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:

Cell-specific pallidal intervention induces long-lasting motor recovery in dopamine-depleted mice

Abstract

The identification of distinct cell types in the basal ganglia has been critical to our understanding of basal ganglia function and the treatment of neurological disorders. The external globus pallidus (GPe) is a key contributor to motor suppressing pathways in the basal ganglia, yet its neuronal heterogeneity has remained an untapped resource for therapeutic interventions. Here we demonstrate that optogenetic interventions that dissociate the activity of two neuronal populations in the GPe, elevating the activity of parvalbumin (PV)-expressing GPe neurons over that of Lim homeobox 6 (Lhx6)-expressing GPe neurons, restores movement in dopamine-depleted mice and attenuates pathological activity of basal ganglia output neurons for hours beyond stimulation. These results establish the utility of cell-specific interventions in the GPe to target functionally distinct pathways, with the potential to induce long-lasting recovery of movement despite the continued absence of dopamine.

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

Figure 1: Global GPe stimulation does not rescue movement in DD mice.
Figure 2: Selective stimulation of PV-GPe neurons rescues movement persistently in DD mice.
Figure 3: Local response during PV-ChR2 stimulation directly inhibits other high-firing GPe neurons.
Figure 4: Selective suppression of Lhx6-GPe neurons rescues movement persistently in DD mice.
Figure 5: The induction of persistent behavioral rescue is cell-type specific.
Figure 6: PV-ChR2 and Lhx6-Arch reverse pathological bursting activity persistently.

Similar content being viewed by others

References

  1. Saunders, A. et al. A direct GABAergic output from the basal ganglia to frontal cortex. Nature 521, 85–89 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Mastro, K.J., Bouchard, R.S., Holt, H.A. & Gittis, A.H. Transgenic mouse lines subdivide external segment of the globus pallidus (GPe) neurons and reveal distinct GPe output pathways. J. Neurosci. 34, 2087–2099 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Kita, H. Globus pallidus external segment. Prog. Brain Res. 160, 111–133 (2007).

    CAS  PubMed  Google Scholar 

  4. Bevan, M.D., Magill, P.J., Terman, D., Bolam, J.P. & Wilson, C.J. Move to the rhythm: oscillations in the subthalamic nucleus-external globus pallidus network. Trends Neurosci. 25, 525–531 (2002).

    CAS  PubMed  Google Scholar 

  5. Mallet, N. et al. Parkinsonian beta oscillations in the external globus pallidus and their relationship with subthalamic nucleus activity. J. Neurosci. 28, 14245–14258 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Bergman, H. et al. Physiological aspects of information processing in the basal ganglia of normal and parkinsonian primates. Trends Neurosci. 21, 32–38 (1998).

    CAS  PubMed  Google Scholar 

  7. Hernández, V.M. et al. Parvalbumin+ neurons and Npas1+ neurons are distinct neuron classes in the mouse external globus pallidus. J. Neurosci. 35, 11830–11847 (2015).

    PubMed  PubMed Central  Google Scholar 

  8. Dodson, P.D. et al. Distinct developmental origins manifest in the specialized encoding of movement by adult neurons of the external globus pallidus. Neuron 86, 501–513 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Abdi, A. et al. Prototypic and arkypallidal neurons in the dopamine-intact external globus pallidus. J. Neurosci. 35, 6667–6688 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Mallet, N. et al. Dichotomous organization of the external globus pallidus. Neuron 74, 1075–1086 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Hegeman, D.J., Hong, E.S., Hernández, V.M. & Chan, C.S. The external globus pallidus: progress and perspectives. Eur. J. Neurosci. 43, 1239–1265 (2016).

    PubMed  PubMed Central  Google Scholar 

  12. Oh, Y.M. et al. Using a novel PV-Cre rat model to characterize pallidonigral cells and their terminations. Brain Struct. Funct. http://dx.doi.org/10.1007/s00429-016-1346-2 (2016).

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

    CAS  PubMed  Google Scholar 

  14. DeLong, M.R. Primate models of movement disorders of basal ganglia origin. Trends Neurosci. 13, 281–285 (1990).

    CAS  PubMed  Google Scholar 

  15. Kravitz, A.V. et al. Regulation of parkinsonian motor behaviours by optogenetic control of basal ganglia circuitry. Nature 466, 622–626 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

  17. Herrera, A.J., Castaño, A., Venero, J.L., Cano, J. & Machado, A. The single intranigral injection of LPS as a new model for studying the selective effects of inflammatory reactions on dopaminergic system. Neurobiol. Dis. 7, 429–447 (2000).

    CAS  PubMed  Google Scholar 

  18. Mallet, N. et al. Arkypallidal cells send a stop signal to striatum. Neuron 89, 308–316 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Wang, J. et al. Coordinated reset deep brain stimulation of subthalamic nucleus produces long-lasting, dose-dependent motor improvements in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine non-human primate model of parkinsonism. Brain Stimul. 9, 609–617 (2016).

    PubMed  Google Scholar 

  20. Soares, J. et al. Role of external pallidal segment in primate parkinsonism: comparison of the effects of 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-induced parkinsonism and lesions of the external pallidal segment. J. Neurosci. 24, 6417–6426 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Rubin, J.E., McIntyre, C.C., Turner, R.S. & Wichmann, T. Basal ganglia activity patterns in parkinsonism and computational modeling of their downstream effects. Eur. J. Neurosci. 36, 2213–2228 (2012).

    PubMed  PubMed Central  Google Scholar 

  22. Gerfen, C.R. et al. D1 and D2 dopamine receptor-regulated gene expression of striatonigral and striatopallidal neurons. Science 250, 1429–1432 (1990).

    CAS  PubMed  Google Scholar 

  23. Smith, Y., Bevan, M.D., Shink, E. & Bolam, J.P. Microcircuitry of the direct and indirect pathways of the basal ganglia. Neuroscience 86, 353–387 (1998).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  25. Kühn, A.A., Kupsch, A., Schneider, G.H. & Brown, P. Reduction in subthalamic 8–35 Hz oscillatory activity correlates with clinical improvement in Parkinson's disease. Eur. J. Neurosci. 23, 1956–1960 (2006).

    PubMed  Google Scholar 

  26. Weinberger, M. et al. Beta oscillatory activity in the subthalamic nucleus and its relation to dopaminergic response in Parkinson's disease. J. Neurophysiol. 96, 3248–3256 (2006).

    PubMed  Google Scholar 

  27. Vitek, J.L., Zhang, J., Hashimoto, T., Russo, G.S. & Baker, K.B. External pallidal stimulation improves parkinsonian motor signs and modulates neuronal activity throughout the basal ganglia thalamic network. Exp. Neurol. 233, 581–586 (2012).

    PubMed  Google Scholar 

  28. Bernheimer, H., Birkmayer, W., Hornykiewicz, O., Jellinger, K. & Seitelberger, F. Brain dopamine and the syndromes of Parkinson and Huntington. Clinical, morphological and neurochemical correlations. J. Neurol. Sci. 20, 415–455 (1973).

    CAS  PubMed  Google Scholar 

  29. Riederer, P. & Wuketich, S. Time course of nigrostriatal degeneration in Parkinson's disease. A detailed study of influential factors in human brain amine analysis. J. Neural Transm. 38, 277–301 (1976).

    CAS  PubMed  Google Scholar 

  30. Betarbet, R., Sherer, T.B. & Greenamyre, J.T. Animal models of Parkinson's disease. Bioessays 24, 308–318 (2002).

    CAS  PubMed  Google Scholar 

  31. Deumens, R., Blokland, A. & Prickaerts, J. Modeling Parkinson's disease in rats: an evaluation of 6-OHDA lesions of the nigrostriatal pathway. Exp. Neurol. 175, 303–317 (2002).

    CAS  PubMed  Google Scholar 

  32. Fahn, S. Description of Parkinson's disease as a clinical syndrome. Ann. NY Acad. Sci. 991, 1–14 (2003).

    CAS  PubMed  Google Scholar 

  33. Tass, P.A. et al. Coordinated reset has sustained aftereffects in Parkinsonian monkeys. Ann. Neurol. 72, 816–820 (2012).

    PubMed  Google Scholar 

  34. Adamchic, I. et al. Coordinated reset neuromodulation for Parkinson's disease: proof-of-concept study. Mov. Disord. 29, 1679–1684 (2014).

    PubMed  PubMed Central  Google Scholar 

  35. Corbit, V.L. et al. Pallidostriatal projections promote β oscillations in a dopamine-depleted biophysical network model. J. Neurosci. 36, 5556–5571 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Holgado, A.J., Terry, J.R. & Bogacz, R. Conditions for the generation of beta oscillations in the subthalamic nucleus-globus pallidus network. J. Neurosci. 30, 12340–12352 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Erez, Y., Czitron, H., McCairn, K., Belelovsky, K. & Bar-Gad, I. Short-term depression of synaptic transmission during stimulation in the globus pallidus of 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-treated primates. J. Neurosci. 29, 7797–7802 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Bar-Gad, I., Elias, S., Vaadia, E. & Bergman, H. Complex locking rather than complete cessation of neuronal activity in the globus pallidus of a 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-treated primate in response to pallidal microstimulation. J. Neurosci. 24, 7410–7419 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Chin, G.D. & Hutchison, W.D. Effects of cobalt and bicuculline on focal microstimulation of rat pallidal neurons in vivo. Brain Stimul. 1, 134–150 (2008).

    PubMed  Google Scholar 

  40. Bugaysen, J., Bar-Gad, I. & Korngreen, A. The impact of stimulation induced short-term synaptic plasticity on firing patterns in the globus pallidus of the rat. Front. Syst. Neurosci. 5, 16 (2011).

    PubMed  PubMed Central  Google Scholar 

  41. Johnson, M.D., Zhang, J., Ghosh, D., McIntyre, C.C. & Vitek, J.L. Neural targets for relieving parkinsonian rigidity and bradykinesia with pallidal deep brain stimulation. J. Neurophysiol. 108, 567–577 (2012).

    PubMed  PubMed Central  Google Scholar 

  42. Chan, C.S. et al. HCN channelopathy in external globus pallidus neurons in models of Parkinson's disease. Nat. Neurosci. 14, 85–92 (2011).

    CAS  PubMed  Google Scholar 

  43. Hardman, C.D. & Halliday, G.M. The external globus pallidus in patients with Parkinson's disease and progressive supranuclear palsy. Mov. Disord. 14, 626–633 (1999).

    CAS  PubMed  Google Scholar 

  44. Gong, S. et al. Targeting Cre recombinase to specific neuron populations with bacterial artificial chromosome constructs. J. Neurosci. 27, 9817–9823 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. 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 

  46. Fogarty, M. et al. Spatial genetic patterning of the embryonic neuroepithelium generates GABAergic interneuron diversity in the adult cortex. J. Neurosci. 27, 10935–10946 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Gradinaru, V., Mogri, M., Thompson, K.R., Henderson, J.M. & Deisseroth, K. Optical deconstruction of parkinsonian neural circuitry. Science 324, 354–359 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Kravitz, A.V., Owen, S.F. & Kreitzer, A.C. Optogenetic identification of striatal projection neuron subtypes during in vivo recordings. Brain Res. 1511, 21–32 (2013).

    CAS  PubMed  Google Scholar 

  49. Willard, A.M., Bouchard, R.S. & Gittis, A.H. Differential degradation of motor deficits during gradual dopamine depletion with 6-hydroxydopamine in mice. Neuroscience 301, 254–267 (2015).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors thank V. Corbit (University of Pittsburgh) and T. Whalen (Carnegie Mellon University) for Matlab analysis scripts and B. Rogowski (Carnegie Mellon University) for surgical support and behavioral video editing. We also thank N. Kessaris (University of College London) and H. Zeng (Allen Institute) for their gifts of the Lhx6-iCre and Pvalb-2A-Cre mice, respectively. This work was supported by NIH grants F31 NS090745-01 (K.M.), F31 NS093944-01 (A.W.) and R00 NS076524, NSF grant DMS 1516288, and grants from the Brain and Behavior Research Foundation (National Alliance for Research on Schizophrenia and Depression Young Investigator Grant), the Parkinson's Disease Foundation, and the NIH Intramural Research Program.

Author information

Authors and Affiliations

Authors

Contributions

K.J.M., K.T.Z., and K.H.L. performed behavioral experiments, including the histological verification, and with A.H.G. analyzed the data. K.J.M. and A.M.W. were responsible for the collection and analysis of the in vivo experiments. All authors discussed results and interpretations. K.J.M. and A.H.G. designed the experiments and wrote the manuscript.

Corresponding author

Correspondence to Aryn H Gittis.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Behavioral and pathophysiological symptoms of bilateral DD are apparent within 3–5 d post-depletion.

(a) Schematic of bilateral DD in the medial forebrain bundle (MFB). (b) Quantification of immobility and bradykinesia induced by unilateral (Uni, n = 4) and bilateral (Bi, n = 51) depletions, as compared to dopamine-intact controls (Naive, n = 4). (*p < 0.02, **p < 0.001, Mann Whitney U). (c) Rasters of single units in the GPe of naïve vs. bilateral DD mice. Scale bar, 500 ms. (d) Box plots showing decreases in firing rates (Naive: 44.3 ± 2.6 Hz, n = 73 across 4 animals, versus Acute: 24.6 ± 1.6 Hz,n = 62 across 3 animals, H(1) = 29.775, **p < 0.001, Kruskal-Wallis H test) and (e) increases in coefficients of variation of the interspike intervals (CVISI) (Naïve: 0.63 ± 0.03 versus Acute: 0.80 ± 0.03, H(1) = 22.615, *p < 0.001, Kruskal-Wallis H test) following bilateral DD. Error bars, sem.

Supplementary Figure 2 Histological verification of TH immunoreactivity, viral expression and fiber placements for behavioral optogenetics in global manipulations.

(a) Representative images of striatal TH immunoreactivity in dorsal striatum of healthy dopamine intact tissue compared to a fully depleted bilateral animal. Scale Bar, 200 μm (b) Epifluorescent images of viral expression and fiber identification (yellow arrow) within the GPe. Scale Bar, 500 μm (c) Superimposed traces of viral expression across animals within hSyn-ChR2 (d) Epifluorescent images of viral expression and fiber identification (yellow arrow) with the dorsal striatum. Scale Bar, 500 μm (e) Superimposed traces of viral expression across animals within D1-ChR2 condition. Ctx = Cortex, Str = Striatum, GPe = globus pallidus externa.

Supplementary Figure 3 Histological verification of viral expression and fiber placements for behavioral optogenetics in cell-type manipulations.

(a) Epifluorescent images of viral expression and fiber identification (yellow arrow) within the GPe. Scale Bar, 500 μm (b-f) Superimposed traces of viral expression across animals within PV-ChR2, Lhx6-Arch, CAG-Arch, Lhx6-ChR2 and PV-Arch conditions.

Supplementary Figure 4 PV-ChR2 stimulation induces transient effects in partially and unilaterally depleted mice.

(a) Percentage of time spent in the immobile state before, during and after PV-ChR2 stimulation in lipopolysaccharide (LPS) injected mice. (b) Schematic representation of striatal location for TH analysis with epifluorescent image of a partial depletion of tyrosine hydroxylase. Scale Bar, 200 μm. (c) Quantification of TH levels, normalized to dopamine intact littermate and subsequent categorization into partially and fully depleted animals.(d) Percentage of time spent in the immobile state before, during, and after PV-ChR2 stimulation in partially depleted mice. (e) Overlay of immobility immediately before (pre), during (stim), and after (post) each light pulse. (f) Percent time spent immobile for each animal (grey x = pre, black circle = post) and degree of rescue (red line) as a function of TH remaining. Note sharp cut-off for induction of behavioral rescue at ~20% dopamine remaining. (g) Percentage of time spent in the immobile state before, during, and after PV-ChR2 stimulation in unilaterally depleted mice. (h) Overlay of immobility immediately before (pre), during (stim), and after (post) each light pulse. (i) Percent time spent immobile over the course of the full experimental trial. Error bars, sem.

Supplementary Figure 5 Optical identification using ChR2 and Arch and their corresponding firing properties and waveforms.

(a) Representative responses over 20 trials from a ChR2+ (putative PV) and CHR2- (putative non-PV) neuron responding to 5 ms optical pulses. Yellow bar denotes first significant bin as compared to baseline (b) Firing rate and coefficient of variation of the interspike interval (CVISI) for ChR2+ and ChR2- neurons in dopamine depleted (FR: p = 0.128, CVISI: p = 0.005, Mann Whitney U) (c) Extracellular waveform analysis of the peak-valley ratio and amplitude of individual units identified as ChR2+ (red, closed circles) and ChR2- (black, open circles). Inset: Average waveforms of ChR2+ and ChR2- (Note: nearly complete overlap). Scale bar: 50 μV(vertical), 220 μsec (horizontal) (d) Firing rate and coefficient of variation of the interspike interval (CVISI) for ChR2+ and ChR2- neurons in dopamine intact animals (Naïve) (e) Representative responses over 20 trials from a Arch+ (putative Lhx6) and Arch- (putative non-Lhx6) neuron responding to 1 s optical pulses. (f) Firing rate and CVISI for Arch+ and Arch- neurons (FR: p = 0.990, CV: p = 0.454, Mann Whitney U). (g) Extracellular waveform analysis of the peak-valley ratio and amplitude of individual units identified as Arch+ (blue, closed squares) and Arch- (black, open squares). Inset: Average waveforms of Arch+ and Arch-. Scale bar: 50 μV(vertical), 220 μsec (horizontal).

Supplementary Figure 6 PV and Lhx6 overlap partially in the Lhx6-Cre transgenic mouse.

(a) Fluorescent images from the GPe showing overlap between Lhx6-iCre and PV. Left: Lhx6-EYFP neurons. Arrows denote position of PV+ neurons (Purple arrows: Lhx6/PV double labeled; Red arrows: PV-only neurons). Note weaker Lhx6-EYFP expression in Lhx6/PV neurons. Middle: PV+ neurons (Blue arrows: Lhx6-only neurons). Right: Overlay. (b) Box diagram summarizing the proportion of GPe neurons (n = 4194 total neurons across 5 animals) counted that expressed either PV, Lhx6, or both.

Supplementary Figure 7 SNr firing rate is unaltered after PV-ChR2 and Lhx6-Arch, but decreased after hSyn-ChR2 manipulation.

(a) Firing rate of single units collected before (pre) and after (post) stimulation in PV-ChR2 (nPre = 80 vs nPost = 58 units across 3 animals, p = 0.976, Mann Whitney U), hSYn-ChR2 (nPre = 55 vs nPost = 69 unit across 3 animals, p = 0.002, Mann Whitney U) and Lhx6-Arch (nPre = 30 vs nPost = 69 units across 3 animals, p = 0.142, Mann Whitney U).

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mastro, K., Zitelli, K., Willard, A. et al. Cell-specific pallidal intervention induces long-lasting motor recovery in dopamine-depleted mice. Nat Neurosci 20, 815–823 (2017). https://doi.org/10.1038/nn.4559

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.4559

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