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:

Divergent medial amygdala projections regulate approach–avoidance conflict behavior

Abstract

Avoidance of innate threats is often in conflict with motivations to engage in exploratory approach behavior. The neural pathways that mediate this approach–avoidance conflict are not well resolved. Here we isolated a population of dopamine D1 receptor (D1R)-expressing neurons within the posteroventral region of the medial amygdala (MeApv) in mice that are activated either during approach or during avoidance of an innate threat stimulus. Distinct subpopulations of MeApv-D1R neurons differentially innervate the ventromedial hypothalamus and bed nucleus of the stria terminalis, and these projections have opposing effects on investigation or avoidance of threatening stimuli. These projections are potently modulated through opposite actions of D1R signaling that bias approach behavior. These data demonstrate divergent pathways in the MeApv that can be differentially weighted toward exploration or evasion of threats.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Fig. 1: Identification of dopamine receptive neurons in the MeApv.
Fig. 2: MeApv-D1R neurons are activated during approach and avoidance to predator odor and robobug.
Fig. 3: Anatomical and synaptic connectivity of MeApv-D1R neurons.
Fig. 4: Activation of distinct MeApv-D1R pathways differentially drives approach and avoidance.
Fig. 5: Inhibition of distinct MeApv-D1R pathways differentially drives approach and avoidance.
Fig. 6: D1R signaling biases activation of MeA→BNST approach pathway.

Similar content being viewed by others

Data availability

Datasets supporting the findings in this study and custom codes used for imaging analysis are available from the corresponding author upon reasonable request.

References

  1. Elliot, A. J. The hierarchical model of approach-avoidance motivation. Motiv. Emot. 30, 111–116 (2006).

    Google Scholar 

  2. Choi, G. B. et al. Lhx6 delineates a pathway mediating innate reproductive behaviors from the amygdala to the hypothalamus. Neuron 46, 647–660 (2005).

    CAS  PubMed  Google Scholar 

  3. Burnett, C. J. et al. Hunger-driven motivational state competition. Neuron 92, 187–201 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Padilla, S. L. et al. Agouti-related peptide neural circuits mediate adaptive behaviors in the starved state. Nat. Neurosci. 19, 734–741 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Blanchard, D. C. & Blanchard, R. J. Innate and conditioned reactions to threat in rats with amygdaloid lesions. J. Comp. Physiol. Psychol. 81, 281–290 (1972).

    CAS  PubMed  Google Scholar 

  6. Martinez, R. C., Carvalho-Netto, E. F., Ribeiro-Barbosa, E. R., Baldo, M. V. & Canteras, N. S. Amygdalar roles during exposure to a live predator and to a predator-associated context. Neuroscience 172, 314–328 (2011).

    CAS  PubMed  Google Scholar 

  7. Olds, M. E. & Olds, J. Approach-avoidance analysis of rat diencephalon. J. Comp. Neurol. 120, 259–295 (1963).

    CAS  PubMed  Google Scholar 

  8. Dielenberg, R. A., Hunt, G. E. & McGregor, I. S. “When a rat smells a cat”: the distribution of Fos immunoreactivity in rat brain following exposure to a predatory odor. Neuroscience 104, 1085–1097 (2001).

    CAS  PubMed  Google Scholar 

  9. McGregor, I. S., Hargreaves, G. A., Apfelbach, R. & Hunt, G. E. Neural correlates of cat odor-induced anxiety in rats: region-specific effects of the benzodiazepine midazolam. J. Neurosci. 24, 4134–4144 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Chung, A. S., Miller, S. M., Sun, Y., Xu, X. & Zweifel, L. S. Sexual congruency in the connectome and translatome of VTA dopamine neurons. Sci. Reports 7, 11120 (2017).

    Google Scholar 

  11. Bromberg-Martin, E. S., Matsumoto, M. & Hikosaka, O. Dopamine in motivational control: rewarding, aversive, and alerting. Neuron 68, 815–834 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Vincenz, D., Wernecke, K. E. A., Fendt, M. & Goldschmidt, J. Habenula and interpeduncular nucleus differentially modulate predator odor-induced innate fear behavior in rats. Behav. Brain Res. 332, 164–171 (2017).

    PubMed  Google Scholar 

  13. Ipser, J. C., Kariuki, C. M. & Stein, D. J. Pharmacotherapy for social anxiety disorder: a systematic review. Expert. Rev. Neurother. 8, 235–257 (2008).

    CAS  PubMed  Google Scholar 

  14. Garcia, R. Neurobiology of fear and specific phobias. Learn. Mem. 24, 462–471 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Dong, H. W., Petrovich, G. D. & Swanson, L. W. Topography of projections from amygdala to bed nuclei of the stria terminalis. Brain Res. Rev. 38, 192–246 (2001).

    CAS  PubMed  Google Scholar 

  16. Swanson, L. W. The amygdala and its place in the cerebral hemisphere. Ann. NY Acad. Sci. 985, 174–184 (2003).

    PubMed  Google Scholar 

  17. Gross, C. T. & Canteras, N. S. The many paths to fear. Nat. Rev. Neurosci. 13, 651–658 (2012).

    CAS  PubMed  Google Scholar 

  18. Lein, E. S. et al. Genome-wide atlas of gene expression in the adult mouse brain. Nature 445, 168–176 (2007).

    CAS  PubMed  Google Scholar 

  19. Heusner, C. L., Beutler, L. R., Houser, C. R. & Palmiter, R. D. Deletion of GAD67 in dopamine receptor-1 expressing cells causes specific motor deficits. Genesis 46, 357–367 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  21. Zhang, T. Y., Chrétien, P., Meaney, M. J. & Gratton, A. Influence of naturally occurring variations in maternal care on prepulse inhibition of acoustic startle and the medial prefrontal cortical dopamine response to stress in adult rats. J. Neurosci. 25, 1493–1502 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Unger, E. K. et al. Medial amygdalar aromatase neurons regulate aggression in both sexes. Cell Reports 10, 453–462 (2015).

    CAS  PubMed  Google Scholar 

  23. Wang, F. et al. RNAscope: a novel in situ RNA analysis platform for formalin-fixed, paraffin-embedded tissues. J. Mol. Diagn. 14, 22–29 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Kollack-Walker, S., Don, C., Watson, S. J. & Akil, H. Differential expression of c-fos mRNA within neurocircuits of male hamsters exposed to acute or chronic defeat. J. Neuroendocrinol. 11, 547–559 (1999).

    CAS  PubMed  Google Scholar 

  25. Lin, D. et al. Functional identification of an aggression locus in the mouse hypothalamus. Nature 470, 221–226 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Liu, Y. X. et al. Psychological stress on female mice diminishes the developmental potential of oocytes: a study using the predatory stress model. PLoS One 7, e48083 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Hong, W., Kim, D. W. & Anderson, D. J. Antagonistic control of social versus repetitive self-grooming behaviors by separable amygdala neuronal subsets. Cell 158, 1348–1361 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Choi, J. S. & Kim, J. J. Amygdala regulates risk of predation in rats foraging in a dynamic fear environment. Proc. Natl Acad. Sci. USA 107, 21773–21777 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Ghosh, K. K. et al. Miniaturized integration of a fluorescence microscope. Nat. Methods 8, 871–878 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Chen, T. W. et al. Ultrasensitive fluorescent proteins for imaging neuronal activity. Nature 499, 295–300 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Pardo-Bellver, C., Cádiz-Moretti, B., Novejarque, A., Martínez-García, F. & Lanuza, E. Differential efferent projections of the anterior, posteroventral, and posterodorsal subdivisions of the medial amygdala in mice. Front. Neuroanat. 6, 33 (2012).

    PubMed  PubMed Central  Google Scholar 

  32. Petrovich, G. D., Canteras, N. S. & Swanson, L. W. Combinatorial amygdalar inputs to hippocampal domains and hypothalamic behavior systems. Brain Res. Rev. 38, 247–289 (2001).

    CAS  PubMed  Google Scholar 

  33. Boyden, E. S., Zhang, F., Bamberg, E., Nagel, G. & Deisseroth, K. Millisecond-timescale, genetically targeted optical control of neural activity. Nat. Neurosci. 8, 1263–1268 (2005).

    CAS  PubMed  Google Scholar 

  34. Blanchard, R. J. & Blanchard, D. C. Attack and defense in rodents as ethoexperimental models for the study of emotion. Prog. Neuropsychopharmacol. Biol. Psychiatry 13, Suppl, S3–S14 (1989).

    PubMed  Google Scholar 

  35. Chuong, A. S. et al. Noninvasive optical inhibition with a red-shifted microbial rhodopsin. Nat. Neurosci. 17, 1123–1129 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Tervo, D. G. et al. A designer aav variant permits efficient retrograde access to projection neurons. Neuron 92, 372–382 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Flagel, S. B. et al. A selective role for dopamine in stimulus-reward learning. Nature 469, 53–57 (2011).

    CAS  PubMed  Google Scholar 

  38. Berridge, K. C. & Robinson, T. E. What is the role of dopamine in reward: hedonic impact, reward learning, or incentive salience? Brain Res. Rev. 28, 309–369 (1998).

    CAS  PubMed  Google Scholar 

  39. Dong, H. W. & Swanson, L. W. Projections from bed nuclei of the stria terminalis, posterior division: implications for cerebral hemisphere regulation of defensive and reproductive behaviors. J. Comp. Neurol. 471, 396–433 (2004).

    PubMed  Google Scholar 

  40. Marowsky, A., Yanagawa, Y., Obata, K. & Vogt, K. E. A specialized subclass of interneurons mediates dopaminergic facilitation of amygdala function. Neuron 48, 1025–1037 (2005).

    CAS  PubMed  Google Scholar 

  41. Schultz, W. Multiple dopamine functions at different time courses. Annu. Rev. Neurosci. 30, 259–288 (2007).

    CAS  PubMed  Google Scholar 

  42. Pezze, M. A. & Feldon, J. Mesolimbic dopaminergic pathways in fear conditioning. Prog. Neurobiol. 74, 301–320 (2004).

    CAS  PubMed  Google Scholar 

  43. Yang, H. et al. Laterodorsal tegmentum interneuron subtypes oppositely regulate olfactory cue-induced innate fear. Nat. Neurosci. 19, 283–289 (2016).

    CAS  PubMed  Google Scholar 

  44. Silva, B. A. et al. Independent hypothalamic circuits for social and predator fear. Nat. Neurosci. 16, 1731–1733 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Hari Dass, S. A. & Vyas, A. Copulation or sensory cues from the female augment Fos expression in arginine vasopressin neurons of the posterodorsal medial amygdala of male rats. Front. Zool. 11, 42 (2014).

    PubMed  PubMed Central  Google Scholar 

  46. Bergan, J. F., Ben-Shaul, Y. & Dulac, C. Sex-specific processing of social cues in the medial amygdala. eLife 3, e02743 (2014).

    PubMed  PubMed Central  Google Scholar 

  47. Wang, L., Chen, I. Z. & Lin, D. Collateral pathways from the ventromedial hypothalamus mediate defensive behaviors. Neuron 85, 1344–1358 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Golden, S. A. et al. Persistent conditioned place preference to aggression experience in adult male sexually-experienced CD-1 mice. Genes Brain. Behav. 16, 44–55 (2017).

    CAS  PubMed  Google Scholar 

  49. Ferrari, P. F., van Erp, A. M., Tornatzky, W. & Miczek, K. A. Accumbal dopamine and serotonin in anticipation of the next aggressive episode in rats. Eur. J. Neurosci. 17, 371–378 (2003).

    CAS  PubMed  Google Scholar 

  50. Couppis, M. H. & Kennedy, C. H. The rewarding effect of aggression is reduced by nucleus accumbens dopamine receptor antagonism in mice. Psychopharmacology (Berl.) 197, 449–456 (2008).

    CAS  Google Scholar 

  51. Zhuang, X., Masson, J., Gingrich, J. A., Rayport, S. & Hen, R. Targeted gene expression in dopamine and serotonin neurons of the mouse brain. J. Neurosci. Methods 143, 27–32 (2005).

    CAS  PubMed  Google Scholar 

  52. Gore, B. B. & Zweifel, L. S. Genetic reconstruction of dopamine D1 receptor signaling in the nucleus accumbens facilitates natural and drug reward responses. J. Neurosci. 33, 8640–8649 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Sparta, D. R. et al. Construction of implantable optical fibers for long-term optogenetic manipulation of neural circuits. Nat. Protoc. 7, 12–23 (2011).

    PubMed  PubMed Central  Google Scholar 

  54. Resendez, S. L. et al. Visualization of cortical, subcortical and deep brain neural circuit dynamics during naturalistic mammalian behavior with head-mounted microscopes and chronically implanted lenses. Nat. Protoc. 11, 566–597 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Sanz, E. et al. Fertility-regulating Kiss1 neurons arise from hypothalamic POMC-expressing progenitors. J. Neurosci. 35, 5549–5556 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Weber, F. et al. Control of REM sleep by ventral medulla GABAergic neurons. Nature 526, 435–438 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Kim, J. G. et al. Leptin signaling in astrocytes regulates hypothalamic neuronal circuits and feeding. Nat. Neurosci. 17, 908–910 (2014).

    CAS  PubMed  Google Scholar 

  58. Chand, A. N., Galliano, E., Chesters, R. A. & Grubb, M. S. A distinct subtype of dopaminergic interneuron displays inverted structural plasticity at the axon initial segment. J. Neurosci. 35, 1573–1590 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Ting, J. T., Daigle, T. L., Chen, Q. & Feng, G. Acute brain slice methods for adult and aging animals: application of targeted patch clamp analysis and optogenetics. Methods Mol. Biol. 1183, 221–242 (2014).

    PubMed  PubMed Central  Google Scholar 

  60. Carter, M. E., Soden, M. E., Zweifel, L. S. & Palmiter, R. D. Genetic identification of a neural circuit that suppresses appetite. Nature 503, 111–114 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Soden, M. E. et al. Genetic isolation of hypothalamic neurons that regulate context-specific male social behavior. Cell Rep. 16, 304–313 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Stamatakis, A. M. & Stuber, G. D. Activation of lateral habenula inputs to the ventral midbrain promotes behavioral avoidance. Nat. Neurosci. 15, 1105–1107 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Zhou, P. et al. Efficient and accurate extraction of in vivo calcium signals from microendoscopic video data. eLife 7, e28728 (2018).

    PubMed  PubMed Central  Google Scholar 

  64. Pnevmatikakis, E. A. et al. Simultaneous denoising, deconvolution, and demixing of calcium imaging data. Neuron 89, 285–299 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Jimenez, J. C. et al. Anxiety cells in a hippocampal-hypothalamic circuit. Neuron 97, 670–683.e6 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank M. Soden for technical advice and assistance with slice electrophysiology and members of the Zweifel lab for scientific discussion. We thank C. Campos, A. Bowen and R. Palmiter for their assistance with calcium imaging studies. We also thank J. Allen for assistance in the production of AAV viral vectors. This work was funded by the US National Institutes of Health (P50MH10642 and R01MH094536 to L.S.Z).

Author information

Authors and Affiliations

Authors

Contributions

S.M.M. and L.S.Z designed the experiments. S.M.M. and L.S.Z wrote the manuscript. S.M.M. performed all viral injection surgeries, behavior experiments, and slice electrophysiology. D.M. analyzed all calcium imaging data. Behavioral analysis and histology was performed by S.M.M. and A.S. L.S.Z purified all viral vectors.

Corresponding author

Correspondence to Larry S. Zweifel.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

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

Integrated supplementary information

Supplementary Figure 1 MeApv D1R neurons are activated by multisensory innate fear stimuli.

(a) Behavioral schematic (b-c) quantification and histological images of Fos levels in MEApv D1R neurons during exposure to threat stimuli. Fos is significantly induced in MEAPV D1R neurons over controls when mice are exposed to predator odor (PO), robobug (RB), and resident-intruder threat stimuli (RI) (n = 3 mice, 1-way ANOVA F(3, 189) = 55.05, P < 0.0001, Bonferroni’s multiple comparisons test). Scale bar: 15 μm. (d) Total number of virally labeled MEApv-D1R neurons did not differ across assays (n = 3 mice, 1-way ANOVA P > 0.05 Bonferroni’s multiple comparisons test). (e) Total Fos was significantly higher following predator odor (PO), robobug (RB), and resident-intruder (RI) assays (n = 3 mice, 1-way ANOVA, F(3,14) = 12.31, P = 0.008). Center values represent mean and error bars represent s.e.m.

Supplementary Figure 2 Histological and behavioral characterization of calcium imaging experiments.

(a) Schematic confirming viral targeting of GCaMP6m and microendoscope placement (n = 4 mice). Out of 24 mice injected with GCaMP6m, 4 mice were successfully implanted with lenses and used for this study. (b) Top: schematic of viral constructs for GCamP6m and HM3-mCherry expression. Bottom: histology showing GCamP6m (green) and HM3 (red) targeting to MeApv of D1R-Cre mice with lens implanted above. Histological verification was replicated in all four experimental animals. Scale bar: 100 μm. (c-f) Distribution of behavioral epochs during approach and hide box behavior for robobug and predator odor assays. (g) Mean time spent during investigation or hide box for predator odor and robobug assays. (n = 4 mice). Center values represent mean, error bars represent s.e.m. (h-K) Comparison of average z-scored ∆F for avoidance cells vs. approach cells during periods of (h) avoidance or (i) investigation of the robobug, (j) avoidance and (k) investigation of predator odor (n = 4 mice; two-tailed Wilcoxon rank sum test; the box represents the 25th–75th percentiles with smallest and largest data points falling within a 1.5 inter quartile range below and above the box, respectively. Whiskers represent the minimum and maximum.). (i) Correlation between percentage of cells active during investigation of robobug and predator odor (n = 4 mice; Pearson correlation test). (j) Same as (i) for male and female odorant assays. (m-n) Overlap of hide box active cells (m) and approach cells (n) in predator odor and robobug assays.

Supplementary Figure 3 MeApv D1R neurons are activated during investigation of conspecific odorant.

(a) Top: probability of distribution of peak fluorescence for all cells active during investigation of conspecific odorant (dashed line represents start of active investigation of male odorant, n = 4 mice); Middle: heat plot of calcium responses to odorant for all the active cells aligned to an investigation (n = 4 mice); Bottom: average traces of all cells activated cells during an approach and investigation epoch ± SEM (n = 4 mice). (b) Selectivity of cells active in mice during investigation of conspecific odorants (n = 4 mice) relative to approach and avoidance investigation of robobug and predator odor. (c and d) Distribution of behavioral epochs during approach and hide box behavior for conspecific odorants. (e) Percent of activated cells was not correlated with investigation time. (n = 4 mice; Pearson correlation test). (f) Time spent in hide box or in active investigation in response to conspecific odorant (n = 4 mice). Center values represent mean and error bars represent s.e.m.

Supplementary Figure 4 Spatial distribution of electrophysiological recordings in VMH and BNST.

(a) Example image and schematic showing location of cells patched in BNST and corresponding responses for ChR2 functional connectivity experiments (n = 31 cells/5 mice). (b) Same as (a) but for VMH (n = 15 cells/ 4 mice). Scale bar: 100 μm. (c) Example trace showing delayed inhibitory response in BNST that was blocked by application of CNQX (delay ~9.1 ms).

Supplementary Figure 5 MeApv D1R pathways are activated by multisensory innate fear stimuli.

(a) Injection schematic, quantification and histological images of Fos levels in VMH and BNST MeApv-projecting neurons during exposure to threat stimuli. (b) More neurons in the MeApv were labeled by RetroBeads injected into the BNST than into the VMH (n = 12 mice, P = 0.0012, two-tailed unpaired t test). (c) Total Fos was significantly higher following predator odor (PO), robobug (RB), and resident-intruder (RI) assays following RetroBead injections into the BNST or VMH (n = 3 mice/group, 2-way ANOVA, effect of assay F(3,16) = 35.56, P < 0.0001). (d) There is significantly more Fos induction in VMH-projecting MeApv neurons compared with BNST-projecting MeApv neurons during predator odor and robobug exposure (n = 3 mice/group, **P = 0.002, *P = 0.0134, two-tailed unpaired t test). (e) Histology showing Fos expression in BNST and VMH projecting MeApv neurons during exposure to threat stimuli. Scale bar: 10 μm. For each behavioral assay, histology was verified in two other animals. Center values represent mean and error bars represent s.e.m.

Supplementary Figure 6 Confirmation of ChR2 viral targeting and optical cannula placement.

(a) Rang of spread of ChR2 viral coverage within MEApv for each of the three experimental groups (n = 51 mice total for all three groups) with example image. Scale bar: 100 μm. Location of optical fibers and ChR2 terminals within (b) VMH (c) BNST and (d) optical fiber placement in MeApv for cell body stimulation. Scale bar: 50 μm.

Supplementary Figure 7 Activation of VMH-, but not BNST-, projecting MeApv D1R neurons elicits real-time place-avoidance behavior.

(a) Representative real time place preference (RTPP) location plots. (b) Quantification of RTPP (right) shows that optogenetic activation of VMH-projecting MEApv D1R neurons reduces time spent in light paired chamber, but activation of BNST-projecting MEApv D1R neurons has no effect (2-way ANOVA F(2, 29) = 3.31, P = 0.05; Bonferroni’s multiple comparisons); alternatively shown as time spent in unpaired side subtracted from paired side (1-way ANOVA F(2, 35) = 4.295, P = 0.036, Tukey’s multiple comparisons) . (c) Total distance traveled was reduced in VMH stimulated versus BNST stimulated MEApv-D1R terminals in the RTPP assay (1-way ANOVA F(2, 29) = 5.766, P = 0.0082; Tukey’s multiple comparison test). Center values represent mean and error bars represent s.e.m.

Supplementary Figure 8 Global activation of MEApv D1R neurons biases approach over avoidance.

(a) Schematic of viral injection of AAV1-FLEX-ChR2-mCherry and optic fiber implant into MEApv of D1RCre/+ mice. (b) Optogenetic activation MeApv-D1R neurons increases approach to predator odor (n = 7 mice/group; Time spent in hide box: P = 0.0459, two-tailed unpaired t test; Latency to approach: *P = 0.0459, unpaired t test; Investigation time: P = 0.0804, unpaired t test; Frequency of investigations: *P = 0.0104, unpaired t test). (c) During exposure to robobug, optogenetic activation of MeAPV-D1R neurons results in significantly shorter latency to sniff predator odor (n = 7 mice/group; *P = 0.0186, two-tailed unpaired t test) and less time in hide box during exposure to predator odor (*P = 0.0186, unpaired t test); investigation of predator odor is not affected (P = 0.11, unpaired t test). (d) Optogenetic activation of MEAPV-D1R neurons does not significantly affect behavior during resident intruder assay. (n = 7, 11 mice/group; grooming: P = 0.4021; fighting: P = 0.1252; investigation: P = 0.9951; unpaired t test). (e) Optogenetic activation of MEAPV-D1R neurons does not significantly affect preference for side of chamber or locomotion during RTPP (n = 8, 10 mice; 2 way ANOVA F(1, 16) = 0.01, P = -0.9420, Bonferroni’s multiple comparison; P = 0.6869, two-tailed unpaired t test). Center values represent mean and error bars represent s.e.m.

Supplementary Figure 9 Confirmation of JAWS viral targeting and optical cannula placement.

(a) Injection site of AAV2-Retro-FLEX-JAWS-GFP into VMH and (b) bilateral cannula placement over MeApv with cell body transfection of JAWS in MeApv (n = 21 animals). Scale bar: 50 μm. (c) Injection site of AAV2.Retro.JAWS into BNST, and (d) bilateral cannula placement over MeApv with cell body transfection of JAWS in MeApv (n = 19 mice). Scale bar: 50 μm.

Supplementary Figure 10 Infusion of SKF 81,297 into MeApv.

(a) Confirmation of bilateral cannula placement over MeApv. (b) SKF 81,297 does not affect locomotion, as seen in both the predator odor and robobug assays (n = 6 mice). Scale bar: 100 μm. Center values represent mean and error bars represent s.e.m.

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Miller, S.M., Marcotulli, D., Shen, A. et al. Divergent medial amygdala projections regulate approach–avoidance conflict behavior. Nat Neurosci 22, 565–575 (2019). https://doi.org/10.1038/s41593-019-0337-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-019-0337-z

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