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:

The locus coeruleus drives disinhibition in the midline thalamus via a dopaminergic mechanism

Abstract

The paraventricular nucleus of the thalamus (PVT) is increasingly being recognized as a critical node linking stress detection to the emergence of adaptive behavioral responses to stress. However, despite growing evidence implicating the PVT in stress processing, the neural mechanisms by which stress impacts PVT neurocircuitry and promotes stressed states remain unknown. Here we show that stress exposure drives a rapid and persistent reduction of inhibitory transmission onto projection neurons of the posterior PVT (pPVT). This stress-induced disinhibition of the pPVT was associated with a locus coeruleus-mediated rise in the extracellular concentration of dopamine in the midline thalamus, required the function of dopamine D2 receptors on PVT neurons, and increased sensitivity to stress. Our findings define the locus coeruleus as an important modulator of PVT function: by controlling the inhibitory tone of the pPVT, it modulates the excitability of pPVT projection neurons and controls stress responsivity.

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: Synaptic inhibition in the pPVT rapidly decreases following stress.
Fig. 2: Activation of D2-like dopamine receptors both mimics and is required for stress-induced downregulation of inhibitory transmission.
Fig. 3: Postsynaptic mechanisms contribute to stress-induced inhibitory plasticity in the pPVT.
Fig. 4: The LC drives stress-evoked dopamine release in the pPVT.
Fig. 5: The LC controls stress-induced inhibitory plasticity in the pPVT.
Fig. 6: LC input to the pPVT amplify neural responses to stress.
Fig. 7: Optogenetic stimulation of LC terminals in the pPVT before stress enhances aversive memory formation.

Similar content being viewed by others

References

  1. Vertes, R. P., Linley, S. B. & Hoover, W. B. Limbic circuitry of the midline thalamus. Neurosci. Biobehav. Rev. 54, 89–107 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  2. Hsu, D. T., Kirouac, G. J., Zubieta, J. K. & Bhatnagar, S. Contributions of the paraventricular thalamic nucleus in the regulation of stress, motivation, and mood. Front. Behav. Neurosci. 8, 73 (2014).

    PubMed  PubMed Central  Google Scholar 

  3. Penzo, M. A. et al. The paraventricular thalamus controls a central amygdala fear circuit. Nature 519, 455–459 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Do-Monte, F. H., Quiñones-Laracuente, K. & Quirk, G. J. A temporal shift in the circuits mediating retrieval of fear memory. Nature 519, 460–463 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Zhu, Y., Wienecke, C. F., Nachtrab, G. & Chen, X. A thalamic input to the nucleus accumbens mediates opiate dependence. Nature 530, 219–222 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Do-Monte, F. H., Minier-Toribio, A., Quiñones-Laracuente, K., Medina-Colón, E. M. & Quirk, G. J. Thalamic regulation of sucrose seeking during unexpected reward omission. Neuron 94, 388–400.e4 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Arcelli, P., Frassoni, C., Regondi, M. C., De Biasi, S. & Spreafico, R. GABAergic neurons in mammalian thalamus: a marker of thalamic complexity? Brain Res. Bull. 42, 27–37 (1997).

    Article  CAS  PubMed  Google Scholar 

  8. Zhang, X. & van den Pol, A. N. Rapid binge-like eating and body weight gain driven by zona incerta GABA neuron activation. Science 356, 853–859 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Betley, J. N., Cao, Z. F., Ritola, K. D. & Sternson, S. M. Parallel, redundant circuit organization for homeostatic control of feeding behavior. Cell 155, 1337–1350 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Giber, K. et al. A subcortical inhibitory signal for behavioral arrest in the thalamus. Nat. Neurosci. 18, 562–568 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Bodor, A. L., Giber, K., Rovó, Z., Ulbert, I. & Acsády, L. Structural correlates of efficient GABAergic transmission in the basal ganglia-thalamus pathway. J. Neurosci. 28, 3090–3102 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Wanaverbecq, N. et al. Contrasting the functional properties of GABAergic axon terminals with single and multiple synapses in the thalamus. J. Neurosci. 28, 11848–11861 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Halassa, M. M. & Acsády, L. Thalamic inhibition: diverse sources, diverse scales. Trends Neurosci. 39, 680–693 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Del Castillo, J. & Katz, B. Quantal components of the end-plate potential. J. Physiol. (Lond.) 124, 560–573 (1954).

    Article  Google Scholar 

  15. Redman, S. Quantal analysis of synaptic potentials in neurons of the central nervous system. Physiol. Rev. 70, 165–198 (1990).

    Article  CAS  PubMed  Google Scholar 

  16. Stevens, C. F. Quantal release of neurotransmitter and long-term potentiation. Cell 72 Suppl, 55–63 (1993).

    Article  CAS  PubMed  Google Scholar 

  17. Goldstein, J. M., Jerram, M., Abbs, B., Whitfield-Gabrieli, S. & Makris, N. Sex differences in stress response circuitry activation dependent on female hormonal cycle. J. Neurosci. 30, 431–438 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  18. Ramikie, T. S. & Ressler, K. J. Mechanisms of sex differences in fear and posttraumatic stress disorder. Biol. Psychiatry 83, 876–885 (2018).

    Article  PubMed  Google Scholar 

  19. Grimley, J. S. et al. Visualization of synaptic inhibition with an optogenetic sensor developed by cell-free protein engineering automation. J. Neurosci. 33, 16297–16309 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Dong, X., Li, S. & Kirouac, G. J. Collateralization of projections from the paraventricular nucleus of the thalamus to the nucleus accumbens, bed nucleus of the stria terminalis, and central nucleus of the amygdala. Brain Struct. Funct. 222, 3927–3943 (2017).

    Article  PubMed  Google Scholar 

  21. Wimmer, R. D. et al. Thalamic control of sensory selection in divided attention. Nature 526, 705–709 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Dobbs, L. K. et al. Dopamine regulation of lateral inhibition between striatal neurons gates the stimulant actions of cocaine. Neuron 90, 1100–1113 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Floran, B., Floran, L., Sierra, A. & Aceves, J. D2 receptor-mediated inhibition of GABA release by endogenous dopamine in the rat globus pallidus. Neurosci. Lett. 237, 1–4 (1997).

    Article  CAS  PubMed  Google Scholar 

  24. Lalchandani, R. R., van der Goes, M. S., Partridge, J. G. & Vicini, S. Dopamine D2 receptors regulate collateral inhibition between striatal medium spiny neurons. J. Neurosci. 33, 14075–14086 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Li, S., Shi, Y. & Kirouac, G. J. The hypothalamus and periaqueductal gray are the sources of dopamine fibers in the paraventricular nucleus of the thalamus in the rat. Front. Neuroanat. 8, 136 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Kempadoo, K. A., Mosharov, E. V., Choi, S. J., Sulzer, D. & Kandel, E. R. Dopamine release from the locus coeruleus to the dorsal hippocampus promotes spatial learning and memory. Proc. Natl. Acad. Sci. USA 113, 14835–14840 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Takeuchi, T. et al. Locus coeruleus and dopaminergic consolidation of everyday memory. Nature 537, 357–362 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Gerfen, C. R., Paletzki, R. & Heintz, N. GENSAT BAC cre-recombinase driver lines to study the functional organization of cerebral cortical and basal ganglia circuits. Neuron 80, 1368–1383 (2013).

    Article  CAS  PubMed  Google Scholar 

  29. Abercrombie, E. D., Keller, R. W. Jr. & Zigmond, M. J. Characterization of hippocampal norepinephrine release as measured by microdialysis perfusion: pharmacological and behavioral studies. Neuroscience 27, 897–904 (1988).

    Article  CAS  PubMed  Google Scholar 

  30. Kirouac, G. J. Placing the paraventricular nucleus of the thalamus within the brain circuits that control behavior. Neurosci. Biobehav. Rev. 56, 315–329 (2015).

    Article  PubMed  Google Scholar 

  31. Moruzzi, G. & Magoun, H. W. Brain stem reticular formation and activation of the EEG. Electroencephalogr. Clin. Neurophysiol. 1, 455–473 (1949).

    Article  CAS  PubMed  Google Scholar 

  32. Cullinan, W. E., Herman, J. P., Battaglia, D. F., Akil, H. & Watson, S. J. Pattern and time course of immediate early gene expression in rat brain following acute stress. Neuroscience 64, 477–505 (1995).

    Article  CAS  PubMed  Google Scholar 

  33. Dallman, M. F. et al. Chronic stress-induced effects of corticosterone on brain: direct and indirect. Ann. NY Acad. Sci. 1018, 141–150 (2004).

    Article  CAS  PubMed  Google Scholar 

  34. Oh, W. C., Lutzu, S., Castillo, P. E. & Kwon, H. B. De novo synaptogenesis induced by GABA in the developing mouse cortex. Science 353, 1037–1040 10 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Chen, J. L. et al. Clustered dynamics of inhibitory synapses and dendritic spines in the adult neocortex. Neuron 74, 361–373 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Villa, K. L. et al. Inhibitory synapses are repeatedly assembled and removed at persistent sites in vivo. Neuron 90, 662–664 (2016).

    Article  CAS  PubMed  Google Scholar 

  37. Li, Y. et al. Orexins in the paraventricular nucleus of the thalamus mediate anxiety-like responses in rats. Psychopharmacology (Berl.) 212, 251–265 (2010).

    Article  CAS  Google Scholar 

  38. Takada, M., Campbell, K. J., Moriizumi, T. & Hattori, T. On the origin of the dopaminergic innervation of the paraventricular thalamic nucleus. Neurosci. Lett. 115, 33–36 (1990).

    Article  CAS  PubMed  Google Scholar 

  39. Uematsu, A. et al. Modular organization of the brainstem noradrenaline system coordinates opposing learning states. Nat. Neurosci. 20, 1602–1611 (2017).

    Article  CAS  PubMed  Google Scholar 

  40. Berridge, C. W. & Waterhouse, B. D. The locus coeruleus-noradrenergic system: modulation of behavioral state and state-dependent cognitive processes. Brain Res. Brain Res. Rev. 42, 33–84 (2003).

    Article  PubMed  Google Scholar 

  41. Martins, A. R. & Froemke, R. C. Coordinated forms of noradrenergic plasticity in the locus coeruleus and primary auditory cortex. Nat. Neurosci. 18, 1483–1492 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Schmitt, L. I. et al. Thalamic amplification of cortical connectivity sustains attentional control. Nature 545, 219–223 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Sherman, S. M. Thalamus plays a central role in ongoing cortical functioning. Nat. Neurosci. 19, 533–541 (2016).

    Article  CAS  PubMed  Google Scholar 

  44. Sherman, S.M. & Guillery, R.W. Exploring the Thalamus. (Academic Press, Cambridge, MA, 2001).

  45. McCall, J. G. et al. CRH engagement of the locus coeruleus noradrenergic system mediates stress-induced anxiety. Neuron 87, 605–620 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Bingham, B. et al. Early adolescence as a critical window during which social stress distinctly alters behavior and brain norepinephrine activity. Neuropsychopharmacology 36, 896–909 (2011).

    Article  CAS  PubMed  Google Scholar 

  47. Curtis, A. L., Lechner, S. M., Pavcovich, L. A. & Valentino, R. J. Activation of the locus coeruleus noradrenergic system by intracoerulear microinfusion of corticotropin-releasing factor: effects on discharge rate, cortical norepinephrine levels and cortical electroencephalographic activity. J. Pharmacol. Exp. Ther. 281, 163–172 (1997).

    CAS  PubMed  Google Scholar 

  48. Skirzewski, M. et al. ErbB4 signaling in dopaminergic axonal projections increases extracellular dopamine levels and regulates spatial/working memory behaviors. Mol. Psychiatry https://doi.org/10.1038/mp.2017.132 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgements

We thank T. Davidson (Stanford University) for his assistance in the design of the fiber photometry system and G. Augustine (Lee Kong Chian School of Medicine, Singapore) for the SuperClomeleon plasmid. This work was supported by the NIMH Intramural Research Program (M.A.P.), NICHD Intramural Research Program (A.B.), and NIH Grant MH107460 (H.-B.K.).

Author information

Authors and Affiliations

Authors

Contributions

B.S.B. performed anatomical and immunohistochemical studies, calcium and chloride imaging experiments, optogenetic experiments, and stereotaxic injections for all experiments. B.J.W. and M.A.P. performed electrophysiological experiments. M.S. performed microdialysis experiments and analyzed collected fractions via HPLC. Y.L. performed RT-PCR and in situ hybridization experiments. J.H.H. performed all procedures for two-photon imaging of gephyrin puncta. O.K. and N.R. performed stereotaxic surgeries and histological procedures. A.B. and H.-B.K. provided critical reagents and suggestions. B.S.B., B.J.W., and M.A.P. designed the study, interpreted results, and wrote the paper.

Corresponding author

Correspondence to Mario A. Penzo.

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 Inhibitory projections to the pPVT.

Compilation of data from the Mouse Brain Connectivity Atlas of the Allen Brain Institute (connectivity.brain-map.org) showing eight different areas of the mouse brain where a Cre-dependent AAV expressing GFP was injected (left panels), and fluorescently-labelled terminals were detected in the pPVT (right panels). Abbreviations: MPO, medial preoptic area; SI, substantia innominata; AHN, anterior hypothalamic nucleus; DMH, dorsomedial hypothalamus; ZI, zona incerta; PAG, periaqueductal gray; MRN, median raphe nucleus; PRNo, pontine reticular nucleus oral part; PRNc, pontine reticular nucleus caudal part. The following mouse Cre-lines were used for these experiments: GAD2-IRES-Cre (MPO, SI, AHN, DMH, ZI, PAG), Slc32a1-IRES-Cre (MRN/PRNo), and Slc6a5-Cre (PRNc).

Supplementary Figure 2 Most neurons of the pPVT express dopamine D2 receptors.

Representative images showing coincident expression of D2-EGFP and the neuronal marker NeuN in the pPVT of wildtype mice (colocalization: 70% D2+/NeuN, n = 3 mice).

Supplementary Figure 3 Sex differences do not account for stress-induced disinhibition of projection neurons of pPVT.

Quantification of mIPSC amplitude and frequency analyzed across both sexes (blue- males; pink- females). Amplitude in pA, naïve male, 49.93 ± 3.62, n = 17 neurons, 2 mice; naïve female, 49.08 ± 4.83, n = 14 neurons, 2 mice; two-sided t-test, non-significant, P = 0.89; restraint male, 39.05 ± 4.66, n = 7 neurons, 2 mice; restraint female, 36.64 ± 1.07, n = 17 neurons, 2 mice; two-sided t-test, non-significant, P = 0.48; footshock male, 41.93 ± 2.26, n = 21 neurons, 2 mice; footshock female, 40.76 ± 1.83, n = 30 neurons, 2 mice; two-sided t-test, non-significant, P = 0.69. Frequency in Hz, naïve male, 5.51 ± 1.06, n = 17 neurons; naïve female, 4.34 ± 0.73, n = 14 neurons, 2 mice; two-sided t-test, non-significant, P = 0.39; restraint male, 5.33 ± 1.33, n = 7 neurons, 2 mice; restraint female, 2.81 ± 0.39, n = 17 neurons, 2 mice; two-sided t-test, *P = 0.023; footshock male, 5.32 ± 0.89, n = 21 neurons, 2 mice; footshock female, 5.21 ± 0.92, n = 30 neurons, 2 mice; two-sided t-test, non-significant, P = 0.93. Data shown as mean ± s.e.m.

Supplementary Figure 4 SuperClomeleon reports GABAA receptor activity.

a. Schematic of the experimental design used for fiber photometry chloride imaging of NAc-projecting pPVT neurons. b. Representative image of SuperClomeleon expression in NAc-projecting neurons of the pPVT. c. Average SuperClomeleon response from NAc-projecting pPVT neurons in animals injected with the GABAA receptor agonist 4,5,6,7-tetrahydroisoxazolo(5,4-c) pyridin-3-ol (THIP, 8 mg/kg; green line) or saline (gray line; n = 3 mice). Data shown as mean ± s.e.m.

Supplementary Figure 5 Optical fiber placements for fiber photometry and optogenetic experiments.

a. Optical fiber placements for GCaMP6s experiment in Fig. 6a-d. b. Optical fiber placement for GCaMP6s experiment in Fig. 6e-h. c. Optical fiber placements for GCaMP6s experiment in Fig. 6i-l. d. Optical fiber placement for GCaMP6s experiment in Fig. 4h. e. Optical fiber placement for GCaMP6s experiment in Supplementary Fig. 8. f. Optical fiber placements for SuperClomeleon experiment in Fig. 1e-h. g. Optical fiber placement for SuperClomeleon experiment in Fig. 5a-d. h. Optical fiber placements for SuperClomeleon experiment in Fig. 5e-h. i. Optical fiber placement for SuperClomeleon experiment in Supplementary Fig. 3. j. Optical fiber placement for ChR2 experiment in Fig. 7. k. Optical fiber placement for halorhodopsin experiment in Fig. 6e-h. All circles depict the lowest position of the optical fibers for each subject.

Supplementary Figure 6 Robust expression of Drd2 and weak expression of Drd3 in the pPVT.

a. Fluorescent 2-plex in situ hybridization experiment showing the expression of Drd2 (red) and Drd3 (green) mRNA in the pPVT. b. Magnified view of the region depicted by the white square (a). 2-plex in situ hybridization experiments were independently repeated three times with samples collected from different subjects and similar results were obtained.

Supplementary Figure 7 Probe placements for microdialysis experiments.

a. Probe placements for microdialysis experiment in Fig. 4b. b. Representative image of hM4Di-mCherry expression in the LC of Dbh-Cre mice used for combined chemogenetic silencing of the LC and microdialysis of the PVT. Histological assessment of hM4Di-mCherry expression in the LC was independently repeated for each mouse included (seven total) and similar results were obtained. c. Probe placements for microdialysis experiment in saline treated mice (Fig. 4j). d. Probe placements for microdialysis experiment in CNO treated mice (Fig. 4j).

Supplementary Figure 8 TH+ afferents of the pPVT originate in the LC.

a. Representative image showing viral-assisted retrograde labelling of pPVT-projecting neurons in the ventral tegmental area (VTA). Antibody staining reveals TH-expressing neurons in VTA. Histological assessment of TH-expressing PVT-projecting neurons in the VTA was independently repeated for two mice and similar results were obtained. b. Magnified portion of the image shown in (a) depicts a single retrogradely-labelled neuron in VTA (arrow) that does not appear to be immunopositive for TH. c. Representative images from two separate experiments of viral-assisted retrograde labelling of pPVT-projecting neurons in the ventral tegmental area in LC. All neurons located within the LC displayed TH immunoreactivity. Histological assessment of TH-expressing PVT-projecting neurons in the LC was independently repeated for two mice and similar results were obtained.

Supplementary Figure 9 Stress-induced increase in extracellular NE in the pPVT is only partially dependent on the LC.

Schematic of the approach utilized for combined chemogenetic silencing of the LC and microdialysis of the pPVT (left). Summary plot depicting stress-induced increases in the extracellular concentration of NE following CNO (black) and saline vehicle (red) I.P. injection in mice expressing hM4Di in LC (n = 7 mice, per group). Data shown as mean ± s.e.m.

Supplementary Figure 10 Pharmacological blockade of the GABAA receptor increases the firing rate of neurons of the pPVT.

a. Sample traces showing the effect of the GABAA receptor blocker picrotoxin (PTX; 100 µM) on cell-attached recorded action potentials from two different pPVT neurons. b. Quantification of firing frequency in Hz (Baseline, 0.86 ± 0.23; PTX, 1.67 ± 0.41; n = 9 neurons, 3 mice, **P = 0.0082; two-sided Paired sample t-test). Data shown as mean ± s.e.m.

Supplementary Figure 11 Stress activates D2+ neurons of the pPVT.

a. Schematic of viral vector injections and optical fiber implantation for GCaMP6 fiber photometry experiments in Drd2-Cre mice. b. Representative image of GCaMP6s expression in D2+ neurons of the pPVT and optical fiber placement. Histological assessment of GCaMP6s expression in D2+ neurons of the pPVT was independently repeated for each mouse included (four total) and similar results were obtained. c. Average GCaMP6s response from D2+ pPVT neurons in animals subjected to footshock stress. Individual footshocks depicted by arrowheads. d. Average change in baseline fluorescence following footshock stress in %dF/F: Before, −0.68 ± 0.46; After, 6.71 ± 1.94; n = 4 mice; *P = 0.049; two-sided Paired sample t-test. Data shown as mean ± s.e.m.

Supplementary Figure 12 Most NAc-projecting neurons of the pPVT express D2 receptors.

a. Schematic of the stereotaxic injections for selectively expressing CTB in NAc-projecting neurons of the pPVT of D2-EGFP mice. b. Representative images showing coincident expression of D2-EGFP and CTB in the pPVT of wildtype mice (colocalization: 80% double-labelled/CTB, n = 4 mice).

Supplementary Figure 13 Tail suspension stress induces rapid activation of NAc-projecting neurons of the pPVT.

a. Schematic of the tail suspension protocol used during GCaMP6s imaging of NAc-projecting pPVT neurons. b. Representative trace showing the effect of brief tail suspensions (~10-12 s) on GCaMP6s fluorescence on a single subject. Arrowheads depict individual tail suspension events. c. Average GCaMP6s response from NAc-projecting pPVT neurons in animals subjected to tail suspension. Individual tail suspension events depicted by arrowheads. d. Average change in baseline fluorescence following tail suspension stress in %dF/F: Before, 0.24 ± 0.45; After, 9.88 ± 2.98; n = 4 mice; *P = 0.039; two-sided Paired sample t-test. Data shown as mean ± s.e.m.

Supplementary Figure 14 Changes in GCaMP6s fluorescence do not correlate with movement.

Correlation between the locomotion index (extracted from the FreezeFrame software, Actimetrics) and the GCaMP6s fluorescent signal (in %dF/F) for two individual subjects (n = 2 mice) from data shown in Fig. 6a-d. R-Square values are 0.008 (left) and 0.0004 (right).

Supplementary Figure 15 Optogenetic stimulation of LC–pPVT projections drives D2-receptor-mediated disinhibition in vitro.

a. Schematic of the experimental procedure employed to assess the effect of optogenetic stimulation of LC terminals onto NAc-projecting neurons of the pPVT. To elicit ChR2-evoked responses in the presence of TTX, 4-aminopyridine (4-AP, 100 µM) was added to the bath. b. Average plot depicting the effect of optogenetic stimulation of LC terminals with ChR2 on the average amplitude of mIPSC recorded from CTB-positive NAc-projecting neurons in the presence (% of Baseline, n = 9 neurons, 5 mice) or absence (% of Baseline, n = 8 neurons, 4 mice) of the D2-like blocker sulpiride (1 µM). c. Average plot depicting the effect of optogenetic stimulation of LC terminals with ChR2 on the average frequency of mIPSC recorded from CTB-positive NAc-projecting neurons in the presence (% of Baseline, n = 9 neurons, 5 mice) or absence (% of Baseline, n = 8 neurons, 4 mice) of the D2-like blocker sulpiride. Data shown as mean ± 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

Beas, B.S., Wright, B.J., Skirzewski, M. et al. The locus coeruleus drives disinhibition in the midline thalamus via a dopaminergic mechanism. Nat Neurosci 21, 963–973 (2018). https://doi.org/10.1038/s41593-018-0167-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-018-0167-4

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