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:

Intrinsic dopamine and acetylcholine dynamics in the striatum of mice

Abstract

External rewards such as food and money are potent modifiers of behaviour1,2. Pioneering studies established that these salient sensory stimuli briefly interrupt the tonic discharge of neurons that produce the neuromodulators dopamine (DA) and acetylcholine (ACh): midbrain DA neurons (DANs) fire a burst of action potentials that broadly elevates DA in the striatum3,4 at the same time that striatal cholinergic interneurons (CINs) produce a characteristic pause in firing5,6. These phasic responses are thought to create unique, temporally limited conditions that motivate action and promote learning7,8,9,10,11. However, the dynamics of DA and ACh outside explicitly rewarded situations remain poorly understood. Here we show that extracellular DA and ACh levels fluctuate spontaneously and periodically at a frequency of approximately 2 Hz in the dorsal striatum of mice and maintain the same temporal relationship relative to one another as that evoked by reward. We show that this neuromodulatory coordination does not arise from direct interactions between DA and ACh within the striatum. Instead, we provide evidence that periodic fluctuations in striatal DA are inherited from midbrain DANs, while striatal ACh transients are driven by glutamatergic inputs, which act to locally synchronize the spiking of CINs. Together, our findings show that striatal neuromodulatory dynamics are autonomously organized by distributed extra-striatal afferents. The dominance of intrinsic rhythms in DA and ACh offers new insights for explaining how reward-associated neural dynamics emerge and how the brain motivates action and promotes learning from within.

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: Spontaneous periodic fluctuations in striatal DA and ACh.
Fig. 2: DA and ACh maintain a consistent temporal relationship across behavioural states.
Fig. 3: Periodic DA and ACh fluctuations are not coordinated locally in the striatum.
Fig. 4: Glutamatergic inputs synchronize CIN spiking and drive periodic ACh fluctuations.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available from the corresponding author on reasonable request. Source data are provided with this paper.

Code availability

All code related to this study was developed in MATLAB and is available at https://github.com/ackrok/Krok-etal_2023.

References

  1. Schultz, W. Behavioral theories and the neurophysiology of reward. Annu. Rev. Psychol. 57, 87–115 (2006).

    Article  PubMed  Google Scholar 

  2. O’Doherty, J. P., Cockburn, J. & Pauli, W. M. Learning, reward, and decision making. Annu. Rev. Psychol. 68, 73–100 (2017).

    Article  PubMed  Google Scholar 

  3. Mirenowicz, J. & Schultz, W. Preferential activation of midbrain dopamine neurons by appetitive rather than aversive stimuli. Nature 379, 449–451 (1996).

    Article  ADS  CAS  PubMed  Google Scholar 

  4. Cohen, J. Y., Haesler, S., Vong, L., Lowell, B. B. & Uchida, N. Neuron-type-specific signals for reward and punishment in the ventral tegmental area. Nature 482, 85–88 (2012).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  5. Morris, G., Arkadir, D., Nevet, A., Vaadia, E. & Bergman, H. Coincident but distinct messages of midbrain dopamine and striatal tonically active neurons. Neuron 43, 133–143 (2004).

    Article  CAS  PubMed  Google Scholar 

  6. Aosaki, T., Graybiel, A. M. & Kimura, M. Effect of the nigrostriatal dopamine system on acquired neural responses in the striatum of behaving monkeys. Science 265, 412–415 (1994).

    Article  ADS  CAS  PubMed  Google Scholar 

  7. Costa, R. M. A selectionist account of de novo action learning. Curr. Opin. Neurobiol. 21, 579–586 (2011).

    Article  CAS  PubMed  Google Scholar 

  8. Coddington, L. T. & Dudman, J. T. Learning from action: reconsidering movement signaling in midbrain dopamine neuron activity. Neuron 104, 63–77 (2019).

    Article  CAS  PubMed  Google Scholar 

  9. Berke, J. D. What does dopamine mean? Nat. Neurosci. 21, 787–793 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Klaus, A., Alves da Silva, J. & Costa, R. M. What, if, and when to move: basal ganglia circuits and self-paced action initiation. Annu. Rev. Neurosci. 42, 459–483 (2019).

    Article  CAS  PubMed  Google Scholar 

  11. Cox, J. & Witten, I. B. Striatal circuits for reward learning and decision-making. Nat. Rev. Neurosci. 20, 482–494 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Shen, W. et al. M4 muscarinic receptor signaling ameliorates striatal plasticity deficits in models of l-DOPA-induced dyskinesia. Neuron 88, 762–773 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Reynolds, J. N. J. et al. Coincidence of cholinergic pauses, dopaminergic activation and depolarisation of spiny projection neurons drives synaptic plasticity in the striatum. Nat. Commun. 13, 1296 (2022).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  14. Howe, M. W. & Dombeck, D. A. Rapid signalling in distinct dopaminergic axons during locomotion and reward. Nature 535, 505–510 (2016).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  15. da Silva, J. A., Tecuapetla, F., Paixao, V. & Costa, R. M. Dopamine neuron activity before action initiation gates and invigorates future movements. Nature 554, 244–248 (2018).

    Article  ADS  PubMed  Google Scholar 

  16. Panigrahi, B. et al. Dopamine is required for the neural representation and control of movement vigor. Cell 162, 1418–1430 (2015).

    Article  CAS  PubMed  Google Scholar 

  17. Engelhard, B. et al. Specialized coding of sensory, motor and cognitive variables in VTA dopamine neurons. Nature 570, 509–513 (2019).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  18. Mohebi, A. et al. Dissociable dopamine dynamics for learning and motivation. Nature 570, 65–70 (2019).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  19. Howe, M. et al. Coordination of rapid cholinergic and dopaminergic signaling in striatum during spontaneous movement. eLife 8, e44903 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  20. Sulzer, D., Cragg, S. J. & Rice, M. E. Striatal dopamine neurotransmission: regulation of release and uptake. Basal Ganglia 6, 123–148 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  21. Threlfell, S. et al. Striatal dopamine release is triggered by synchronized activity in cholinergic interneurons. Neuron 75, 58–64 (2012).

    Article  CAS  PubMed  Google Scholar 

  22. Cachope, R. et al. Selective activation of cholinergic interneurons enhances accumbal phasic dopamine release: setting the tone for reward processing. Cell Rep. 2, 33–41 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Liu, C. et al. An action potential initiation mechanism in distal axons for the control of dopamine release. Science 375, 1378–1385 (2022).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  24. Straub, C., Tritsch, N. X., Hagan, N. A., Gu, C. & Sabatini, B. L. Multiphasic modulation of cholinergic interneurons by nigrostriatal afferents. J. Neurosci. 34, 8557–8569 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  25. Chuhma, N., Mingote, S., Moore, H. & Rayport, S. Dopamine neurons control striatal cholinergic neurons via regionally heterogeneous dopamine and glutamate signaling. Neuron 81, 901–912 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Sun, F. et al. Next-generation GRAB sensors for monitoring dopaminergic activity in vivo. Nat. Methods 17, 1156–1166 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Jing, M. et al. An optimized acetylcholine sensor for monitoring in vivo cholinergic activity. Nat. Methods 17, 1139–1146 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Tritsch, N. X. & Sabatini, B. L. Dopaminergic modulation of synaptic transmission in cortex and striatum. Neuron 76, 33–50 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Hnasko, T. S. et al. Vesicular glutamate transport promotes dopamine storage and glutamate corelease in vivo. Neuron 65, 643–656 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Tritsch, N. X., Ding, J. B. & Sabatini, B. L. Dopaminergic neurons inhibit striatal output through non-canonical release of GABA. Nature 490, 262–266 (2012).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  31. Shin, J. H., Adrover, M. F., Wess, J. & Alvarez, V. A. Muscarinic regulation of dopamine and glutamate transmission in the nucleus accumbens. Proc. Natl Acad. Sci. USA 112, 8124–8129 (2015).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  32. Joshua, M. et al. Synchronization of midbrain dopaminergic neurons is enhanced by rewarding events. Neuron 62, 695–704 (2009).

    Article  CAS  PubMed  Google Scholar 

  33. Liu, C., Goel, P. & Kaeser, P. S. Spatial and temporal scales of dopamine transmission. Nat. Rev. Neurosci. 22, 345–358 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  34. Fujisawa, S. & Buzsaki, G. A 4 Hz oscillation adaptively synchronizes prefrontal, VTA, and hippocampal activities. Neuron 72, 153–165 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. 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 

  36. Guo, Q. et al. Whole-brain mapping of inputs to projection neurons and cholinergic interneurons in the dorsal striatum. PLoS ONE 10, e0123381 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  37. Neske, G. T. Sleepy circuits in vigilant mice? A slow cortical oscillation occurring during multiple arousal states. J. Neurosci. 37, 7294–7296 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Nacher, V., Ledberg, A., Deco, G. & Romo, R. Coherent delta-band oscillations between cortical areas correlate with decision making. Proc. Natl Acad. Sci. USA 110, 15085–15090 (2013).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  39. Lee, K. et al. Gain modulation by corticostriatal and thalamostriatal input signals during reward-conditioned behavior. Cell Rep. 29, 2438–2449 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Quick, M. W. & Lester, R. A. Desensitization of neuronal nicotinic receptors. J. Neurobiol. 53, 457–478 (2002).

    Article  CAS  PubMed  Google Scholar 

  41. Maskos, U. et al. Nicotine reinforcement and cognition restored by targeted expression of nicotinic receptors. Nature 436, 103–107 (2005).

    Article  ADS  CAS  PubMed  Google Scholar 

  42. Choi, S. J. et al. Alterations in the intrinsic properties of striatal cholinergic interneurons after dopamine lesion and chronic l-DOPA. eLife 9, e56920 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Mamaligas, A. A., Barcomb, K. & Ford, C. P. Cholinergic transmission at muscarinic synapses in the striatum is driven equally by cortical and thalamic inputs. Cell Rep. 28, 1003–1014 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Zhang, Y. F., Reynolds, J. N. J. & Cragg, S. J. Pauses in cholinergic interneuron activity are driven by excitatory input and delayed rectification, with dopamine modulation. Neuron 98, 918–925 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Beeler, J. A. & Kisbye Dreyer, J. Synchronicity: the role of midbrain dopamine in whole-brain coordination. eNeuro 6, ENEURO.0345-18.2019 (2019).

  46. Shen, W., Flajolet, M., Greengard, P. & Surmeier, D. J. Dichotomous dopaminergic control of striatal synaptic plasticity. Science 321, 848–851 (2008).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  47. Liu, Y., Mattar, M. G., Behrens, T. E. J., Daw, N. D. & Dolan, R. J. Experience replay is associated with efficient nonlocal learning. Science 372, eabf1357 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Graybiel, A. M. The basal ganglia and chunking of action repertoires. Neurobiol. Learn. Mem. 70, 119–136 (1998).

    Article  CAS  PubMed  Google Scholar 

  49. Soares, S., Atallah, B. V. & Paton, J. J. Midbrain dopamine neurons control judgment of time. Science 354, 1273–1277 (2016).

    Article  ADS  CAS  PubMed  Google Scholar 

  50. Liu, Y., Nour, M. M., Schuck, N. W., Behrens, T. E. J. & Dolan, R. J. Decoding cognition from spontaneous neural activity. Nat. Rev. Neurosci. 23, 204–214 (2022).

    Article  PubMed  Google Scholar 

  51. Patel, J. C., Rossignol, E., Rice, M. E. & Machold, R. P. Opposing regulation of dopaminergic activity and exploratory motor behavior by forebrain and brainstem cholinergic circuits. Nat. Commun. 3, 1172 (2012).

    Article  ADS  PubMed  Google Scholar 

  52. Burbridge, T. J. et al. Visual circuit development requires patterned activity mediated by retinal acetylcholine receptors. Neuron 84, 1049–1064 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Backman, 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 

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

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  55. Thiele, S. L., Warre, R. & Nash, J. E. Development of a unilaterally-lesioned 6-OHDA mouse model of Parkinson’s disease. J. Vis. Exp. https://doi.org/10.3791/3234 (2012).

  56. Yang, C. F. et al. Sexually dimorphic neurons in the ventromedial hypothalamus govern mating in both sexes and aggression in males. Cell 153, 896–909 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Warren, R. A. et al. A rapid whisker-based decision underlying skilled locomotion in mice. eLife 10, e63596 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Yang, L., Lee, K., Villagracia, J. & Masmanidis, S. C. Open source silicon microprobes for high throughput neural recording. J. Neural Eng. 17, 016036 (2020).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  59. Stanford Research Systems. About lock-in amplifiers. https://www.thinksrs.com/downloads/pdfs/applicationnotes/AboutLIAs.pdf (2016).

  60. Balakrishnan, H. & Verghese, G. Modulation and demodulation. http://web.mit.edu/6.02/www/s2012/handouts/14.pdf (2012).

  61. Zutshi, I., Valero, M., Fernandez-Ruiz, A. & Buzsaki, G. Extrinsic control and intrinsic computation in the hippocampal CA1 circuit. Neuron 110, 658–673 (2022).

    Article  CAS  PubMed  Google Scholar 

  62. Berke, J. D., Okatan, M., Skurski, J. & Eichenbaum, H. B. Oscillatory entrainment of striatal neurons in freely moving rats. Neuron 43, 883–896 (2004).

    Article  CAS  PubMed  Google Scholar 

  63. Schmitzer-Torbert, N. C. & Redish, A. D. Task-dependent encoding of space and events by striatal neurons is dependent on neural subtype. Neuroscience 153, 349–360 (2008).

    Article  CAS  PubMed  Google Scholar 

  64. Sharott, A., Doig, N. M., Mallet, N. & Magill, P. J. Relationships between the firing of identified striatal interneurons and spontaneous and driven cortical activities in vivo. J. Neurosci. 32, 13221–13236 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Yamin, H. G., Stern, E. A. & Cohen, D. Parallel processing of environmental recognition and locomotion in the mouse striatum. J. Neurosci. 33, 473–484 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Peters, A. J., Fabre, J. M. J., Steinmetz, N. A., Harris, K. D. & Carandini, M. Striatal activity topographically reflects cortical activity. Nature 591, 420–425 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  67. Gage, G. J., Stoetzner, C. R., Wiltschko, A. B. & Berke, J. D. Selective activation of striatal fast-spiking interneurons during choice execution. Neuron 67, 466–479 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank G. Buzsaki, M. Long, S. Shoham, R. Tsien and members of the Tritsch laboratory for comments on the manuscript and M. Crair (Yale University) and R. Machold (NYULH) for providing the β2flox and ChATflox;Nkx2.1Cre mice, respectively. This work was supported by the National Institutes of Health (DP2NS105553 and R01MH130658 to N.X.T.; T32NS086750, T32GM007308 and T32GM136573 to A.C.K.), the Paolo and Marlene Fresco, Alfred P. Sloan, Dana, Whitehall and Feldstein Medical foundations to N.X.T. and a Vilcek Scholars Award to A.C.K. We acknowledge the New York University Langone Health Rodent Genetic Engineering Laboratory for rederivation, the Genotyping Core Laboratory for mouse genotyping, the Department of Comparative Medicine for animal care and maintenance, and the Neuroscience Institute’s imaging facilities for microscope availability. We apologize to those whose work we were unable to cite due to length limits.

Author information

Authors and Affiliations

Authors

Contributions

A.C.K. and N.X.T. conceived of the project, designed and performed experiments, analysed and interpreted the data, and wrote the manuscript. A.C.K. and P.M. wrote code for photometry data acquisition and analyses, M.M. performed DA neuron lesions and helped with photometry data collection and analyses, and X.M. and Y.L. performed experiments to determine the kinetics of ACh3.0 and provided the GRAB sensors rDA1m and ACh3.0.

Corresponding author

Correspondence to Nicolas X. Tritsch.

Ethics declarations

Competing interests

Y.L. is listed as an inventor on a patent application (PCT/CN2018/107533) describing GRAB probes. The other authors declare no competing interests.

Peer review

Peer review information

Nature thanks Anna Beyeler, Sean Ostlund and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

Additional information

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 Fig. 1 Characterization of rDA1m and ACh3.0 expression.

a, Epifluorescence images of endogenous rDA1m (left) and ACh3.0 (middle) fluorescence in a coronal section of the striatum. Right, merged image overlaid with DAPI nuclear stain (blue). Similar results were obtained in n = 16 other brains recovered for post-hoc validation of GRAB sensor expression and fiber optic placement. b, Epifluorescence images of immuno-enhanced rDA1m and ACh3.0 expressed in the dorsal striatum. Note the presence of rDA1m- and ACh3.0-positive axonal projections in the globus pallidus externus (GPe) and internus (GPi), consistent with the subcellular distribution of dopamine and muscarinic receptors in both direct and indirect-pathway SPNs. c, High magnification confocal images (single optical plane) of immuno-enhanced rDA1m and ACh3.0 in a coronal striatal section. Note the localization of both sensors to the somatic membranes of striatal neurons and throughout the neuropil, and their exclusion from corticofugal axon bundles (asterisks). Note also that some rDA1m forms intracellular aggregates. d, Same as c for a sagittal section of the substantia nigra pars reticulata. Subcellular localization of immuno-enhanced rDA1m and ACh3.0 by high magnification confocal microscopy was independently confirmed in four mice. e, Schematic showing the tip of all recovered fiber optic implants in the DLS (blue) and DMS (green) at 3 different anterior-posterior levels relative to bregma (in mm).

Extended Data Fig. 2 Kinetics of ACh3.0 sensor in the striatum in vivo.

a, Experimental preparation to optogenetically evoke ACh release in vivo using ChrimsonR while simultaneously monitoring ACh levels with ACh3.0 photometry via the same fibre optic. ChrimsonR and ACh3.0 were both virally-expressed. b, Example ACh3.0 recording (teal). ChrimsonR was stimulated every 200 s with 635 nm light flashes (20 Hz, 1 s; shown in magenta). c, detail of dashed box in b showing ChrimsonR-evoked ACh release. d, Overlay of individual ACh3.0 transients (light teal) from b aligned to optogenetic stimulation onset. Group mean shown in dark teal with overlaid exponential fits used to calculate onset (τon) and offset (τoff) time constants in black. e, ACh3.0 τon (left) and τoff (right). Mean (± s.e.m.) shown in black (τon: 138 ± 10 ms; τoff: 247 ± 32 ms; n = 12 trials from 3 mice).

Source Data

Extended Data Fig. 3 Reward-evoked DA and ACh responses in the DLS.

a, Example continuous recording of rDA1m (magenta) and ACh3.0 (teal) fluorescence, lick events (black) and treadmill acceleration (gray) illustrating the temporal dynamics of both sensors. Dashed blue lines depict uncued solenoid valve opening, which in this example did not generate large DA and ACh reward transients. b, Other reward response from same mouse. c, Example DA and ACh reward responses for a different mouse. d, Mean rDA1m (magenta) and ACh3.0 (teal) fluorescence aligned to solenoid valve opening for all uncued water deliveries followed within 1 s by consummatory licking (n = 13 mice). e, Same as d for water deliveries not followed by consummatory licking. Gray: 95% confidence interval. f, Same as d, but aligned to first lick after solenoid valve opening (i.e., first rewarded lick). g, Mean rDA1m and ACh3.0 fluorescence aligned to licks occurring between consummatory licking bouts (i.e., unrewarded licks). Gray: 95% confidence interval. h–j, Latency of first rewarded lick from solenoid valve opening (h; P = 0.52), fraction of rewards consumed within 1 s of valve opening (i; P = 0.14) and number of consummatory licks per reward delivery (j; P = 0.59) during the first (early) and last (late) quartiles of each imaging session in n = 13 mice suggest that the motivation to consume rewards is stable. k, Mean reward-aligned DA and ACh responses imaged during the first (early; gray) and last (late; colored) quartiles of each imaging session are similar in amplitude and kinetics (n = 13 mice). l, Scatterplot of the mean amplitude of reward-evoked DA peaks (left; P = 0.24) and ACh throughs (right; P = 0.74) at the beginning (early) vs. end (late) of each imaging session (n = 13 mice). m, Scatterplot of the amplitude of individual DA peaks and concurrent ACh troughs imaged during reward (blue) and periods of immobility (orange) in one example recording session. Amplitude distribution for ACh (top) and DA (right) with overlap shown in gray. n, Mean distribution of DA peak (left) and ACh trough (right) amplitudes for all mice (n = 13) normalized to the mode of reward distributions showing considerable overlap between reward and immobility states. o, Left, power spectrum of DA signal recorded in the DLS during immobility in n = 10 female mice (black) and n = 8 male mice (magenta). Right, Area under the curve of DA power spectra in the 0.5–4 Hz frequency band (P = 0.65). p, Same as o, for ACh (P = 0.84). Group means (± s.e.m.) in h–j, o and p are shown in black. Shaded regions in dg, k, n, o and p denote s.e.m. Statistical comparisons are Student’s paired t-tests in h–j and l, and Student’s two-sample t-tests in o and p.

Source Data

Extended Data Fig. 4 Locomotion-evoked DA and ACh responses in the DLS.

a–b, Experimental setup: virally-expressed rDA1m and ACh3.0 were imaged in the DLS (a) of mice head-fixed on a transparent cylindrical treadmill (b). c, Example rDA1m and ACh3.0 fluorescence from 2 separate mice during immobility (left) and spontaneous locomotion on the treadmill (right). d, Mean standard deviation of photometry signal during immobility and locomotion for ACh3.0 (top; P = 6.5 x 10−8, Student’s paired t-test) and rDA1m (bottom; P = 3.6x10−4; Student’s paired t-test). Group mean (± s.e.m.) shown in black (n = 13 mice). e, Left, mean ACh (top) and DA (bottom) power density spectrum during locomotion (n = 13 mice). Right, Area under the curve (AUC) in 0.5–4 Hz frequency band. Group mean (± s.e.m.) shown in black. f, Heatmap of mean ACh3.0 (top) and rDA1m (bottom) fluorescence aligned to momentary peaks in treadmill acceleration during locomotion across mice. g, Group-averaged ACh3.0 (top) and rDA1m (bottom) fluorescence aligned to acceleration peaks stratified into tertiles according to acceleration magnitude. ACh and DA levels co-vary with acceleration magnitude. ACh levels peaking near acceleration maxima, whereas DA levels dip shortly before and crest shortly after acceleration maxima. h, Mean cross-correlation between treadmill acceleration and either ACh3.0 (teal) or rDA1m (magenta) fluorescence (n = 13 mice). i, Mean treadmill acceleration at different phases of periodic ACh fluctuations (dashed teal line) in the 0.5–4 Hz frequency band during locomotion (black) or immobility (gray) in n = 13 mice. During locomotion, periodic ACh fluctuations occur in phase with positive treadmill acceleration. Note that periodic ACh fluctuations during immobility are not associated with movements of the treadmill. j, Same as i for DA fluorescence. Note that DA fluctuations are phase-delayed relative to treadmill acceleration during locomotion, and not associated with micro-movements of the treadmill during immobility. Shaded regions in e, gj denote s.e.m.

Source Data

Extended Data Fig. 5 Spontaneous fluctuations in striatal DA and ACh in unrestrained mice.

a, Simultaneous photometry recording of rDA1m (magenta) and ACh3.0 (teal) fluorescence from a mouse in an open field arena. b, Mean cross-correlation between simultaneously recorded DA and ACh during locomotion (green) and immobility (orange) in n = 4 mice. c, Peak correlation coefficient between DA and ACh (P = 0.93) during immobility and locomotion (n = 4 mice). d, Same as c for time lag of negative cross-correlation peak (p = 0.06). e, Example coherence between DA and ACh signals across frequency and time domains during a recording in an open field arena. f, Left, mean coherence between DA and ACh at different frequencies (n = 4 mice). Vertical lines depict 0.5 and 4 Hz. Right, median coherence in 0.5–4 Hz frequency band in individual mice (P = 0.25). g–h, Same as e-f for phase offset between DA and ACh (P = 0.93). i, Mean DA fluorescence at different phases of periodic DA fluctuations in the 0.5-4 Hz frequency band (n = 4 mice). j, Same as h for ACh fluorescence. k, Peak-normalized DA fluorescence vs. phase of periodic ACh fluctuations (gray dotted line) in 0.5–4 Hz frequency band during locomotion (green) and immobility (orange). Group means (± s.e.m.) in c, d, f and h shown in black. Shaded areas in b, f, h–k reflect s.e.m. All statistical comparisons are Student’s paired t-tests.

Source Data

Extended Data Fig. 6 DA and ACh levels also fluctuate periodically in DMS.

a, Experimental setup. b, Left, mean power spectrum of DA signal recorded simultaneously from DMS (magenta) and DLS (black) during immobility in n = 5 mice. Right, area under the DA power spectra in the 0.5–4 Hz frequency band in DMS and DLS. Group averages (± s.e.m.) shown in black (p = 0.8, Student’s paired t-test). c, Same as b for ACh (p = 0.3). d, Left, mean cross-correlation between DA and ACh in DMS across different behavioral states. Middle, peak correlation coefficient (Pearson’s r) in individual mice. Group averages (± s.e.m.) shown in black (**p = 0.006 vs. reward and immobility, one-way balanced ANOVA, Dunn’s multiple comparisons; n = 5 mice). Right, time lag of negative cross-correlation peak (p = 0.4, one-way balanced ANOVA). e, Magnitude of coherence between DA and ACh in DMS across frequency and time domains for an example recording. f, Left, mean coherence at different frequencies across behavioral states (n = 5 mice). Vertical lines depict 0.5 and 4 Hz. Right, median coherence in 0.5–4 Hz frequency band in individual mice. Group averages (± s.e.m.) shown in black (p = 0.09, one-way unbalanced ANOVA; n = 5 mice). g, Same as e for phase offset between DA and ACh in DMS. h, Same as f for phase offset between DA and ACh in DMS (p = 0.2, one-way unbalanced ANOVA). i, Left, mean DA fluorescence at different phase of DA fluctuations in the 0.5–4 Hz frequency band in DMS. Right, same for ACh fluorescence vs. phase of ACh fluctuations. j, Peak-normalized DA fluorescence vs. phase of ACh in DMS. k, Left, mean coherence between simultaneously recorded DADLS and AChDMS (n = 5 mice). Right, same for coherence between DADMS and AChDLS. l, Median coherence in 0.5–4 Hz band in individual mice. Group averages (± s.e.m.) shown in black (reward: DADLS+AChDMS p = 8.14 x 10−7; DADMS+AChDLS p = 6.07 x 10−7; locomotion: DADLS+AChDMS p = 7.46 x 10−5; DADMS+AChDLS p = 6.07 x 10−5; immobility: DADLS+AChDMS p = 0.004; DADMS+AChDLS p = 0.005; all vs. DADLS+AChDLS, one-way unbalanced ANOVA, Dunn’s multiple comparisons; n = 5 mice). Shaded areas in b–d, f, h–k reflect s.e.m.

Source Data

Extended Data Fig. 7 DA does not drive spontaneous fluctuations in ACh.

a, Left, example session-averaged DA fluorescence aligned to solenoid valve opening for all uncued water deliveries followed within 1 s by consummatory licking after intra-striatal infusion of saline (black) or a cocktail of D1 and D2 receptor antagonists (blue). Right, mean reward-evoked DA amplitude (n = 4 mice). b, Left, same as a for ACh. Right, scatter plot of mean reward-evoked ACh dip amplitude and timing in mice infused with saline or D1/2R antagonists (amplitude: p = 0.2; timing: p = 0.9, both vs. saline). c, Frequency (left) and amplitude (right) of spontaneous dips in ACh fluorescence during immobility (n = 4 mice). d, Same as c for peaks in ACh. e, Experimental preparation to lesion midbrain DA neurons (DANs) unilaterally while imaging DA and ACh in the DLS ipsilaterally. DANs were lesioned using either 6OHDA infusions in wild type mice (n = 7), or AAV-mediated expression of Cre-dependent caspase-3 (AAV-FLEX-taCasp3) in the midbrain DANs of DatCre mice (n = 4). f, Epifluorescence image of ACh3.0 (green) and DA transporter (DAT; red) immunofluorescence overlaid with DAPI nuclear stain (blue) in a coronal section from a DatCre mouse injected with AAV-FLEX-taCasp3 into the right midbrain. Similar results were obtained in other 10 mice. g, Example DA (magenta) and ACh (teal) fluorescence following DAN lesion during immobility. h, Left, mean ACh power spectrum during immobility before (black) and after DAN lesion (teal; n = 11 mice). Right, ACh power area under the curve (AUC) in 0.5–4 Hz frequency band. i, Same as c before and after DAN lesion (n = 11 mice). Chronic DA depletion causes phasic dips in ACh to become larger and more frequent, hinting at a negative influence of DA on ACh release. j, Same as d before and after DAN lesion (n = 11 mice). Group means (± s.e.m.) in ad, hj are shown in black. Shaded areas in a,b and h reflect s.e.m. All statistical comparisons are Student’s paired t-tests.

Source Data

Extended Data Fig. 8 ACh signaling is not required for periodic DA fluctuations.

a, Left, example session-averaged ACh fluorescence aligned to solenoid opening for all uncued water deliveries followed within 1 s by consummatory licking in control (black) and β2 cKODAN (blue) mice. Right, scatter plot of mean reward-evoked ACh dip amplitude and timing in control (gray; n = 13) and β2 cKODAN (blue; n = 6) mice. Group means (± s.e.m.) overlaid (amplitude: p = 0.29; timing: p = 0.89, both vs. control, Student’s two-sample t-test). b, Left, same as a for reward-evoked DA peak (amplitude: p = 0.58; timing, p = 0.33). c, Left, average coherence between DA and ACh signals in β2 cKODAN mice (n = 6) at different frequencies during reward (blue), locomotion (green) and immobility (orange). Vertical lines depict 0.5 and 4 Hz. Right, same for phase offset. d, Median DA–ACh coherence and phase in 0.5–4 Hz band in β2 cKODAN mice (n = 6) normalized to control mice across conditions (coherence: reward p = 0.9, locomotion p = 0.3, immobility p = 0.9; phase: reward p = 0.4, locomotion p = 0.6, immobility p = 0.2; all comparisons vs. control, Student’s two-sample t-test). Group mean (± s.e.m.) shown in black. e, Left, average DA fluorescence at different phases of periodic DA fluctuations in the 0.5–4 Hz frequency band across behavioral states in β2 cKODAN mice (n = 6 mice). Right, same for ACh fluorescence vs. phase of periodic ACh fluctuations. f, Experimental preparation for local pharmacological inhibition of nAChR signaling. g, Same as a after DLS infusion of saline (black) or the nAChR antagonist DHβE (blue; amplitude: p = 0.4; timing: p = 0.13, both vs. saline, Student’s paired t-test, n = 5 mice). h, Same as g for DA (amplitude: p = 0.9; timing: p = 0.15). i, Example DA (magenta) and ACh (teal) traces during infusion of DHβE. j, Left, mean ACh power spectrum during immobility following intra-striatal infusion of saline (black) or DHβE (blue; n = 8 mice). Right, ACh power area under the curve (AUC) in 0.5–4 Hz frequency band (p = 0.6, Student’s paired t-test). k, Same as d for Pearson’s correlation coefficient r (reward: p = 0.2, locomotion: p = 0.5, immobility: p = 0.06), coherence (reward: p = 0.04, locomotion: p = 0.6, immobility: p = 0.09) and phase (reward: p = 0.2, locomotion: p = 0.5, immobility: p = 0.6) in DHβE normalized to saline (all comparisons vs. saline, Student’s paired t-test; n = 8 mice). l, Experimental preparation for local pharmacological inhibition of mAChR signaling in the DLS. m, Left, same as g following infusion of saline (black) or the mAChR antagonist scopolamine (blue). Right, amplitude of reward-evoked dip in ACh fluorescence (p = 0.02; n = 5 mice). n, same as j after infusion of saline or mAChR antagonist (p = 2.5 x 10−4, Student’s paired t-test; n = 5 mice). This effect confirms that scopolamine reaches the volume of tissue imaged by the fiber optic. o, Same as m for DA (amplitude: p = 0.6, timing: p = 0.09; n = 5 mice). p, Same as n for DA (p = 0.048). This effect is consistent with the role of facilitatory effects of presynaptic mAChRs on DAN axons (see ref. 31). Shaded areas in a,b,c,e,g,h,j and m–p reflect s.e.m.

Source Data

Extended Data Fig. 9 Midbrain DAN activity underlies periodic fluctuations in striatal DA.

a, Experimental preparation for simultaneously imaging the activity of DANs in the SNc and their axons in the DLS. b, Example photometry recording of GCaMP6f in the SNc (black) and DLS (blue) during immobility. c, Left, average (± s.e.m. shown as shaded region) cross-correlation between SNc and DLS signals during immobility in n = 6 mice. 95% confidence interval shown in gray. Right, peak correlation coefficient between SNc and DLS signals. Group mean (± s.e.m.) in black. d, Experimental preparation for pharmacological silencing of SNc while imaging of rDA1m and ACh3.0 in the DLS. e, Left, Example continuous recording of ACh3.0 (teal) and rDA1m (magenta). Black line depicts muscimol infusion. Right, Detail of dashed boxes showing ACh and DA fluctuations during immobility before (e1) and following (e2) infusion of muscimol. Note that muscimol abolishes phasic DA transients as well as ‘global’ DA levels in the DLS, whereas ACh global levels and phasic transients persist. f, Average global ACh and DA fluorescence during immobility after infusion of saline and muscimol. Group mean (± s.e.m.) shown in black (ACh: p = 0.021; DA: p = 0.0001, Student’s paired t-test; n = 6 mice). g, Frequency (left) and amplitude (right) of spontaneous dips in ACh fluorescence recorded during immobility after infusion of saline (gray) or muscimol (teal). Group mean (± s.e.m.) shown in black (n = 6 mice). p-values (Student’s paired t-tests) indicated in figure. h, Same as g for peaks in ACh fluorescence. Collectively, these data confirm that spontaneous DA fluctuations in the DLS are driven by the somatic activity of midbrain DANs, whereas spontaneous ACh transients do not require DAN activity. i, Experimental preparation for local pharmacological inhibition of neurotransmitter receptors in the DLS. j, Average global ACh fluorescence measured during immobility after intra-striatal infusion of saline, the mAChR antagonist scopolamine (p = 6.9 x 10−6; n = 5 mice), the D1/2R antagonists SCH23390 and sulpiride (p = 0.24; n = 4 mice) or the iGluR antagonists NBQX and APV (p = 0.41; n = 8 mice; all comparisons are Student’s paired t-test vs. saline control). Data are expressed as a fraction of global ACh fluorescence pre-infusion. Group means (± s.e.m.) are shown in black. Note that mAChR antagonist blocks ACh3.0 (and therefore serves a positive control for maximal possible change in ACh signal) and that blocking iGluRs in the DLS does not significantly alter overall ACh3.0 fluorescence compared to saline, suggesting that CINs continue to release ACh via spontaneous, cell-autonomous firing. k, Same as l for global DA fluorescence after infusion of saline, scopolamine (p = 0.24), D1/2R antagonists (p = 1.1 x 10−4) or iGluR antagonists (p = 0.53). Note that blocking iGluRs in the DLS does not alter overall rDA1m fluorescence either, suggesting that DA axons continue to release DA.

Source Data

Extended Data Fig. 10 Characterization of electrophysiological recordings in the DLS.

a, Experimental preparation for acute in vivo extracellular recordings in the DLS. b, Scatter plot of spike properties used to distinguish units as putative SPNs (pSPNs; gray), putative CINs (pCINs; blue), or other putative interneurons (green). c, Average waveform (left) and auto-correlograms (right) for pSPNs (black), pCINs (blue), and other neurons (green). d–g, Distribution of firing rates (d), coefficient of variation of inter-spike intervals (e), phasic activity index (f), and waveform duration for pSPNs (black), pCINs (blue), and other neurons (green). h, Mean (± s.e.m.) pCIN-pCIN cross correlogram (CCG) computed from all units shown in Fig. 4c (n = 157 pairs). 95% confidence interval shown in gray. i, Proportion of CCGs from Fig. 4c with firing rates above or below the 95% confidence interval (bin size: 20 ms). j, Left, experimental preparation for simultaneous ACh photometry and acute in vivo extracellular recordings from the DLS. Right, location of all recovered fiber optic implant tips (blue) and electrode tracks (red) in the DLS merged across anterior-posterior levels +0.5 to 0.75 mm relative to bregma. k, Men firing rate of two simultaneously recorded pCIN units aligned to ACh fluorescence peaks during immobility. l, Instantaneous (mean normalized) firing rate of pCINs (n = 22) aligned to ACh fluorescence peaks during immobility. Dots indicate units shown in k. m, Average firing rate (normalized to mean; ± s.e.m. shown in shaded region) of all pCINs shown in l. 95% confidence interval shown in gray. n–p, Same as k–m aligned to ACh fluorescence troughs.

Supplementary information

Source data

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Krok, A.C., Maltese, M., Mistry, P. et al. Intrinsic dopamine and acetylcholine dynamics in the striatum of mice. Nature 621, 543–549 (2023). https://doi.org/10.1038/s41586-023-05995-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-023-05995-9

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