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:

Rabies screen reveals GPe control of cocaine-triggered plasticity

Abstract

Identification of neural circuit changes that contribute to behavioural plasticity has routinely been conducted on candidate circuits that were preselected on the basis of previous results. Here we present an unbiased method for identifying experience-triggered circuit-level changes in neuronal ensembles in mice. Using rabies virus monosynaptic tracing, we mapped cocaine-induced global changes in inputs onto neurons in the ventral tegmental area. Cocaine increased rabies-labelled inputs from the globus pallidus externus (GPe), a basal ganglia nucleus not previously known to participate in behavioural plasticity triggered by drugs of abuse. We demonstrated that cocaine increased GPe neuron activity, which accounted for the increase in GPe labelling. Inhibition of GPe activity revealed that it contributes to two forms of cocaine-triggered behavioural plasticity, at least in part by disinhibiting dopamine neurons in the ventral tegmental area. These results suggest that rabies-based unbiased screening of changes in input populations can identify previously unappreciated circuit elements that critically support behavioural adaptations.

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

Access options

Buy this article

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

Figure 1: Cocaine-induced changes to VTA neuron inputs.
Figure 2: Cocaine triggers increase in GPe-PV neuron activity and excitability.
Figure 3: Bidirectional modulation of rabies labelling by activity manipulations.
Figure 4: GPe-PV neuron activity is required for cocaine-induced LMS and CPP.
Figure 5: GPe-PV neurons disinhibit VTA-DA neurons.

Similar content being viewed by others

References

  1. 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)

    Article  CAS  PubMed  Google Scholar 

  2. Armbruster, B. N., Li, X., Pausch, M. H., Herlitze, S. & Roth, B. L. Evolving the lock to fit the key to create a family of G protein-coupled receptors potently activated by an inert ligand. Proc. Natl Acad. Sci. USA 104, 5163–5168 (2007)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  4. Ye, L. et al. Wiring and molecular features of prefrontal ensembles representing distinct experiences. Cell 165, 1776–1788 (2016)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Watabe-Uchida, M., Zhu, L., Ogawa, S. K., Vamanrao, A. & Uchida, N. Whole-brain mapping of direct inputs to midbrain dopamine neurons. Neuron 74, 858–873 (2012)

    Article  CAS  PubMed  Google Scholar 

  6. Weissbourd, B. et al. Presynaptic partners of dorsal raphe serotonergic and GABAergic neurons. Neuron 83, 645–662 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. 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 

  8. Schwarz, L. A. et al. Viral-genetic tracing of the input-output organization of a central noradrenaline circuit. Nature 524, 88–92 (2015)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  9. Lerner, T. N. et al. Intact-brain analyses reveal distinct information carried by SNc dopamine subcircuits. Cell 162, 635–647 (2015)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Lammel, S. et al. Input-specific control of reward and aversion in the ventral tegmental area. Nature 491, 212–217 (2012)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  11. Kalivas, P. W. & Volkow, N. D. The neural basis of addiction: a pathology of motivation and choice. Am. J. Psychiatry 162, 1403–1413 (2005)

    Article  PubMed  Google Scholar 

  12. Nestler, E. J. & Carlezon, W. A. Jr. The mesolimbic dopamine reward circuit in depression. Biol. Psychiatry 59, 1151–1159 (2006)

    Article  CAS  PubMed  Google Scholar 

  13. Albin, R. L., Young, A. B. & Penney, J. B. The functional anatomy of basal ganglia disorders. Trends Neurosci. 12, 366–375 (1989)

    Article  CAS  PubMed  Google Scholar 

  14. Yin, H. H. & Knowlton, B. J. The role of the basal ganglia in habit formation. Nat. Rev. Neurosci. 7, 464–476 (2006)

    Article  CAS  PubMed  Google Scholar 

  15. Hammond, C., Bergman, H. & Brown, P. Pathological synchronization in Parkinson’s disease: networks, models and treatments. Trends Neurosci. 30, 357–364 (2007)

    Article  CAS  PubMed  Google Scholar 

  16. Lüscher, C. & Malenka, R. C. Drug-evoked synaptic plasticity in addiction: from molecular changes to circuit remodeling. Neuron 69, 650–663 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Wickersham, I. R. et al. Monosynaptic restriction of transsynaptic tracing from single, genetically targeted neurons. Neuron 53, 639–647 (2007)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Smith, Y. & Bolam, J. P. The output neurones and the dopaminergic neurones of the substantia nigra receive a GABA-containing input from the globus pallidus in the rat. J. Comp. Neurol. 296, 47–64 (1990)

    Article  CAS  PubMed  Google Scholar 

  19. Zhou, F.-M. & Lee, C. R. Intrinsic and integrative properties of substantia nigra pars reticulata neurons. Neuroscience 198, 69–94 (2011)

    Article  CAS  PubMed  Google Scholar 

  20. Oliet, S. H. R., Malenka, R. C. & Nicoll, R. A. Bidirectional control of quantal size by synaptic activity in the hippocampus. Science 271, 1294–1297 (1996)

    Article  ADS  CAS  PubMed  Google Scholar 

  21. Raab-Graham, K. F., Radeke, C. M. & Vandenberg, C. A. Molecular cloning and expression of a human heart inward rectifier potassium channel. Neuroreport 5, 2501–2505 (1994)

    Article  CAS  PubMed  Google Scholar 

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

    Article  ADS  CAS  PubMed  Google Scholar 

  23. Stachniak, T. J., Ghosh, A. & Sternson, S. M. Chemogenetic synaptic silencing of neural circuits localizes a hypothalamus→midbrain pathway for feeding behavior. Neuron 82, 797–808 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Lammel, S., Ion, D. I., Roeper, J. & Malenka, R. C. Projection-specific modulation of dopamine neuron synapses by aversive and rewarding stimuli. Neuron 70, 855–862 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Miyamichi, K. et al. Dissecting local circuits: parvalbumin interneurons underlie broad feedback control of olfactory bulb output. Neuron 80, 1232–1245 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Rajasethupathy, P. et al. Projections from neocortex mediate top-down control of memory retrieval. Nature 526, 653–659 (2015)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  27. Gittis, A. H. et al. New roles for the external globus pallidus in basal ganglia circuits and behavior. J. Neurosci. 34, 15178–15183 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Pierce, R. C. & Kumaresan, V. The mesolimbic dopamine system: the final common pathway for the reinforcing effect of drugs of abuse? Neurosci. Biobehav. Rev. 30, 215–238 (2006)

    Article  CAS  PubMed  Google Scholar 

  29. Stepien, A. E., Tripodi, M. & Arber, S. Monosynaptic rabies virus reveals premotor network organization and synaptic specificity of cholinergic partition cells. Neuron 68, 456–472 (2010)

    Article  CAS  PubMed  Google Scholar 

  30. Miyamichi, K. et al. Cortical representations of olfactory input by trans-synaptic tracing. Nature 472, 191–196 (2011)

    Article  ADS  CAS  PubMed  Google Scholar 

  31. Yonehara, K. et al. Spatially asymmetric reorganization of inhibition establishes a motion-sensitive circuit. Nature 469, 407–410 (2011)

    Article  ADS  CAS  PubMed  Google Scholar 

  32. Wall, N. R ., De La Parra, M ., Callaway, E. M. & Kreitzer, A. C. Differential innervation of direct- and indirect-pathway striatal projection neurons. Neuron 79, 347–360 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Franklin, K. B. J. & Paxinos, G. The Mouse Brain in Stereotaxic Coordinates 4th edn (Academic, 2012)

  34. Bäckman, C. M. et al. Characterization of a mouse strain expressing Cre recombinase from the 3′ untranslated region of the dopamine transporter locus. Genesis 44, 383–390 (2006)

    Article  CAS  PubMed  Google Scholar 

  35. Taniguchi, H. et al. A resource of Cre driver lines for genetic targeting of GABAergic neurons in cerebral cortex. Neuron 71, 995–1013 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Hippenmeyer, S. et al. A developmental switch in the response of DRG neurons to ETS transcription factor signaling. PLoS Biol. 3, e159 (2005)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Madisen, L. et al. Transgenic mice for intersectional targeting of neural sensors and effectors with high specificity and performance. Neuron 85, 942–958 (2015)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Rothwell, P. E. et al. Autism-associated neuroligin-3 mutations commonly impair striatal circuits to boost repetitive behaviors. Cell 158, 198–212 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Shuen, J. A., Chen, M., Gloss, B. & Calakos, N. Drd1a-tdTomato BAC transgenic mice for simultaneous visualization of medium spiny neurons in the direct and indirect pathways of the basal ganglia. J. Neurosci. 28, 2681–2685 (2008)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Krashes, M. J. et al. Rapid, reversible activation of AgRP neurons drives feeding behavior in mice. J. Clin. Invest. 121, 1424–1428 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Xue, M., Atallah, B. V. & Scanziani, M. Equalizing excitation-inhibition ratios across visual cortical neurons. Nature 511, 596–600 (2014)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  42. Mosca, T. J. & Luo, L. Synaptic organization of the Drosophila antennal lobe and its regulation by the Teneurins. eLife 3, e03726 (2014)

    Article  PubMed  PubMed Central  Google Scholar 

  43. Gunaydin, L. A. et al. Natural neural projection dynamics underlying social behavior. Cell 157, 1535–1551 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Petreanu, L., Mao, T., Sternson, S. M. & Svoboda, K. The subcellular organization of neocortical excitatory connections. Nature 457, 1142–1145 (2009)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  45. Park, J. S. et al. Synthetic control of mammalian-cell motility by engineering chemotaxis to an orthogonal bioinert chemical signal. Proc. Natl Acad. Sci. USA 111, 5896–5901 (2014)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This study was supported by grants from the Howard Hughes Medical Institute (Hughes Collaborative Innovation Award), National Institutes of Health (R01-NS50835, PO1 DA008227, TR01-MH099647, F32-DA038913, K99-DC013059 and K99-DA041445), and the Stanford Neurosciences Institute.

Author information

Authors and Affiliations

Authors

Contributions

K.T.B. performed the majority of experiments and data analysis; C.K.K. assisted with fibre photometry experiments and data analysis with support from K.D; P.H. assisted with electrophysiological recordings and data analysis; L.W.H. performed surgeries for ChR2 stimulation and Fos counting; B.D.H. assisted with terminal inhibition experiments; T.J.M. assisted with assay design for puncta quantification; K.E.D. and S.N. provided technical support; K.T.B., L.L. and R.C.M. designed the experiments, interpreted the results and wrote the paper, which was edited by all authors.

Corresponding authors

Correspondence to Liqun Luo or Robert C. Malenka.

Ethics declarations

Competing interests

R.C.M. and K.D. are on the scientific advisory board of Circuit Therapeutics, Inc., a biotech dedicated to development of novel therapeutics for brain disorders. All other authors declare no competing financial interests.

Additional information

Reviewer Information Nature thanks P. Kenny, M. Wolf and the other anonymous reviewer(s) for their contribution to the peer review of this work.

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

Extended data figures and tables

Extended Data Figure 1 Changes to VTA-DA inputs induced by drugs of abuse.

a, Quantification of monosynaptic inputs to VTA-DA neurons labelled in animals receiving single dose administration of cocaine, amphetamine, nicotine, morphine, saline, or fluoxetine one day before injection of RVdG into the VTA. Data were combined to generate Fig. 1c. b, Sample images of GPe neurons labelled by RVdG and co-stained for parvalbumin. Pie graph shows proportion of labelled cells that co-stained for parvalbumin. c, Sample images of NAcMedS neurons labelled by RVdG in DAT-Cre;D1-tdTomato mice. Pie graph shows proportion of labelled cells that were D1+ as defined by presence of tdTomato (scale bars, 50 μm).

Extended Data Figure 2 Axonal projections of GPe-PV neurons.

a, AAV-FLExloxP-mGFP was injected into the GPe of Pvalb-Cre animals, and mGFP+ axons were quantified throughout the brain. b, Quantification of fraction of mGFP+ axons in the indicated brain regions. c, Sample image of mGFP+ axons in the ventral midbrain (scale bar, 500 μm). d, Quantification of fraction of mGFP+ axons in the indicated ventral midbrain brain regions. The schematics of the mouse brain in this figure were adapted from ref. 33. DLStr, dorsolateral striatum; DMStr, dorsomedial striatum.

Extended Data Figure 3 Inhibition of GPe-PV neuron activity modestly affects cocaine-induced locomotion.

a, Cre-dependent AAVs expressing YFP, hM4Di, Kir2.1 or TeTxLc were injected into the GPe of Pvalb-Cre animals. b, Quantification of effects of CNO on basal locomotion in animals expressing YFP or hM4Di (compared to YFP + saline: YFP + CNO, P = 0.36; hM4Di + CNO, P = 0.59). c, Quantification of basal locomotion during GPe-PV neuron inhibition (hM4Di, P = 0.54; Kir2.1, P = 0.66; TeTxLc, P = 0.27). d, Quantification of cocaine-induced locomotion during GPe-PV neuron inhibition (hM4Di, P = 0.37; Kir2.1, P = 0.12; TeTxLc, P = 0.002). The schematics of the mouse brain in this figure were adapted from ref. 33.

Extended Data Figure 4 Labelled GPe inputs to the VTA correlated with LMS.

a, AAV-FLExloxP-TC and AAV-FLExloxP-G were injected into the VTA of DAT-Cre mice. Eleven days later, animals were habituated for two days to an open field chamber, and given a drug injection the following day. RVdG was injected one day after the drug. Five days after RVdG injection, the animal was given a second injection of the same drug in the open field. b, Normalized labelled GPe inputs plotted against the relative locomotion in session 2 vs. session 1 for cocaine (n = 34), amphetamine (n = 5), nicotine (n = 5) and morphine (n = 5). Regression line is plotted for all drugs combined. ce, Labelled GPe inputs after a single dose of cocaine significantly correlated with LMS (c) but not total locomotion after the first (d) or second (e) dose of cocaine. f–h, Plots of labelled GPe inputs vs. LMS for amphetamine (f; 1 mg kg−1), nicotine (g; 0.5 mg kg−1), or morphine (h; 10 mg kg−1). The schematics of the mouse brain in this figure were adapted from ref. 33.

Extended Data Figure 5 Inhibition of the GPe prevents morphine LMS and CPP.

a, A Cre-dependent AAV expressing either YFP or hM4Di was injected into the GPe of Pvalb-Cre mice. b, c, Quantification of LMS (b; P = 0.022) and CPP (c; P = 0.0005) in animals in which YFP or hM4Di (activated by CNO) were expressed in GPe-PV neurons. The schematics of the mouse brain in this figure were adapted from ref. 33.

Extended Data Figure 6 Inhibition of GPe-PV neurons that project to the midbrain blocks cocaine-induced CPP and LMS.

a, CAV-FLExloxP-Flp was injected into the ventral midbrain, and AAV-FLExFRT-Kir2.1 or AAV-FLExFRT-YFP was injected into the GPe of Pvalb-Cre mice. b, c, Quantification of LMS (b; P = 0.019) and CPP (c; P = 0.0067) in animals in which YFP or Kir2.1 were expressed in GPe-PV neurons projecting to the ventral midbrain. The schematics of the mouse brain in this figure were adapted from ref. 33.

Extended Data Figure 7 GPe-PV neurons that project to the midbrain collateralize to multiple subcortical targets.

a, CAV-FLExloxP-Flp was injected into the ventral midbrain, and AAV-FLExFRT-mGFP was injected into in the GPe of Pvalb-Cre mice. b, Representative image of mGFP+ collaterals in the thalamus and subthalamic nucleus (STN) (scale bar, 500 μm). c, Quantification of projection fraction of collaterals to indicated target regions. The schematics of the mouse brain in this figure were adapted from ref. 33.

Extended Data Figure 8 Dopamine neuron activity is required for the development of LMS and CPP.

a, Breeding scheme for experiments. LSL, loxP stop loxP. b, c, Quantification of LMS (b; P = 0.001) and CPP (c; P = 0.005) in control animals or animals expressing hM4Di in dopamine neurons receiving CNO.

Extended Data Figure 9 Map of anatomical location of ventral midbrain cells from which whole cell recordings were made.

Individual dots indicate location of cells in which ChR2-evoked IPSCs due to ChR2 expression in GPe-PV neurons could be detected (connected) or not (not connected) in NAcLat-projecting (a) or NAcMed-projecting (b) VTA-DA cells, and SNr-GABA (c) or VTA-GABA (d) cells. The schematics of the mouse brain in this figure were adapted from ref. 33.

Extended Data Figure 10 GPe-PV neurons mediate their effects through SNr-GABA neurons.

a, Procedure to test LMS and CPP during SNr-GABA activation. b, c, Activating SNr-GABA neurons with CNO prevented LMS (b; P = 0.010) and CPP (c; P = 0.015). d, Injection strategy to test whether SNr-GABA neurons are downstream of GPe-PV neurons. e, f, While expression of Kir2.1 in GPe-PV neurons prevented LMS and CPP, this suppression was overcome by concurrent inhibition of SNr-GABA neurons (e; P = 0.035, f; P = 0.036). The schematics of the mouse brain in this figure were adapted from ref. 33.

Supplementary information

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Beier, K., Kim, C., Hoerbelt, P. et al. Rabies screen reveals GPe control of cocaine-triggered plasticity. Nature 549, 345–350 (2017). https://doi.org/10.1038/nature23888

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature23888

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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