002).

    Article  CAS  PubMed  Google Scholar 

  6. Ludwig, M. & Leng, G. Dendritic peptide release and peptide-dependent behaviours. Nat. Rev. Neurosci. 7, 126–136 (2006).

    Article  CAS  PubMed  Google Scholar 

  7. Kennedy, M. J. & Ehlers, M. D. Mechanisms and function of dendritic exocytosis. Neuron 69, 856–875 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Pow, D. V. & Morris, J. F. Dendrites of hypothalamic magnocellular neurons release neurohypophysial peptides by exocytosis. Neuroscience 32, 435–439 (1989).

    Article  CAS  PubMed  Google Scholar 

  9. Ludwig, M. Dendritic release of vasopressin and oxytocin. J. Neuroendocrinol. 10, 881–895 (1998).

    Article  CAS  PubMed  Google Scholar 

  10. Rhodes, C. H., Morrell, J. I. & Pfaff, D. W. Immunohistochemical analysis of magnocellular elements in rat hypothalamus: distribution and numbers of cells containing neurophysin, oxytocin, and vasopressin. J. Comp. Neurol. 198, 45–64 (1981).

    Article  CAS  PubMed  Google Scholar 

  11. Gimpl, G. & Fahrenholz, F. The oxytocin receptor system: structure, function, and regulation. Physiol. Rev. 81, 629–683 (2001).

    Article  CAS  PubMed  Google Scholar 

  12. Cunningham, E. T. Jr. & Sawchenko, P. E. Reflex control of magnocellular vasopressin and oxytocin secretion. Trends Neurosci. 14, 406–411 (1991).

    Article  CAS  PubMed  Google Scholar 

  13. Jurek, B. & Neumann, I. D. The oxytocin receptor: from intracellular signaling to behavior. Physiol. Rev. 98, 1805–1908 (2018).

    Article  CAS  PubMed  Google Scholar 

  14. Ludwig, M. & Pittman, Q. J. Talking back: dendritic neurotransmitter release. Trends Neurosci. 26, 255–261 (2003).

    Article  CAS  PubMed  Google Scholar 

  15. Ferguson, J. N. et al. Social amnesia in mice lacking the oxytocin gene. Nat. Genet. 25, 284–288 (2000).

    Article  CAS  PubMed  Google Scholar 

  16. Andari, E. et al. Promoting social behavior with oxytocin in high-functioning autism spectrum disorders. Proc. Natl Acad. Sci. USA 107, 4389–4394 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Nishioka, T., Anselmo-Franci, J. A., Li, P., Callahan, M. F. & Morris, M. Stress increases oxytocin release within the hypothalamic paraventricular nucleus. Brain Res. 781, 57–61 (1998).

    Article  CAS  PubMed  Google Scholar 

  18. Rosenfeld, A. J., Lieberman, J. A. & Jarskog, L. F. Oxytocin, dopamine, and the amygdala: a neurofunctional model of social cognitive deficits in schizophrenia. Schizophr. Bull. 37, 1077–1087 (2011).

    Article  PubMed  Google Scholar 

  19. Bito, L., Davson, H., Levin, E., Murray, M. & Snider, N. The concentrations of free amino acids and other electrolytes in cerebrospinal fluid, in vivo dialysate of brain, and blood plasma of the dog. J. Neurochem. 13, 1057–1067 (1966).

    Article  CAS  PubMed  Google Scholar 

  20. Robinson, I. C. & Jones, P. M. Oxytocin and neurophysin in plasma and CSF during suckling in the guinea-pig. Neuroendocrinology 34, 59–63 (1982).

    Article  CAS  PubMed  Google Scholar 

  21. Neumann, I., Russell, J. A. & Landgraf, R. Oxytocin and vasopressin release within the supraoptic and paraventricular nuclei of pregnant, parturient and lactating rats: a microdialysis study. Neuroscience 53, 65–75 (1993).

    Article  CAS  PubMed  Google Scholar 

  22. Pinol, R. A., Jameson, H., Popratiloff, A., Lee, N. H. & Mendelowitz, D. Visualization of oxytocin release that mediates paired pulse facilitation in hypothalamic pathways to brainstem autonomic neurons. PLoS ONE 9, e112138 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  23. Mignocchi, N., Krüssel, S., Jung, K., Lee, D. & Kwon, H.-B. Development of a genetically-encoded oxytocin sensor. Preprint at bioRxiv https://doi.org/10.1101/2020.07.14.202598 (2020).

  24. Ino, D., Tanaka, Y., Hibino, H. & Nishiyama, M. A fluorescent sensor for real-time measurement of extracellular oxytocin dynamics in the brain. Nat Methods 19, 1286–1294 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Jing, M. et al. A genetically encoded fluorescent acetylcholine indicator for in vitro and in vivo studies. Nat. Biotechnol. 36, 726–737 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Patriarchi, T. et al. Ultrafast neuronal imaging of dopamine dynamics with designed genetically encoded sensors. Science 360, eaat4422 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  27. Sun, F. et al. A genetically encoded fluorescent sensor enables rapid and specific detection of dopamine in flies, fish, and mice. Cell 174, 481–496 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Feng, J. et al. A genetically encoded fluorescent sensor for rapid and specific in vivo detection of norepinephrine. Neuron 102, 745–761 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Jing, M. et al. An optimized acetylcholine sensor for monitoring in vivo cholinergic activity. Nat. Methods 17, 1139–1146 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Patriarchi, T. et al. An expanded palette of dopamine sensors for multiplex imaging in vivo. Nat. Methods 17, 1147–1155 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Peng, W. et al. Regulation of sleep homeostasis mediator adenosine by basal forebrain glutamatergic neurons. Science 369, eabb0556 (2020).

    Article  CAS  PubMed  Google Scholar 

  32. Sun, F. et al. Next-generation GRAB sensors for monitoring dopaminergic activity in vivo. Nat. Methods 17, 1156–1166 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Dong, A. et al. A fluorescent sensor for spatiotemporally resolved imaging of endocannabinoid dynamics in vivo. Nat. Biotechnol. 40, 787–798 (2022).

  34. Wan, J. et al. A genetically encoded sensor for measuring serotonin dynamics. Nat. Neurosci. 24, 746–752 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Wu, Z. et al. A sensitive GRAB sensor for detecting extracellular ATP in vitro and in vivo. Neuron 110, 770–782 (2022).

  36. Kroeze, W. K. et al. PRESTO-Tango as an open-source resource for interrogation of the druggable human GPCRome. Nat. Struct. Mol. Biol. 22, 362–369 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Busnelli, M. et al. Functional selective oxytocin-derived agonists discriminate between individual G protein family subtypes. J. Biol. Chem. 287, 3617–3629 (2012).

    Article  CAS  PubMed  Google Scholar 

  38. Nagai, Y. et al. Deschloroclozapine, a potent and selective chemogenetic actuator enables rapid neuronal and behavioral modulations in mice and monkeys. Nat. Neurosci. 23, 1157–1167 (2020).

    Article  CAS  PubMed  Google Scholar 

  39. Beier, K. T. et al. Circuit architecture of VTA dopamine neurons revealed by systematic input–output mapping. Cell 162, 622–634 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Hung, L. W. et al. Gating of social reward by oxytocin in the ventral tegmental area. Science 357, 1406–1411 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Xiao, L., Priest, M. F., Nasenbeny, J., Lu, T. & Kozorovitskiy, Y. Biased oxytocinergic modulation of midbrain dopamine systems. Neuron 95, 368–384 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Barg, S. et al. Delay between fusion pore opening and peptide release from large dense-core vesicles in neuroendocrine cells. Neuron 33, 287–299 (2002).

    Article  CAS  PubMed  Google Scholar 

  43. Marvin, J. S. et al. Stability, affinity, and chromatic variants of the glutamate sensor iGluSnFR. Nat. Methods 15, 936–939 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  45. Lewis, E. M. et al. Parallel social information processing circuits are differentially impacted in autism. Neuron 108, 659–675 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Fisher, T. E. & Bourque, C. W. Calcium-channel subtypes in the somata and axon terminals of magnocellular neurosecretory cells. Trends Neurosci. 19, 440–444 (1996).

    Article  CAS  PubMed  Google Scholar 

  47. Xu, S. et al. Behavioral state coding by molecularly defined paraventricular hypothalamic cell type ensembles. Science 370, eabb2494 (2020).

    Article  CAS  PubMed  Google Scholar 

  48. Romanov, R. A. et al. Molecular interrogation of hypothalamic organization reveals distinct dopamine neuronal subtypes. Nat. Neurosci. 20, 176–188 (2017).

    Article  CAS  PubMed  Google Scholar 

  49. Tobin, V. A., Douglas, A. J., Leng, G. & Ludwig, M. The involvement of voltage-operated calcium channels in somato-dendritic oxytocin release. PLoS ONE 6, e25366 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Wheeler, D. B., Randall, A. & Tsien, R. W. Roles of N-type and Q-type Ca2+ channels in supporting hippocampal synaptic transmission. Science 264, 107–111 (1994).

    Article  CAS  PubMed  Google Scholar 

  51. Hirasawa, M., Kombian, S. B. & Pittman, Q. J. Oxytocin retrogradely inhibits evoked, but not miniature, EPSCs in the rat supraoptic nucleus: role of N- and P/Q-type calcium channels. J. Physiol. 532, 595–607 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Sudhof, T. C. & Rothman, J. E. Membrane fusion: grappling with SNARE and SM proteins. Science 323, 474–477 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  53. Kasai, H., Takahashi, N. & Tokumaru, H. Distinct initial SNARE configurations underlying the diversity of exocytosis. Physiol. Rev. 92, 1915–1964 (2012).

    Article  CAS  PubMed  Google Scholar 

  54. Schiavo, G. et al. Tetanus and botulinum-B neurotoxins block neurotransmitter release by proteolytic cleavage of synaptobrevin. Nature 359, 832–835 (1992).

    Article  CAS  PubMed  Google Scholar 

  55. Blasi, J. et al. Botulinum neurotoxin A selectively cleaves the synaptic protein SNAP-25. Nature 365, 160–163 (1993).

    Article  CAS  PubMed  Google Scholar 

  56. Bennett, M. K., Calakos, N. & Scheller, R. H. Syntaxin: a synaptic protein implicated in docking of synaptic vesicles at presynaptic active zones. Science 257, 255–259 (1992).

    Article  CAS  PubMed  Google Scholar 

  57. Yamasaki, S. et al. Cleavage of members of the synaptobrevin/VAMP family by types D and F botulinal neurotoxins and tetanus toxin. J. Biol. Chem. 269, 12764–12772 (1994).

    Article  CAS  PubMed  Google Scholar 

  58. Klapoetke, N. C. et al. Independent optical excitation of distinct neural populations. Nat. Methods 11, 338–346 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Oti, T. et al. Oxytocin influences male sexual activity via non-synaptic axonal release in the spinal cord. Curr. Biol. 31, 103–114 (2021).

    Article  CAS  PubMed  Google Scholar 

  60. Hillegaart, V., Alster, P., Uvnas-Moberg, K. & Ahlenius, S. Sexual motivation promotes oxytocin secretion in male rats. Peptides 19, 39–45 (1998).

    Article  CAS  PubMed  Google Scholar 

  61. Hughes, A. M., Everitt, B. J., Lightman, S. L. & Todd, K. Oxytocin in the central nervous system and sexual behaviour in male rats. Brain Res. 414, 133–137 (1987).

    Article  CAS  PubMed  Google Scholar 

  62. Witt, D. M. & Insel, T. R. Increased Fos expression in oxytocin neurons following masculine sexual behavior. J. Neuroendocrinol. 6, 13–18 (1994).

    Article  CAS  PubMed  Google Scholar 

  63. Argiolas, A. & Melis, M. R. The role of oxytocin and the paraventricular nucleus in the sexual behaviour of male mammals. Physiol. Behav. 83, 309–317 (2004).

    Article  CAS  PubMed  Google Scholar 

  64. Clement, P. et al. Brain oxytocin receptors mediate ejaculation elicited by 7-hydroxy-2-(di-N-propylamino) tetralin (7-OH-DPAT) in anaesthetized rats. Br. J. Pharmacol. 154, 1150–1159 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Gil, M., Bhatt, R., Picotte, K. B. & Hull, E. M. Oxytocin in the medial preoptic area facilitates male sexual behavior in the rat. Horm. Behav. 59, 435–443 (2011).

    Article  CAS  PubMed  Google Scholar 

  66. Yu, P., Zhang, M., Nan, X., Zhao, H. & Gong, D. Differences in the number of oxytocin, vasopressin, and tyrosine hydroxylase cells in brain regions associated with mating among great, midday, and Mongolian gerbils. Brain Res. 1733, 146677 (2020).

    Article  CAS  PubMed  Google Scholar 

  67. Melis, M. R. et al. Oxytocin injected into the ventral tegmental area induces penile erection and increases extracellular dopamine in the nucleus accumbens and paraventricular nucleus of the hypothalamus of male rats. Eur. J. Neurosci. 26, 1026–1035 (2007).

    Article  PubMed  Google Scholar 

  68. Argiolas, A. & Melis, M. R. Central control of penile erection: role of the paraventricular nucleus of the hypothalamus. Prog. Neurobiol. 76, 1–21 (2005).

    Article  CAS  PubMed  Google Scholar 

  69. Miesenbock, G., De Angelis, D. A. & Rothman, J. E. Visualizing secretion and synaptic transmission with pH-sensitive green fluorescent proteins. Nature 394, 192–195 (1998).

    Article  CAS  PubMed  Google Scholar 

  70. Ludwig, M. et al. Regulation of activity-dependent dendritic vasopressin release from rat supraoptic neurones. J. Physiol. 564, 515–522 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Neumann, I. D. & Landgraf, R. Balance of brain oxytocin and vasopressin: implications for anxiety, depression, and social behaviors. Trends Neurosci. 35, 649–659 (2012).

    Article  CAS  PubMed  Google Scholar 

  72. Busnelli, M., Bulgheroni, E., Manning, M., Kleinau, G. & Chini, B. Selective and potent agonists and antagonists for investigating the role of mouse oxytocin receptors. J. Pharmacol. Exp. Ther. 346, 318–327 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Manning, M. et al. Oxytocin and vasopressin agonists and antagonists as research tools and potential therapeutics. J. Neuroendocrinol. 24, 609–628 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Postina, R., Kojro, E. & Fahrenholz, F. Separate agonist and peptide antagonist binding sites of the oxytocin receptor defined by their transfer into the V2 vasopressin receptor. J. Biol. Chem. 271, 31593–31601 (1996).

    Article  CAS  PubMed  Google Scholar 

  75. Kimura, T. et al. Molecular characterization of a cloned human oxytocin receptor. Eur. J. Endocrinol. 131, 385–390 (1994).

    Article  CAS  PubMed  Google Scholar 

  76. Freund-Mercier, M. J., Stoeckel, M. E. & Klein, M. J. Oxytocin receptors on oxytocin neurones: histoautoradiographic detection in the lactating rat. J. Physiol. 480, 155–161 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Moos, F. et al. Release of oxytocin and vasopressin by magnocellular nuclei in vitro: specific facilitatory effect of oxytocin on its own release. J. Endocrinol. 102, 63–72 (1984).

    Article  CAS  PubMed  Google Scholar 

  78. Brown, C. H., Ludwig, M., Tasker, J. G. & Stern, J. E. Somato-dendritic vasopressin and oxytocin secretion in endocrine and autonomic regulation. J. Neuroendocrinol. 32, e12856 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Chen, R., Wu, X., Jiang, L. & Zhang, Y. Single-cell RNA-seq reveals hypothalamic cell diversity. Cell Rep. 18, 3227–3241 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Tobin, V. et al. Expression of exocytosis proteins in rat supraoptic nucleus neurones. J. Neuroendocrinol. 24, 629–641 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Jourdain, P., Poulain, D. A., Theodosis, D. T. & Israel, J. M. Electrical properties of oxytocin neurons in organotypic cultures from postnatal rat hypothalamus. J. Neurophysiol. 76, 2772–2785 (1996).

    Article  CAS  PubMed  Google Scholar 

  82. Wang, G., Dayanithi, G., Newcomb, R. & Lemos, J. R. An R-type Ca2+ current in neurohypophysial terminals preferentially regulates oxytocin secretion. J. Neurosci. 19, 9235–9241 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Waldherr, M. & Neumann, I. D. Centrally released oxytocin mediates mating-induced anxiolysis in male rats. Proc. Natl Acad. Sci. USA 104, 16681–16684 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. de Kock, C. P. et al. Somatodendritic secretion in oxytocin neurons is upregulated during the female reproductive cycle. J. Neurosci. 23, 2726–2734 (2003).

    Article  PubMed  PubMed Central  Google Scholar 

  85. Althammer, F. & Grinevich, V. Diversity of oxytocin neurons: beyond magno- and parvocellular cell types? J. Neuroendocrinol. https://doi.org/10.1111/jne.12549 (2017).

  86. Luther, J. A. & Tasker, J. G. Voltage-gated currents distinguish parvocellular from magnocellular neurones in the rat hypothalamic paraventricular nucleus. J. Physiol. 523, 193–209 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Zhang, B. et al. Reconstruction of the hypothalamo-neurohypophysial system and functional dissection of magnocellular oxytocin neurons in the brain. Neuron 109, 331–346 (2020).

    Article  PubMed  Google Scholar 

  88. Swanson, L. W. & Sawchenko, P. E. Hypothalamic integration: organization of the paraventricular and supraoptic nuclei. Annu. Rev. Neurosci. 6, 269–324 (1983).

    Article  CAS  PubMed  Google Scholar 

  89. Eliava, M. et al. A new population of parvocellular oxytocin neurons controlling magnocellular neuron activity and inflammatory pain processing. Neuron 89, 1291–1304 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Shimojo, M. et al. SNAREs controlling vesicular release of BDNF and development of callosal axons. Cell Rep. 11, 1054–1066 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Leng, G. & Ludwig, M. Intranasal oxytocin: myths and delusions. Biol. Psychiatry 79, 243–250 (2016).

    Article  CAS  PubMed  Google Scholar 

  92. Lee, H. J., Macbeth, A. H., Pagani, J. H. & Young, W. S. 3rd Oxytocin: the great facilitator of life. Prog. Neurobiol. 88, 127–151 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  93. Bartz, J. A., Zaki, J., Bolger, N. & Ochsner, K. N. Social effects of oxytocin in humans: context and person matter. Trends Cogn. Sci. 15, 301–309 (2011).

    CAS  PubMed  Google Scholar 

  94. Guastella, A. J. & Hickie, I. B. Oxytocin treatment, circuitry, and autism: a critical review of the literature placing oxytocin into the autism context. Biol. Psychiatry 79, 234–242 (2016).

    Article  CAS  PubMed  Google Scholar 

  95. Meyer-Lindenberg, A., Domes, G., Kirsch, P. & Heinrichs, M. Oxytocin and vasopressin in the human brain: social neuropeptides for translational medicine. Nat. Rev. Neurosci. 12, 524–538 (2011).

    Article  CAS  PubMed  Google Scholar 

  96. Falkner, A. L., Grosenick, L., Davidson, T. J., Deisseroth, K. & Lin, D. Hypothalamic control of male aggression-seeking behavior. Nat. Neurosci. 19, 596–604 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Tang, Y. et al. Viral vectors for opto-electrode recording and photometry-based imaging of oxytocin neurons in anesthetized and socially interacting rats. STAR Protoc. 3, 101032 (2022).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This research was supported by the National Key R&D Program of China (2019YFE011781), the National Natural Science Foundation of China (31925017 and 31871087), the NIH BRAIN Initiative (1U01NS113358 and 1U01NS120824), the Feng Foundation of Biomedical Research, the Clement and Xinxin Foundation, and grants from the Peking–Tsinghua Center for Life Sciences and the State Key Laboratory of Membrane Biology at Peking University School of Life Sciences (to Y.L.); the National Natural Science Foundation of China (81871134) (to P.W.); NIH grants (R01MH101377, R01HD092596, U19NS107616 and U01NS113358), the Mathers Foundation and the Vulnerable Brain Project (to D.L.); the Levy Leon fellowship (to M.L.); the Uehara Memorial Foundation, a JSPS Overseas Research Fellowship and the Osamu Hayaishi Memorial Scholarship (to T.O.); Swiss National Science Foundation grants (CRSK-3_190779, 310030_192463, 310030E_173565) and the Synapsis Foundation (to R.S.); German Research Foundation (DFG) grants (GR 3619/13-1, GR 3619/15-1, GR 3619/16-1) and the SFB Consortium 1158-2 (to V.G.); and the Humboldt Research Fellowship (to A.K.). We thank Y. Rao for sharing the two-photon microscope and X. Lei at PKU-CLS and the National Center for Protein Sciences at Peking University for providing support for the Opera Phenix high-content screening system and imaging platform. We thank L. Barteczko for packaging the viral vectors used for the in vivo experiments of A.K. and V.G. We thank X. Yu at Peking University, B. Li at Cold Spring Harbor Laboratory, D. Anderson at the California Institute of Technology, M. Andermann at Harvard University, P. Kalugin at Harvard University and members of the Li laboratory for helpful suggestions and comments on the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

Y.L. supervised the project. H.W. and L.G. performed the experiments related to the development, optimization and characterization of the sensors in cultured cells. T.Q. performed the two-photon imaging of OT dynamics in acute brain slices. L.W. performed the immunostaining experiments in brain slices. L.M., T.O. and A.K. performed the in vivo ICV infusion experiments in mice and rats under the supervision of D.L. and V.G. P.W. performed the in vivo fiber photometry recording and optogenetic experiments in mice under the supervision of M.L. Y.T. performed the optogenetic experiments in rats under the supervision of R.S. All authors contributed to the interpretation and analysis of the data. T.Q. and Y.L. wrote the manuscript with input from all authors.

Corresponding author

Correspondence to Yulong Li.

Ethics declarations

Competing interests

Y.L. and H.W. have filed patent applications whose value might be affected by this publication. The remaining authors declare no competing interests.

Peer review

Peer review information

Nature Biotechnology thanks Yongsoo Kim and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

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

Extended data

Extended Data Fig. 1 Characterization of GRABOT sensors in HEK293T cells.

a. Representative expression and response images of OT1.0 to 100 nM OT in HEK293T cells. Scale bar, 20 μm. b. Example traces (left) and summary data (right) of OT1.0-expressing HEK293T cells pre-incubated with saline, 10 μM YM022, or 10 μM Atosiban in response to OT. Saline: n = 30 cells from 3 coverslips [30/3], YM022: n = 39/4, Atosiban: n = 44/3. Two-tailed Student’s t-tests, p = 0.84 (between Saline and YM022); p = 9.8 × 10−18 (between Saline and Atosiban). c. Excitation (Ex) and emission (Em) spectra of the OT1.0 sensor with or without 100 nM OT. d. Representative images of OT1.0 and OTmut expressed in HEK293T cells in saline and in the presence of 100 nM OT. Also shown is RFP-CAAX expression, showing localization at the plasma membrane. The images at the right show the change in OT1.0 and OTmut fluorescence in response to OT application. White rectangle with enhanced contrast showing OTmut expressing HEK293T cells in saline. Scale bars, 20 μm. e. Summary of the peak change in OT1.0 and OTmut fluorescence measured in HEK293T cells in response to 100 nM OT. OT1.0: n = 45 cells from 3 coverslips; OTmut: n = 31 cells from 4 coverslips. Two-tailed Student’s t-tests were performed (p = 5.5 × 10−31). f. Dose–response curves for OT1.0 and OTmut expressed in HEK293T cells in response to the indicated concentrations of OT and AVP, with the corresponding EC50 values shown. The data were normalized to the maximal response measured in OT group. The dosage curves of OT1.0 to OT/AVP were averaged from 9 individual trials, with 3–4 wells per trial. g. Dose-response curves for OT1.0 expressed in HEK293T cells in response to the indicated concentrations of OT and its orthologous peptides, with amino acid sequence alignment shown. n = 4 coverslips for each group. h. Summary of the peak change in OT1.0 fluorescence measured in HEK293T cells in response to the indicated compounds applied at 1 μM (CRF, NTS, NPY, and VIP) or 10 μM (Glu, GABA, Gly, DA, NE, and 5-HT), normalized to the peak response measured in OT; n = 4 wells per group. CRF, corticotropin-releasing factor; NTS, neurotensin; NPY, neuropeptide Y; VIP, vasoactive intestinal peptide; Glu, glutamate; GABA, γ-aminobutyric acid; Gly, glycine; DA, dopamine; NE, norepinephrine; and 5-HT, 5-hydroxytryptamine (serotonin). One-way ANOVA test was performed for all groups (F (10, 3.88) =132.3, p = 2.0 × 10−4); Dunnett’s T3 multiple comparisons tests were performed (p = 4.6 × 10−3, 3.8 × 10−3, 4.4 × 10−3, 4.0 × 10−3, 4.2 × 10−3, 4.2 × 10−3, 4.6 × 10−3, 4.8 × 10−3, 4.5 × 10−3, and 4.5 × 10−3 (between OT and CRF, NTS, SST, NPY, CCK, VIP, Glu, GABA, ACh, DA)). i. Representative traces of the OT1.0 signal evoked by OT puffing at indicated concentrations. j. Summary of the OT1.0 signal time constant at indicated OT concentrations (n = 19 cells for 30 nM OT, n = 11 cells for 100 nM OT and n = 9 cells for 1000 nM OT). k. Representative traces and summary ΔF/F0 of OT1.0 when bath application of OT at saturated concentration using line-scanning mode. n = 6 trials. l. The calibration curve of OT1.0 dose-dependent fluorescence response in line-scanning mode, which is used to estimate the local OT concentration reaching the cells during local puffing experiments. m. The association rate constant of the OT1.0 sensor for OT. Local OT concentrations were estimated from i (n = 19 cells for 30 nM OT, n = 11 cells for 100 nM OT and n = 9 cells for 1000 nM OT). **p < 0.01, and ***p < 0.001. Summary data are presented as the mean±s.e.m.

Source data

Extended Data Fig. 2 Negligible downstream signaling coupling of OT1.0 sensor in HEK293T cells.

a, b. Representative expression images in BFP channel, pseudocolor images (top) and ΔF/F0 traces (bottom) showing the Ca2+ response to the indicated concentrations of OT or ATP in HEK293T cells expressing bOTR-BFP (a) or OT1.0-BFP (b). c. Summary of peak Ca2+ ΔF/F0 for bOTR or OT1.0 expressed HEK293T cells corresponding to (a and b) at indicated OT concentrations, with the corresponding EC50 value shown. The data were normalized to the peak response measured in 100 μM ATP. n = 3 coverslips for each group. Two-tailed Student’s t-tests were performed between bOTR and OT1.0 (p = 0.56, 0.29, 0.018, and 0.013 for 0.01, 0.1, 1, and 10 nM OT). d. β-arrestin coupling was measured using the Tango assay in cells expressing the bovine OTR (bOTR), OT1.0, bOTR and OT1.0, or no receptor (Control). n = 3 wells each. Two-tailed Student’s t-tests, p = 0.12 (between bOTR and bOTR+OT1.0); p = 6.8 × 10−7 (between bOTR and OT1.0); p = 2.9 × 10−3 (between OT1.0 and Control). e. Representative images (left) and summary (right) of the fluorescence change in OT1.0-expressing neurons in response to a 2-hour continuous OT application. n = 5 cultures with >30 cells each. Row matched one-way ANOVA, with the Geisser-Greenhouse correction, F (1.509,6.037) =0.16, p = 0.79. *p < 0.05, **p < 0.01, ***p < 0.001, and n.s., not significant. All scale bars represent 20 μm. Summary data are presented as the mean±s.e.m.

Source data

Extended Data Fig. 3 Chemogenetic activation of oxytocinergic neurons induces OT release.

a. Schematic diagram depicting the chemogenetic activation experiments. A mixture of AAVs (EF1α-dio-hM3Dq-mCherry and hSyn-OT1.0) was injected into the PVN of OT-Cre mice. As a control, hSyn-OT1.0 was injected into the PVN of OT-Cre x Ai14 mice (no-hM3Dq). The PVN and third ventricle (3 V) are indicated. b. Left: representative 2-photon microscopy merged images of OT1.0 (green channel) and the RFP channel (red, mCherry expression for OT-hM3Dq and tdTomato for no-hM3Dq). Right: responses of the OT1.0 sensor measured in ACSF (baseline), 60 nM DCZ, and 100 nM OT. Scale bars, 100 μm. c, d. Example OT1.0 traces (c) and peak change (d) in OT1.0 fluorescence; where indicated, DCZ or OT were applied to the slices. n = 7 slices from 2 mice for OT-hM3Dq and n = 5 slices from 2 mice for no-hM3Dq. Two-tailed Student’s t-tests were performed (p = 8.9 × 10−6 (left) and 0.41 (right)). ***p < 0.001 and n.s., not significant. Summary data are presented as the mean±s.e.m.

Source data

Extended Data Fig. 4 Immunostaining of OT neurites in the VTA and PVN.

a. Representative images of OT-positive axons in the VTA of WT and OT-KO mice. Green, OT antibody; blue, DAPI. Scale bars: left, 1 mm; right, 100 μm. b-d. Schematic drawings depicting the experimental strategy (b), representative images (c), and quantification data (d) showing the colocalization of PSD95-EGFP with OT-positive neurites in the PVN or VTA of coronal slices. Red, OT (indicated by OT antibody in the VTA or tdTomato in the PVN); green, PSD95 (indicated by GFP antibody); blue, DAPI. n = 3 slices from 1 mouse. Two-tailed Student’s t-tests were performed (p = 0.014). Scale bars: left, 200 μm; right, 20 μm. *p < 0.05. Summary data are presented as the mean±s.e.m.

Source data

Extended Data Fig. 5 Electrical stimulation evoked OT and Glutamate release in the PVN, S1 and dStr.

a. Schematic illustration depicting the experimental setup for panels (b-g). b and d. Representative fluorescence images (left) and ΔF/F0 pseudocolor images (right) showing the expression and electrical stimulation-induced response of OT1.0 or iGluSnFR in the PVN, S1, and dStr. c and e. Representative traces of electrical stimulation evoked and 100 nM OT or 100 μM Glu perfusion induced OT1.0 or iGluSnFR signals. f and g. Summary of the change in OT1.0 or iGluSnFR fluorescence in response to 100 pulses or ligand application (n = 5 slices from 4 mice [5/4], 6/3, and 6/3 mice for OT1.0 in the PVN, S1, and dStr, respectively; n = 12 slices from 4 mice [12/4], 6/2, and 7/2 mice for iGluSnFR in the PVN, S1, and dStr, respectively.) Two-tailed Student’s t-tests were performed (f: for electrical stimulation, p = 2.9 × 10−3, 0.11, and 0.034 between basal and peak ΔF/F0 for PVN, S1, and dStr, respectively; p = 3.1 × 10−3 between PVN and S1; p = 9.5 × 10−3 between PVN and dStr; for OT perfusion, p = 0.037 between PVN and S1; p = 0.27 between PVN and dStr; g: for electrical stimulation, p = 0.50 between PVN and S1; p = 0.63 between PVN and dStr; p = 0.77 between S1 and dStr; for Glu perfusion, p = 0.92 between PVN and S1; p = 0.12 between PVN and dStr; p = 0.26 between S1 and dStr). The data of OT1.0 and iGluSnFR in the PVN are reused from Fig. 5. *p < 0.05, **p < 0.01, and n.s., not significant. All scale bars represent 100 μm. Summary data are presented as the mean±s.e.m.

Source data

Extended Data Fig. 6 Application of OT with different concentrations in OT1.0 expressing slices.

a-c. Example pseudocolor images (top) and fluorescence traces (bottom) of OT1.0-expressing slices containing the PVN (a) or VTA (b, c) before and after application of 100 nM OT (a, b) or AVP (c). d, f. Representative fluorescence images showing the expression of OT1.0 in the VTA (d) and PVN (f). e, g. Example traces of OT1.0 signals in response to 1, 10, 100, and 1000 nM OT application in the VTA (e) and PVN (g). h, i. Summary of peak and post-application (wash) ΔF/F0 in response to 1, 10, 100, and 1000 nM OT in the VTA (h) and PVN (i). n = 3 slices from 3 mice [3/3] and 3/2 mice for OT1.0 in the VTA and PVN, respectively. Two-tailed Student’s t-tests were performed (h: p = 0.057, 0.024, 0.18, and 0.31 (between Peak and Wash) for 1, 10, 100, and 1000 nM OT, respectively; i: p = 0.011, 8.5 × 10−3, 0.021, and 9.2 × 10−3 (between Peak and Wash) for 1, 10, 100, and 1000 nM OT, respectively). *p < 0.05, **p < 0.01, and n.s., not significant. All scale bars represent 100 μm. Summary data are presented as the mean±s.e.m.

Source data

Extended Data Fig. 7 Dissecting the Ca2+ sources underlying somatodendritic OT release.

a. Top left: representative image of OT1.0 expressed in the PVN (left). Also shown are example pseudocolor images (top row), corresponding traces (bottom row), and summary of the peak OT1.0 response (right) to 100 electrical stimuli delivered at 20 Hz in ACSF, nimodipine (Nim; 10 μM), Cd2+ (200 μM), or 10 nM OT. n = 5 slices from 3 mice for ACSF, Nim, and Cd2+; n = 3 slices from 2 mice for OT. Paired two-tailed Student’s t-tests were performed (p = 9.7 × 10−4 (left) and 0.031 (right)). b. Representative pseudocolor images (top row) and corresponding traces (bottom row) of OT1.0 expressed in the PVN in response to 100 electrical stimuli delivered at 20 Hz in ACSF, ω-Agx-IVA (0.3 μM), ω-CTx (1 μM), or SNX-482 (100 nM) to block P/Q-, N-, and R-type VGCCs, respectively. c. Representative fluorescence image of iGluSnFR (top left) and schematic drawing depicting the experimental strategy (bottom left), related to Fig. 3e, f. Example pseudocolor images (top) and traces (bottom) of the change in iGluSnFR fluorescence in response to 100 electrical pulses delivered at 20 Hz in ACSF, ω-Agx-IVA (0.3 μM), ω-CTx (1 μM), nimodipine (Nim; 10 μM), or Cd2+ (200 μM) to block P/Q-, N-, L-type or all VGCCs, respectively (slices were sequentially perfused with the indicated blockers). *p < 0.05 and **p < 0.01. All scale bars represent 100 μm. Summary data are presented as the mean±s.e.m.

Source data

Extended Data Fig. 8 Somatodendritic OT release is insensitive to cell type-specific expression of TeNT.

a-j. Schematic drawings depicting the experimental strategy (a, f), representative images showing the expression (b, g) and peak response of OT1.0 sensor (c, h), average traces (d, i), and summary (e, j) of OT1.0 in response to 100 pulses stimulation delivered at 20 Hz or OT perfusion applied under the indicated conditions. n = 7 slices from 2 mice [7/2] and 5/2 mice for PVN and VTA, respectively. Two-tailed Student’s t-tests were performed (e: p = 0.83 and 0.47 between Ctrl and DIO-TeNT for electrical stimulation and OT perfusion; j: p = 0.038 and 0.78 between Ctrl and DIO-TeNT for electrical stimulation and OT perfusion). The data of Ctrl groups in the PVN and VTA are reused from Fig. 5. Scale bars, 100 μm. *p < 0.05 and n.s., not significant. Summary data are presented as the mean±s.e.m.

Source data

Extended Data Fig. 9 OT1.0 can detect intraventricularly injected OT in the PVN of rats.

a. Schematic diagram depicting the experimental strategy for in vivo recording of OT1.0 in rats. An AAV expressing hSyn-OT1.0 was injected into the PVN of WT Sprague Dawley female rats; optical fibers were placed in the above PVN 2 weeks later, 10 mM OT (1 µL) was injected into the lateral ventricle during recording, and 470- nm light was used to excite the OT1.0 sensor together with isosbestic control signal (405 nm). b. Exemplary histological verification of the optic fiber placement and the OT1.0 expression in the periPVN area. OT1.0 was stained with anti-GFP antibody for visualization. OF, optic fiber. Scale bar, 400 µm. c. Average trace and quantification of OT1.0 signal. The OT1.0 and isosbestic signals were sampled at 1 Hz. n = 4 rats. Paired two-tailed Student’s t-tests were performed (p = 1.3 × 10−3). **p < 0.01. Summary data are presented as the mean±s.e.m.

Source data

Extended Data Fig. 10 Optogenetic activation of neurons induces somatodendritic and axonal OT release in vivo in freely moving rats.

a. Schematic illustrations depicting the optogenetic activation experiments with both the OT sensor and the optogenetic stimulation by ChrimsonR in the PVN. b. Representative traces recorded in the rat PVN of changes in normalized fluorescent emission ΔF/F0 (525 nm) during excitation at isosbestic control (405 nm, in purple) or sensor wavelength (465 nm, in green) before, during and after stimulation of ChrimsonR with pulses (at 593 nm, in orange) of 10 ms at a frequency of 1, 7 or 30 Hz for a total duration of 5 s). c. Summary of the peak changes in OT1.0 fluorescence emission (at 525 nm) during excitation at the sensor wavelength (465 nm, in green) or isosbestic control (405 nm, in purple) in rats PVN (at 1, 7, or 30 Hz photostimulation). n = 4 rats per group. Paired two-tailed Student’s t-tests were performed between 465 nm and 405 nm (p = 9.1 × 10−3, 6.1 × 10−4, and 6.6 × 10−4 for 1, 7, and 30 Hz, respectively). d. Summary of the rise time (‘on’) and decay time (‘off’) constants (T1/2) of the OT1.0 response to photostimulation. n = 4 rats per group. e. Schematic illustrations depicting the optogenetic activation experiments with OT sensor expressed in the SON and the optogenetic stimulator ChrimsonR expressed in the PVN. f. Representative traces recorded in the rat SON of changes in normalized fluorescent emission ΔF/F0 (525 nm) during excitation at isosbestic control (405 nm, in purple) or sensor wavelength (465 nm, in green) before, during and after stimulation of ChrimsonR (at 593 nm, in orange) with pulses of 10 ms at a frequency of 1, 7, or 30 Hz for a total duration of 5 s). g. Summary of the peak change in OT1.0 fluorescence emission (at 525 nm) during excitation at the sensor wavelength (465 nm, in green) or isosbestic control (405 nm, in purple) in rats SON (at 1, 7, or 30 Hz photostimulation). n = 4 rats per group. Paired two-tailed Student’s t-tests were performed between 465 nm and 405 nm (p = 0.043, 8.8 × 10−3, and 0.034 for 1, 7, and 30 Hz, respectively). h. Summary of the rise time (‘on’) and decay time (‘off’) constants (T1/2) of the OT1.0 response to photostimulation. n = 4 rats per group. *p < 0.05, **p < 0.01 and ***p < 0.001. Summary data are presented as the mean±s.e.m.

Source data

Supplementary information

Reporting Summary

Supplementary Table 1

Primer sequences used in this study.

Supplementary Video 1

Single-trial measurement of OT1.0 fluorescence in the VTA in the male mouse brain during mating (related to Fig. 5).

Supplementary Video 2

Single-trial measurement of OT1.0 fluorescence in the PVN in the male mouse brain during mating (related to Fig. 5).

Supplementary Video 3

Single-trial measurement of OT1.0 fluorescence in the PFC in the male mouse brain during mating (related to Fig. 5).

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 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.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 9

Statistical source data.

Source Data Extended Data Fig. 10

Statistical source data.

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Qian, T., Wang, H., Wang, P. et al. A genetically encoded sensor measures temporal oxytocin release from different neuronal compartments. Nat Biotechnol 41, 944–957 (2023). https://doi.org/10.1038/s41587-022-01561-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41587-022-01561-2

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