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:

Cortico-cortical transfer of socially derived information gates emotion recognition

Abstract

Emotion recognition and the resulting responses are important for survival and social functioning. However, how socially derived information is processed for reliable emotion recognition is incompletely understood. Here, we reveal an evolutionarily conserved long-range inhibitory/excitatory brain network mediating these socio-cognitive processes. Anatomical tracing in mice revealed the existence of a subpopulation of somatostatin (SOM) GABAergic neurons projecting from the medial prefrontal cortex (mPFC) to the retrosplenial cortex (RSC). Through optogenetic manipulations and Ca2+ imaging fiber photometry in mice and functional imaging in humans, we demonstrate the specific participation of these long-range SOM projections from the mPFC to the RSC, and an excitatory feedback loop from the RSC to the mPFC, in emotion recognition. Notably, we show that mPFC-to-RSC SOM projections are dysfunctional in mouse models relevant to psychiatric vulnerability and can be targeted to rescue emotion recognition deficits in these mice. Our findings demonstrate a cortico-cortical circuit underlying emotion recognition.

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: A subpopulation of SOM-positive cells within the mPFC shows long-range projections to the RSC.
Fig. 2: mPFC-to-RSC SOM terminals are involved in emotion recognition.
Fig. 3: mPFC-to-RSC SOM terminals amplify RSC differential activation responses between encounter of a neutral versus emotionally altered mouse.
Fig. 4: RSC pyramidal neurons generally encode social information.
Fig. 5: RSC pyramidal neurons projecting to the mPFC regulate emotion recognition.
Fig. 6: FC of the DMN components during emotion recognition in humans.
Fig. 7: Emotion recognition deficits in genetic and pharmacological schizophrenia-relevant mouse models are associated with functional alterations of the mPFC–RSC inhibitory–excitatory circuit.

Similar content being viewed by others

Data availability

All scripts used in the paper were based on the provided scripts from the Tucker-Davis Technology database and were adjusted for our needs. The individual data and statistical analyses are shared on the open repository at https://doi.org/10.48557/FEXUTZ. All raw images, photometry recordings or videos are shared on reasonable request to the corresponding author due to their large sizes. Human data are available on the IMAGEN database via https://imagen-project.org/the-imagen-dataset/.

References

  1. Ferretti, V. & Papaleo, F. Understanding others: emotion recognition in humans and other animals. Genes Brain Behav. 18, e12544 (2019).

    Article  PubMed  Google Scholar 

  2. Panksepp, J. & Panksepp, J. B. Toward a cross-species understanding of empathy. Trends Neurosci. 36, 489–496 (2013).

    Article  CAS  PubMed  Google Scholar 

  3. Keysers, C., Knapska, E., Moita, M. A. & Gazzola, V. Emotional contagion and prosocial behavior in rodents. Trends Cogn. Sci. 26, 688–706 (2022).

    Article  PubMed  Google Scholar 

  4. Weinreb, S., Li, F. & Kurtz, M. M. A meta-analysis of social cognitive deficits in schizophrenia: does world region matter? Schizophrenia Res. 243, 206–213 (2022).

    Article  Google Scholar 

  5. Yeung, M. K. A systematic review and meta-analysis of facial emotion recognition in autism spectrum disorder: the specificity of deficits and the role of task characteristics. Neurosci. Biobehav. Rev. 133, 104518 (2022).

    Article  PubMed  Google Scholar 

  6. Behrens, T. E., Hunt, L. T. & Rushworth, M. F. The computation of social behavior. Science 324, 1160–1164 (2009).

    Article  CAS  PubMed  Google Scholar 

  7. Green, M. F., Horan, W. P. & Lee, J. Social cognition in schizophrenia. Nat. Rev. Neurosci. 16, 620–631 (2015).

    Article  CAS  PubMed  Google Scholar 

  8. Mars, R. B. et al. On the relationship between the “default mode network” and the “social brain. Front. Hum. Neurosci. 6, 189 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  9. Whitesell, J. D. et al. Regional, layer, and cell-type-specific connectivity of the mouse default mode network. Neuron 109, 545–559 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Broyd, S. J. et al. Default-mode brain dysfunction in mental disorders: a systematic review. Neurosci. Biobehav. Rev. 33, 279–296 (2009).

    Article  PubMed  Google Scholar 

  11. Li, P. et al. Structural and functional brain network of human retrosplenial cortex. Neurosci. Lett. 674, 24–29 (2018).

    Article  CAS  PubMed  Google Scholar 

  12. Greicius, M. D., Supekar, K., Menon, V. & Dougherty, R. F. Resting-state functional connectivity reflects structural connectivity in the default mode network. Cereb. Cortex 19, 72–78 (2009).

    Article  PubMed  Google Scholar 

  13. Wager, T. D. et al. A Bayesian model of category-specific emotional brain responses. PLoS Comput. Biol. 11, e1004066 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  14. Li, W., Mai, X. & Liu, C. The default mode network and social understanding of others: what do brain connectivity studies tell us. Front. Hum. Neurosci. 8, 74 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Pessoa, L. Emotion and the interactive brain: insights from comparative neuroanatomy and complex systems. Emot. Rev. 10, 204–216 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Hiser, J. & Koenigs, M. The multifaceted role of the ventromedial prefrontal cortex in emotion, decision making, social cognition, and psychopathology. Biol. Psychiatry 83, 638–647 (2018).

    Article  PubMed  Google Scholar 

  17. Scheggia, D. et al. Somatostatin interneurons in the prefrontal cortex control affective state discrimination in mice. Nat. Neurosci. 23, 47–60 (2020).

    Article  CAS  PubMed  Google Scholar 

  18. Urban-Ciecko, J. & Barth, A. L. Somatostatin-expressing neurons in cortical networks. Nat. Rev. Neurosci. 17, 401–409 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Ferretti, V. et al. Oxytocin signaling in the central amygdala modulates emotion discrimination in mice. Curr. Biol. 29, 1938–1953 (2019).

    Article  CAS  PubMed  Google Scholar 

  20. Gozzi, A. & Schwarz, A. J. Large-scale functional connectivity networks in the rodent brain. NeuroImage 127, 496–509 (2016).

    Article  PubMed  Google Scholar 

  21. Huang, S., Wu, S. J., Sansone, G., Ibrahim, L. A. & Fishell, G. Layer 1 neocortex: gating and integrating multidimensional signals. Neuron 112, 184–200 (2024).

    Article  CAS  PubMed  Google Scholar 

  22. Katona, L. et al. Sleep and movement differentiates actions of two types of somatostatin-expressing GABAergic interneuron in rat hippocampus. Neuron 91, 1183 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Bertero, A., Feyen, P. L. C., Zurita, H. & Apicella, A. J. A non-canonical cortico-amygdala inhibitory loop. J. Neurosci. 39, 8424–8438 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Assous, M., Dautan, D., Tepper, J. M. & Mena-Segovia, J. Pedunculopontine glutamatergic neurons provide a novel source of feedforward inhibition in the striatum by selectively targeting interneurons. J. Neurosci. 39, 4727–4737 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  25. Dautan, D. et al. Segregated cholinergic transmission modulates dopamine neurons integrated in distinct functional circuits. Nat. Neurosci. 19, 1025–1033 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Vander Weele, C. M. et al. Dopamine enhances signal-to-noise ratio in cortical-brainstem encoding of aversive stimuli. Nature 563, 397–401 (2018).

    Article  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Czajkowski, R. et al. Encoding and storage of spatial information in the retrosplenial cortex. Proc. Natl Acad. Sci. USA 111, 8661–8666 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Ahrlund-Richter, S. et al. A whole-brain atlas of monosynaptic input targeting four different cell types in the medial prefrontal cortex of the mouse. Nat. Neurosci. 22, 657–668 (2019).

    Article  PubMed  Google Scholar 

  30. Allen, E. A. et al. Tracking whole-brain connectivity dynamics in the resting state. Cereb. Cortex 24, 663–676 (2014).

    Article  PubMed  Google Scholar 

  31. Tu, Y. et al. Distinct thalamocortical network dynamics are associated with the pathophysiology of chronic low back pain. Nat. Commun. 11, 3948 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Bluhm, R. L. et al. Retrosplenial cortex connectivity in schizophrenia. Psychiatry Res. 174, 17–23 (2009).

    Article  PubMed  Google Scholar 

  33. Manago, F. et al. Genetic disruption of Arc/Arg3.1 in mice causes alterations in dopamine and neurobehavioral phenotypes related to schizophrenia. Cell Rep. 16, 2116–2128 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Jentsch, J. D. & Roth, R. H. The neuropsychopharmacology of phencyclidine: from NMDA receptor hypofunction to the dopamine hypothesis of schizophrenia. Neuropsychopharmacol. 20, 201–225 (1999).

    Article  CAS  Google Scholar 

  35. Scheggia, D., Bebensee, A., Weinberger, D. R. & Papaleo, F. The ultimate intra-/extra-dimensional attentional set-shifting task for mice. Biol. Psychiatry 75, 660–670 (2014).

    Article  PubMed  Google Scholar 

  36. Vesuna, S. et al. Deep posteromedial cortical rhythm in dissociation. Nature 586, 87–94 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Keum, S. et al. A missense variant at the Nrxn3 locus enhances empathy fear in the mouse. Neuron 98, 588–601 (2018).

    Article  CAS  PubMed  Google Scholar 

  38. Zhang, C. et al. Dynamics of a disinhibitory prefrontal microcircuit in controlling social competition. Neuron 110, 516–531 (2022).

    Article  CAS  PubMed  Google Scholar 

  39. Melzer, S. et al. Long-range-projecting GABAergic neurons modulate inhibition in hippocampus and entorhinal cortex. Science 335, 1506–1510 (2012).

    Article  CAS  PubMed  Google Scholar 

  40. Yamawaki, N. et al. Long-range inhibitory intersection of a retrosplenial thalamocortical circuit by apical tuft-targeting CA1 neurons. Nat. Neurosci. 22, 618–626 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Lee, A. T., Vogt, D., Rubenstein, J. L. & Sohal, V. S. A class of GABAergic neurons in the prefrontal cortex sends long-range projections to the nucleus accumbens and elicits acute avoidance behavior. J. Neurosci. 34, 11519–11525 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Melzer, S. et al. Distinct corticostriatal GABAergic neurons modulate striatal output neurons and motor activity. Cell Rep. 19, 1045–1055 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Vogt, B. A., Vogt, L. & Laureys, S. Cytology and functionally correlated circuits of human posterior cingulate areas. NeuroImage 29, 452–466 (2006).

    Article  PubMed  Google Scholar 

  44. Munoz, W., Tremblay, R., Levenstein, D. & Rudy, B. Layer-specific modulation of neocortical dendritic inhibition during active wakefulness. Science 355, 954–959 (2017).

    Article  CAS  PubMed  Google Scholar 

  45. Kawaguchi, Y. & Kubota, Y. GABAergic cell subtypes and their synaptic connections in rat frontal cortex. Cereb. Cortex 7, 476–486 (1997).

    Article  CAS  PubMed  Google Scholar 

  46. Brockway, D. F. et al. Somatostatin peptide signaling dampens cortical circuits and promotes exploratory behavior. Cell Rep. 42, 112976 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Dienel, S. J. et al. Diagnostic specificity and association with cognition of molecular alterations in prefrontal somatostatin neurons in schizophrenia. JAMA Psychiatry 80, 1235–1245 (2023).

    Article  PubMed  Google Scholar 

  48. Biedermann, F., Frajo-Apor, B. & Hofer, A. Theory of mind and its relevance in schizophrenia. Curr. Opin. Psychiatry 25, 71–75 (2012).

    Article  PubMed  Google Scholar 

  49. Fisher, M. H., Moskowitz, A. L. & Hodapp, R. M. Differences in social vulnerability among individuals with autism spectrum disorder, Williams syndrome, and Down syndrome. Res Autism Spectr. Disord. 7, 931–937 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  50. Toth, M. The other side of the coin: hypersociability. Genes Brain Behav. 18, e12512 (2019).

    Article  PubMed  Google Scholar 

  51. Scheggia, D. et al. Reciprocal cortico-amygdala connections regulate prosocial and selfish choices in mice. Nat. Neurosci. 25, 1505–1518 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Rocchi, F. et al. Increased fMRI connectivity upon chemogenetic inhibition of the mouse prefrontal cortex. Nat. Commun. 13, 1056 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Bertero, A. et al. Autism-associated 16p11.2 microdeletion impairs prefrontal functional connectivity in mouse and human. Brain 141, 2055–2065 (2018).

    Article  PubMed  Google Scholar 

  54. Pagani, M. et al. mTOR-related synaptic pathology causes autism spectrum disorder-associated functional hyperconnectivity. Nat. Commun. 12, 6084 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Vecchia, D. et al. Temporal sharpening of sensory responses by layer V in the mouse primary somatosensory cortex. Curr. Biol. 30, 1589–1599 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Schumann, G. et al. The IMAGEN study: reinforcement-related behaviour in normal brain function and psychopathology. Mol. Psychiatry 15, 1128–1139 (2010).

    Article  CAS  PubMed  Google Scholar 

  57. Grosbras, M. H. & Paus, T. Brain networks involved in viewing angry hands or faces. Cereb. Cortex 16, 1087–1096 (2006).

    Article  PubMed  Google Scholar 

  58. Fu, Z. et al. Transient increased thalamic-sensory connectivity and decreased whole-brain dynamism in autism. NeuroImage 190, 191–204 (2019).

    Article  PubMed  Google Scholar 

  59. Calhoun, V. D., Adali, T., Pearlson, G. D. & Pekar, J. J. A method for making group inferences from functional MRI data using independent component analysis. Hum. Brain Mapp. 14, 140–151 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Preti, M. G., Bolton, T. A. & Van De Ville, D. The dynamic functional connectome: state-of-the-art and perspectives. NeuroImage 160, 41–54 (2017).

    Article  PubMed  Google Scholar 

  61. Xie, H. et al. Efficacy of different dynamic functional connectivity methods to capture cognitively relevant information. NeuroImage 188, 502–514 (2019).

    Article  PubMed  Google Scholar 

  62. Gonzalez-Castillo, J. & Bandettini, P. A. Task-based dynamic functional connectivity: recent findings and open questions. NeuroImage 180, 526–533 (2018).

    Article  PubMed  Google Scholar 

  63. Xie, H. et al. Time-varying whole-brain functional network connectivity coupled to task engagement. Netw. Neurosci. 3, 49–66 (2019).

    Article  PubMed  Google Scholar 

  64. Cole, M. W. et al. Task activations produce spurious but systematic inflation of task functional connectivity estimates. NeuroImage 189, 1–18 (2019).

    Article  PubMed  Google Scholar 

  65. Fair, D. A. et al. A method for using blocked and event-related fMRI data to study “resting state” functional connectivity. NeuroImage 35, 396–405 (2007).

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

We are grateful to G. Contarini, M. Morini, D. Cantatore, G. Pruzzo, B. Chiarenza, A. Monteforte and C. Chiabrera for technical support. We thank O. Yizhar for suggestions and discussions on optogenetics experiments. This work was supported by funding from the Istituto Italiano di Tecnologia, the Ministero della Salute (project GR-2016-02362413) and Fondazione Telethon Italia (project GGP19103) to F.P. Human experiments were supported by the ERC Advanced Grant STRATIFY (695313), Human Brain Project (HBP SGA 945539), the Horizon Europe project environMENTAL (101057429), an NSFC RFIS grant to G.S. and the National Natural Science Foundation of China (82102138) to X.C.

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization: D.D. and F.P. Methodology and investigation: D.D., A.M., F. Maltese, X.C., D.M., C.M., F.A., A.B., F.D., G.M.L., A. Galbusera, D.V., M.P., A. Gozzi, G.S., F. Managò and F.P. Resource: T.F., A. Gozzi, G.S. and F.P. Writing: D.D., X.C., D.V., M.P., A. Gozzi and F.P. Visualization and analysis: D.D., A.M., F. Maltese, X.C., D.M., C.M., A. Galbusera, D.V., M.P., A. Gozzi, G.S., F. Managò and F.P. Human experiments: X.C. and G.S. Supervision: F.P. All the authors revised the manuscript. F. M. and X.C. participated equally as co-second authors.

Corresponding author

Correspondence to Francesco Papaleo.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Neuroscience thanks Ewelina Knapska, Nancy Padilla-Coreano 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 Refer to Fig. 1.

In mice expressing TdTomato in all somatostatin neurons (SOM-tdT) quantification of immunolabelled somatostatin cell of the mPFC (a) revealed a high-colocalization (n = 3 mice, 91.31 + /-2.02% overlapping). The same immunostaining for (b) Calretinin (CR), (c) Nitric oxide synthase (NOS), or (d) Parvalbumin (PV) revealed (e) an overall low colocalization with TdTomato positive cells (CR:13.24 + /-1.29, PV: 7.32 + /-0.94, NOS: 9.50 + /-1.73%). (f-g) Following injection of Ctb-560 in the RSC of SOM-TdT mice, we observed CtB expressing cells across the entire cortex, including the different subregions of the mPFC (the cingulate, prelimbic and infralimbic cortex, repeated in 3 animals). (h-i) The overlapping of all sections containing the mPFC and the mapping of CtB-Positive neurons revealed a distribution along the different layers and the dorsoventral axis. Data are represented as means + /-SEM. Cg, cingulate cortex; PrL, Prelimbic cortex, iFl, infralimbic cortex. Bar plots are represented as mean ± SEM. * is used for P < 0.05.

Extended Data Fig. 2 Refer to Figs. 23.

(a) In SOM-cre mice AAV-DIO-Chrimson and AAV-DIO-GCaMP were injected in the mPFC to record and activate SOM neurons in the mPFC and their terminals in the RSC. (b) Activation of the soma induces an increase of the signal in both the mPFC and the RSC while (c) stimulation of the terminals results in an increased GCAmP signal only in the terminals but not in the soma. (d) SOM-cre mice receiving inhibition (red, AAV-DIO-NphR) or control (black, AAV-DIO-TdTomato) of the mPFC-to-RSC terminals when in the stress demonstrator zone, and sacrificed 30 min following the task for cFOS immunohistochemistry. (e) Ex vivo quantification of cFOS positive neurons in the RSC (n = 6 mice, 8 to 13 sections per mice within the PFC, *P < 0.0001). (f) SOM-cre mice (n = 4 males) were injected with AAV-DIO-Chrimson in the mPFC, receiving retroorbital administration of PHPeb-CAG-GCamP and implanted with a widefield miniscope above the RSC, while control mice (WT n = 2) received only the retroorbital administration of PHPeb-CAG-GCamP and implanted with a widefield miniscope above the RSC. (g) GCaMP-activity in the RSC during activation (5 Hz, 2 seconds long) of mPFC-to-RSC SOM terminals (red, n = 386 stimulations, *P < 0.0001 2 sec light period versus 5 sec baseline). Control mice were treated in the same way but not having injected the chrimson opsin (gray line, n = 100 stimulation). Left panel represent the average GCamP-Signal during the baseline (5 s) and the stimulation (2 s) across all stimulation trials. The cake graph shows that in opsin injected animals, 91.5% of the stimulation pulses delivered induced a reduction in the signal (353/386 stim) while a minor part was associated with an increase (20/386, 5.2%) or no significant changes (13/386, 3.4%). (h) SOM-cre mice injected in the mPFC with AAV-DIO-ChR2 (n = 8 mice) or AAV-DIO-NphR (n = 14), and tested for discrimination between an object and a mouse without light, with optostimulation (blue background), or photoinhibition (orange background), (*P < 0.05 mouse versus object). Light stimulation was delivered only during the first 2 minutes of the test. (i) Discrimination between a familiar and an unfamiliar mouse (n = 11) without light, with optostimulation (blue background) or photoinhibition (orange background), (*P < 0.05 familiar versus unfamiliar). Light stimulation was delivered during the first 2 minutes of the test. (j) Distance travelled during the emotion recognition with photoactivation (blue background) or photoinhibition (orange background, n = 6). (k-l) SOM-cre mice injected with AAV-DIO-YFP (Control, n = 6) and implanted with a LED on the RSC to deliver blue (blue background) or red light (orange background) during the first 2 minutes of the emotion recognition task. (*P < 0.05 neutral versus emotionally altered). (m) In all tested mice, most social contacts (95.95 + /-0.71%) between the observer and a demonstrator mouse occurred with entries of ≥2.5 s from crossing the stimulus mouse zone (n = 24 mice). (n) Calcium signal in WT mice (n = 9) injected with AAV-CamKII-GCamP in the RSC or SOM cre mice (same as Fig.3d) injected with AAV-DIO-Chrimson in the mPFC, before, during and after the delivery of red light (560 nm). Bar plots are represented as mean ± SEM. Whisker-plot presenting upper and lower quartile, mean and median. Complete statistics in Supplementary Table 3.

Extended Data Fig. 3 Refer to Figs. 2 and 7.

Optogenetic inhibition (NphR) and stimulation (ChR2) of mPFC-to-RSC SOM terminals during (a) stress emotion recognition in male (n = 5 for ChR2 and 6 for NphR) and female (n = 4 for ChR2 and 8 for NphR) observers, or (b) during relief emotion recognition in male (n = 5 for ChR2 and 6 for NphR) and female (n = 4 for ChR2 and 8 for NphR) observers (*P < 0.05 versus no light). (c) Fiber photometry signal of mPFC-to-RSC SOM terminals during emotion recognition when entering the neutral mouse zone (grey lines) or the zone of emotionally altered demonstrators (red lines) in all animals, or divided by males and females (n = 23; #P < 0.0001 versus baseline; *P < 0.005 versus neutral). Whisker-plot are presenting upper and lower quartile, mean and median. (d) Optogenetic ChR2 stimulation of the mPFC-to-RSC SOM terminals of male (n = 5) and female (n = 5) observers in neutral versus neutral recognition with entries to a randomly selected neutral mouse zone paired with 5 Hz stimulation (Stim). (e) Control group with no stimulation of the mPFC-to-RSC SOM terminals of male (n = 5) and female (n = 5) observers in neutral versus neutral recognition. (f) Observer mice tested with two neutral demonstrators. Entrance in one of these demonstrators (randomly chosen) triggered optogenetic inhibition (NphR) of the mPFC-to-RSC SOM terminals (continuous light). (*P < 0.05 neutral + stim versus neutral). (g) Dual stimulation of the mPFC-to-RSC SOM terminals in neutral versus neutral condition with entries to a randomly selected demonstrator paired with a short activation (200 ms, 5 Hz) follow by inhibition (continuous). (*P < 0.05 neutral + dual stim versus neutral). (h) Stress emotion recognition in ARC mice (+/+, + /- and -/-) in males (+/+ n = 10, +/- n = 9, -/- n = 9) and females (+/+ n = 5, +/- n = 6, -/- n = 4). (*P < 0.05 ARC + /+ neutral versus ARC + /+ stress). (i) Relief emotion recognition in ARC mice (+/+, + /- and -/-) in males and female (*P = 0.01 ARC + /+ neutral versus ARC+ /+ relief). (j) Emotion recognition with no PCP (PCP day 0), following acute PCP (PCP day 1) and 14 days PCP + 7 days washout (PCP day 21) in males (n = 5) and females (n = 6). Bar plots are represented as mean ± SEM. Whisker-plot presenting upper and lower quartile, mean and median. * is used for P < 0.05. Complete statistics in Supplementary Table 3.

Extended Data Fig. 4 Refer to Figs. 2 and 4.

(a-b) Number of body rotation, head-tail distance, and number of entries without light and following optogenetic inhibition of mPFC-to-RSC SOM terminals (NphR-terminals) during recognition for (a) stress or (b) relief (*P < 0.0001 no light versus light). (c-d) Number of body rotation, head-tail distance, and number of entries without light and following optogenetic stimulation of mPFC-to-RSC SOM terminals (ChR2-terminals) during recognition for (c) stress or (d) relief (*P < 0.001 no light versus light). (e-f) Number of body rotation, head-tail distance, and number of entries without light and following optogenetic stimulation of RSC-pyramidal neurons during recognition for (e) stress or (f) relief (*P < 0.05 no light versus light). The increase in the number of entries in both the emotionally altered is associated with a reduction of social-interaction (Fig. 2) and less sniffing (see Fig. 4b-c). (g-h) Time spent in the no demonstrator mice zone subdivided by central area or close to the walls without light or following optogenetic stimulation of RSC-pyramidal neurons in (g) stress or (h) relief emotion recognition (Figures show only 0-2 min for clarity). (i-j) Time spent in the no demonstrator mice zone subdivided by central area or close to the walls without light or following optogenetic stimulation of mPFC-to-RSC SOM terminals in (i) stress or (j) relief emotion recognition. (k-l) Time spent in the no demonstrator mice zone subdivided by central area or close to the walls without light or following optogenetic inhibition of mPFC-to-RSC SOM terminals in (k) stress or (l) relief emotion recognition. Together with increased entrance (Extended Data Fig. 9c-d), these data suggest that opto inhibition of mPFC-to-RSC SOM terminals increase mouse willingness to interact with other mice, consistent with its increased emotion recognition effect (Fig. 2), unrelated to anxiety-like behaviors (effects both in center and walls time). (m) Grooming time in no stimulated side versus the side with optogenetic stimulation of RSC pyramidal neurons in a neutral versus neutral condition. (n) Grooming time in no stimulated side versus the side with optogenetic stimulation of mPFC-to-RSC SOM terminals in a neutral versus neutral condition. (o) Grooming time in no stimulated side versus the side with optogenetic inhibition of mPFC-to-RSC SOM terminals in a neutral versus neutral condition. Bar plots are represented as mean ± SEM. * is used for P < 0.05. Complete statistics in Supplementary Table 4.

Extended Data Fig. 5 Refer to Figs. 234.

Photometry recording of mPFC-to-RSC SOM terminals during recognition (a) between an object and a novel mouse (n = 4; object: green; mouse: grey. n = 25 entries to object, n = 34 entries to neutral) or (b) between a familiar and unfamiliar mouse (n = 29 entries to unfamiliar, n = 18 entries to familiar). (c-d) Photometry recording in SOM-cre mice (n = 6) injected with AAV-DIO-YFP in the mPFC and recorded in RSC during recognition (c) for stress (; stress: purple; neutral: grey; n = 199 entries to stress, n = 97 entries to neutral) or (d) relief (n = 198 entries to relief, n = 194 entries to neutral). (e) SOM-Cre mice (n = 4) were injected with AAV-DIO-GCAmP in the mPFC, and implanted with optic fibers both in the mPFC and RSC to record calcium activity of SOM cells at the level of the soma (Soma) or their terminals in the RSC (Term), allowing a direct correlation of both signals. (f) Calcium events in the Soma were selected as the threshold for the peristimulus analyses, showing that GCamP-signal increased in the mPFC (grey line), while presenting two patterns in the RSC, either an increase in 45% of the events (blue line) or a decrease (red line). (g) Som-cre mice were injected in the mPFC with AAV-DIO-ChR2 and AAV-DIO-NphR (n = 5 mice) and bilateral optic fiber was put on top of the RSC to test if activation or inhibition of mPFC-to-RSC SOM terminals could induce side preference in the absence of stimulus mice (empty) which are normally neutral in their side bias. (h) Short (250 ms) inhibition of the terminals (entire 6 minutes, n = 5 mice) prior to entry to a randomly selected side. (i) Activation of the terminals (n = 5 mice) only when in a randomly selected side zone. (j) Inhibition of the terminals (n = 5 mice) when in a randomly selected side zone. (k) Activation of the terminals (entire 6 minutes, n = 5 mice) prior to entry to a randomly selected side. (l) Number of entries in each zone during no stimulation (black bars), activation (blue bars) or inhibition (red bars) (n = 5 mice). For experiment empty-empty all mice were tested in all stimulation conditions. Whisker-plot are presenting upper and lower quartile, mean and median. Bar plots are represented as mean ± SEM. * is used for P < 0.05. Complete statistics in Supplementary Table 4.

Extended Data Fig. 6 Refer to Fig. 3.

(a) Schematic of the experiment in cortical slices containing mPFC. The blue oval indicates ChR2-positive SOM neurons. The gray triangle indicates ChR2-negative excitatory cells. (b) Representative current-clamp patch-clamp recording of a ChR2-negative excitatory cell showing the membrane potential responses to current injections (-50, +250, +300, +350 pA). (c) Left: five representative current traces (black) and their average (red) recorded at –50 mV in voltage-clamp configuration from a ChR2-negative excitatory neuron during photoactivation of ChR2-positive SOM cells in PFC. Currents are recorded in the presence of antagonists of glutamatergic ionotropic receptors (see methods for details). Laser power: 3.5 mW. Light pulse duration: 10 ms (top) and 2 ms (bottom). Right: same as in left using 15 mW of laser power. Note that current amplitude evoked with 10 ms long of blue light at 3.5 mW (top left) and 15 mW (top right) was similar. (d) Top left: schematic of the experiment in cortical slices containing RSC. The gray triangle indicates a ChR2-negative excitatory cell. The blue lines indicate ChR2-positive axonal projections from SOM interneurons in mPFC. Bottom left: representative current-clamp patch-clamp recording of a ChR2-negative excitatory cell showing the membrane potential responses to current injections (-50, +100, +150, +200, +250 pA). Right: representative current traces recorded at – 50 mV in voltage-clamp configuration from a RSC excitatory cell during 10 ms or 2 ms of blue light stimulation at 22 mW. No currents were recorded from this as well from other 9 excitatory cells in RSC during photoactivation of ChR2-positive axons originating from SOM neurons located in mPFC. (e) Representative current traces recorded at +10 mV in voltage-clamp configuration from a RSC excitatory cell during 10 ms or 2 ms of blue light stimulation at 15 mW. Bar plots are represented as mean ± SEM. * is used for P < 0.05. Complete statistics in Supplementary Table 4.

Extended Data Fig. 7 Refer to Fig. 4.

(a-b) Wild type mice (n = 6 males) were injected with AAV-CamKII-GCamP6 in the RSC to allow the recording of GCamP signal from RSC pyramidal cells during recognition (a) for stress (stress: purple; neutral: grey; #P < 0.001 versus baseline; *P = 0.03 neutral versus stress; n = 77 entries to stress, n = 57 entries to neutral) and (b) relief (relief: yellow; neutral: grey; #P = 0.02 relief versus baseline; n = 122 entries to relief, n = 120 entries to neutral). Isosbetic signal was collected on the red wavelength to remove artifact and movement control. (c-d) Wild-type mice (n = 6 males) injected bilaterally in the RSC with AAV-CamKII-ChR2 with light on or off for the first 2 minutes of the (c) stress and (d) relief emotion recognition test (*P < 0.05 versus no stimulation). (e) Same stimulation protocol, but this time in the second two minutes interval of the task (*P < 0.05 stress light off versus all other groups). (f-i) WT mice (n = 12 females) were bilaterally injected in the RSC with either AAV-CamKII-NphR (orange, n = 6 females) or AAV-CamKII-ChR2 (blue, n = 6 females), and after at least 4 weeks a 560 nm continuous light or 450 nm 5 Hz light was delivered during the first two minutes of the object/mouse and familiar/unfamiliar recognition task for inhibition or activation, respectively, of RSC pyramidal neurons. Time (in seconds) spent sniffing a mouse/familiar (gray bars), or an unfamiliar (violet bars) or object (green bars) displayed by the same observer mice without light stimulation (first column), with photoinhibition (second column, orange background) or photostimulation (third column, blue background) performed for 2 minutes from the beginning of the test (f, *P < 0.05 object versus mouse) (h, *P < 0.01 familiar versus unfamiliar). (g,i) A direct comparison of the no light (gray circles), photoinhibition (orange triangle), and photostimulation (blue squares) conditions by the discrimination index (*P < 0.05 versus ChR2). (j-m) These same mice were then tested with optogenetic stimulation in an empty versus empty discrimination (n = 6 mice) (P > 0.05 stimulated side versus no stimulated side). Empty-side paired stimulation had similarly no effects in (k) the total distance traveled (P = 0.59), (l) the number of zones entries (P = 0.59), and (m) the time spent in the zone (P = 0.99). (n-q) When these same mice were tested in these same conditions but in the presence of two neutral mice (one being paired to 5 Hz stimulation), we found no significant effect (n) on the time spent in the zone (P = 0.11), (o) the distance traveled (P = 0.09), and (q) the time in each zone (P = 0.08), but a significant effect of the stimulation in (p) the number of entries (P = 0.01 stimulated versus no stimulated side). The same animals were tested in both empty versus empty and Neutral versus Neutral. Bar plots are represented as mean ± SEM. Complete statistics in Supplementary Table 5.

Extended Data Fig. 8 Refer to Fig. 4.

(a) Following injection in the RSC of AAV-CamKII-YFP in wild-type mice (n = 3), we found YFP-positive signal in different brain regions including the medial prefrontal cortex (mPFC), the dorsomedial striatum (DMS), the cingulate (CG), the contralateral retrosplenial cortex (RSC), and the lateroposterior thalamic nuclei (LP).

Extended Data Fig. 9 Refer to Fig. 6.

(a) Seed-based connectivity mapping of the right (upper row) and left (bottom row) somatosensory cortex (SS) in ARC + /+ and ARC-/- littermates. Red-yellow represents brain regions showing significant rsfMRI functional connectivity with the right and left SS cortex in control  ARC+ /+ (left) and ARC -/- mice (middle). Seed regions are depicted in green. No changes of functional connectivity of the right and left SS were identified in ARC-/- mutants compared to  ARC+ /+ mice (right). (b) Unimpaired rsfMRI connectivity between right and left SS cortex was also quantified in reference volumes of interest (t = 0.70, P = 0.49) (n = 11 ARC+/+, n = 11 ARC-/-, unpaired two-tailed T-test). (c) SOM- ARC+ /- mice (n = 8) injected with AAV-DIO-NphR in the mPFC and implanted with a LED above the RSC present no emotion recognition in absence of light stimulation (paired two-tailed T-test; 0-2 min: P = 0.40 neutral versus stress; 2-4 min: P = 0.11 neutral versus stress; 4-6 min: P = 0.38 neutral versus stress). Bar plots are represented as mean ± SEM. * is used for P < 0.05. (d) Graphical resume of the photometry results obtained in the PCP-treated animals (n = 11 mice) during no drug condition (No PCP), acute (PCP day1) and chronic (PCP day 21) treatments. All same animals were tested in no PCP, PCP day 1 and PCP day 21. Arrow represents the variation of GCamP signal during entries to the stress (purple) or the neutral (grey) demonstrators. The thickness of the arrow is used for the relative strength of the signal variation.

Extended Data Table 1 Refer to Fig 6

Supplementary information

Supplementary Information

Supplementary Tables 1–5. Detailed statistical analyses for Figs. 1–5 and Extended Data Figs. 2–7.

Reporting Summary

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

Dautan, D., Monai, A., Maltese, F. et al. Cortico-cortical transfer of socially derived information gates emotion recognition. Nat Neurosci (2024). https://doi.org/10.1038/s41593-024-01647-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1038/s41593-024-01647-x

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