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:

Circadian rhythms in neuronal activity propagate through output circuits

Abstract

Twenty-four hour rhythms in behavior are organized by a network of circadian pacemaker neurons. Rhythmic activity in this network is generated by intrinsic rhythms in clock neuron physiology and communication between clock neurons. However, it is poorly understood how the activity of a small number of pacemaker neurons is translated into rhythmic behavior of the whole animal. To understand this, we screened for signals that could identify circadian output circuits in Drosophila melanogaster. We found that leucokinin neuropeptide (LK) and its receptor (LK-R) were required for normal behavioral rhythms. This LK/LK-R circuit connects pacemaker neurons to brain areas that regulate locomotor activity and sleep. Our experiments revealed that pacemaker neurons impose rhythmic activity and excitability on LK- and LK-R-expressing neurons. We also found pacemaker neuron–dependent activity rhythms in a second circadian output pathway controlled by DH44 neuropeptide–expressing neurons. We conclude that rhythmic clock neuron activity propagates to multiple downstream circuits to orchestrate behavioral rhythms.

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: Lk and Lkr mutant circadian phenotypes and location of LK- and LK-R-expressing neurons.
Figure 2: The anatomy of LHLK LK-R neurons suggests they lie downstream of clock neurons.
Figure 3: LHLK neurons are downstream of LNvs.
Figure 4: LHLK neuronal activity is rhythmically regulated by clock neurons.
Figure 5: LHLK activity rhythms propagate to LK-R neurons.
Figure 6: LK and LK-R neuron signaling controls locomotor activity levels.
Figure 7: Clock electrical rhythms propagate through multiple output circuits.

Similar content being viewed by others

References

  1. Herzog, E.D. Neurons and networks in daily rhythms. Nat. Rev. Neurosci. 8, 790–802 (2007).

    Article  CAS  PubMed  Google Scholar 

  2. Nitabach, M.N. & Taghert, P.H. Organization of the Drosophila circadian control circuit. Curr. Biol. 18, R84–R93 (2008).

    Article  CAS  PubMed  Google Scholar 

  3. Colwell, C.S. Linking neural activity and molecular oscillations in the SCN. Nat. Rev. Neurosci. 12, 553–569 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Dibner, C., Schibler, U. & Albrecht, U. The mammalian circadian timing system: organization and coordination of central and peripheral clocks. Annu. Rev. Physiol. 72, 517–549 (2010).

    Article  CAS  PubMed  Google Scholar 

  5. Morin, L.P. Neuroanatomy of the extended circadian rhythm system. Exp. Neurol. 243, 4–20 (2013).

    Article  PubMed  Google Scholar 

  6. Myers, E.M., Yu, J. & Sehgal, A. Circadian control of eclosion: interaction between a central and peripheral clock in Drosophila melanogaster. Curr. Biol. 13, 526–533 (2003).

    Article  CAS  PubMed  Google Scholar 

  7. Xu, K., Zheng, X. & Sehgal, A. Regulation of feeding and metabolism by neuronal and peripheral clocks in Drosophila. Cell Metab. 8, 289–300 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Martin, J.R., Raabe, T. & Heisenberg, M. Central complex substructures are required for the maintenance of locomotor activity in Drosophila melanogaster. J. Comp. Physiol. A 185, 277–288 (1999).

    Article  CAS  PubMed  Google Scholar 

  9. Foltenyi, K., Greenspan, R.J. & Newport, J.W. Activation of EGFR and ERK by rhomboid signaling regulates the consolidation and maintenance of sleep in Drosophila. Nat. Neurosci. 10, 1160–1167 (2007).

    Article  CAS  PubMed  Google Scholar 

  10. Cavanaugh, D.J. et al. Identification of a circadian output circuit for rest:activity rhythms in Drosophila. Cell 157, 689–701 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Martin, J.R., Ernst, R. & Heisenberg, M. Mushroom bodies suppress locomotor activity in Drosophila melanogaster. Learn. Mem. 5, 179–191 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Pitman, J.L., McGill, J.J., Keegan, K.P. & Allada, R. A dynamic role for the mushroom bodies in promoting sleep in Drosophila. Nature 441, 753–756 (2006).

    Article  CAS  PubMed  Google Scholar 

  13. Joiner, W.J., Crocker, A., White, B.H. & Sehgal, A. Sleep in Drosophila is regulated by adult mushroom bodies. Nature 441, 757–760 (2006).

    Article  CAS  PubMed  Google Scholar 

  14. Pírez, N., Christmann, B.L. & Griffith, L.C. Daily rhythms in locomotor circuits in Drosophila involve PDF. J. Neurophysiol. 110, 700–708 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  15. Kunst, M. et al. Calcitonin gene-related peptide neurons mediate sleep-specific circadian output in Drosophila. Curr. Biol. 24, 2652–2664 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Seluzicki, A. et al. Dual PDF signaling pathways reset clocks via TIMELESS and acutely excite target neurons to control circadian behavior. PLoS Biol. 12, e1001810 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  17. Choi, C. et al. Autoreceptor control of peptide/neurotransmitter corelease from PDF neurons determines allocation of circadian activity in Drosophila. Cell Rep. 2, 332–344 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Al-Anzi, B. et al. The leucokinin pathway and its neurons regulate meal size in Drosophila. Curr. Biol. 20, 969–978 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. de Haro, M. et al. Detailed analysis of leucokinin-expressing neurons and their candidate functions in the Drosophila nervous system. Cell Tissue Res. 339, 321–336 (2010).

    Article  PubMed  Google Scholar 

  20. Tanoue, S., Krishnan, P., Krishnan, B., Dryer, S.E. & Hardin, P.E. Circadian clocks in antennal neurons are necessary and sufficient for olfaction rhythms in Drosophila. Curr. Biol. 14, 638–649 (2004).

    Article  CAS  PubMed  Google Scholar 

  21. Bertet, C. et al. Temporal patterning of neuroblasts controls Notch-mediated cell survival through regulation of Hid or Reaper. Cell 158, 1173–1186 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Helfrich-Förster, C. Neurobiology of the fruit fly's circadian clock. Genes Brain Behav. 4, 65–76 (2005).

    Article  PubMed  Google Scholar 

  23. Nicolaï, L.J. et al. Genetically encoded dendritic marker sheds light on neuronal connectivity in Drosophila. Proc. Natl. Acad. Sci. USA 107, 20553–20558 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  24. Zhang, Y.Q., Rodesch, C.K. & Broadie, K. Living synaptic vesicle marker: Synaptotagmin-GFP. Genesis 34, 142–145 (2002).

    Article  CAS  PubMed  Google Scholar 

  25. Gorostiza, E.A., Depetris-Chauvin, A., Frenkel, L., Pírez, N. & Ceriani, M.F. Circadian pacemaker neurons change synaptic contacts across the day. Curr. Biol. 24, 2161–2167 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Herrero, P., Magariños, M., Torroja, L. & Canal, I. Neurosecretory identity conferred by the apterous gene: lateral horn leucokinin neurons in Drosophila. J. Comp. Neurol. 457, 123–132 (2003).

    Article  CAS  PubMed  Google Scholar 

  27. Cognigni, P., Bailey, A.P. & Miguel-Aliaga, I. Enteric neurons and systemic signals couple nutritional and reproductive status with intestinal homeostasis. Cell Metab. 13, 92–104 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Kahsai, L., Martin, J.R. & Winther, A.M. Neuropeptides in the Drosophila central complex in modulation of locomotor behavior. J. Exp. Biol. 213, 2256–2265 (2010).

    Article  CAS  PubMed  Google Scholar 

  29. Lima, S.Q. & Miesenböck, G. Remote control of behavior through genetically targeted photostimulation of neurons. Cell 121, 141–152 (2005).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Sheeba, V. et al. Large ventral lateral neurons modulate arousal and sleep in Drosophila. Curr. Biol. 18, 1537–1545 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Shang, Y., Griffith, L.C. & Rosbash, M. Light-arousal and circadian photoreception circuits intersect at the large PDF cells of the Drosophila brain. Proc. Natl. Acad. Sci. USA 105, 19587–19594 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Yagodin, S., Pivovarova, N.B., Andrews, S.B. & Sattelle, D.B. Functional characterization of thapsigargin and agonist-insensitive acidic Ca2+ stores in Drosophila melanogaster S2 cell lines. Cell Calcium 25, 429–438 (1999).

    Article  CAS  PubMed  Google Scholar 

  34. Shafer, O.T. et al. Widespread receptivity to neuropeptide PDF throughout the neuronal circadian clock network of Drosophila revealed by real-time cyclic AMP imaging. Neuron 58, 223–237 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Gong, Z. et al. Two pairs of neurons in the central brain control Drosophila innate light preference. Science 330, 499–502 (2010).

    Article  CAS  PubMed  Google Scholar 

  36. Frenkel, L. & Ceriani, M.F. Circadian plasticity: from structure to behavior. Int. Rev. Neurobiol. 99, 107–138 (2011).

    Article  PubMed  Google Scholar 

  37. O'Donnell, M.J. et al. Hormonally controlled chloride movement across Drosophila tubules is via ion channels in stellate cells. Am. J. Physiol. 274, R1039–R1049 (1998).

    CAS  PubMed  Google Scholar 

  38. Lelito, K.R. & Shafer, O.T. Reciprocal cholinergic and GABAergic modulation of the small ventrolateral pacemaker neurons of Drosophila's circadian clock neuron network. J. Neurophysiol. 107, 2096–2108 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Cao, G. & Nitabach, M.N. Circadian control of membrane excitability in Drosophila melanogaster lateral ventral clock neurons. J. Neurosci. 28, 6493–6501 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Sheeba, V., Gu, H., Sharma, V.K., O'Dowd, D.K. & Holmes, T.C. Circadian- and light-dependent regulation of resting membrane potential and spontaneous action potential firing of Drosophila circadian pacemaker neurons. J. Neurophysiol. 99, 976–988 (2008).

    Article  PubMed  Google Scholar 

  41. Flourakis, M. et al. A conserved bicycle model for circadian clock control of membrane excitability. Cell 162, 836–848 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Bushey, D., Tononi, G. & Cirelli, C. Sleep- and wake-dependent changes in neuronal activity and reactivity demonstrated in fly neurons using in vivo calcium imaging. Proc. Natl. Acad. Sci. USA 112, 4785–4790 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Petsakou, A., Sapsis, T.P. & Blau, J. Circadian rhythms in Rho1 activity regulate neuronal plasticity and network hierarchy. Cell 162, 823–835 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Pulver, S.R., Pashkovski, S.L., Hornstein, N.J., Garrity, P.A. & Griffith, L.C. Temporal dynamics of neuronal activation by Channelrhodopsin-2 and TRPA1 determine behavioral output in Drosophila larvae. J. Neurophysiol. 101, 3075–3088 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  45. Kitamoto, T. Conditional modification of behavior in Drosophila by targeted expression of a temperature-sensitive shibire allele in defined neurons. J. Neurobiol. 47, 81–92 (2001).

    Article  CAS  PubMed  Google Scholar 

  46. Merighi, A. Costorage and coexistence of neuropeptides in the mammalian CNS. Prog. Neurobiol. 66, 161–190 (2002).

    Article  CAS  PubMed  Google Scholar 

  47. Baines, R.A., Uhler, J.P., Thompson, A., Sweeney, S.T. & Bate, M. Altered electrical properties in Drosophila neurons developing without synaptic transmission. J. Neurosci. 21, 1523–1531 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Gao, X.B. & Horvath, T. Function and dysfunction of hypocretin/orexin: an energetics point of view. Annu. Rev. Neurosci. 37, 101–116 (2014).

    Article  CAS  PubMed  Google Scholar 

  49. Cook, R.K. et al. The generation of chromosomal deletions to provide extensive coverage and subdivision of the Drosophila melanogaster genome. Genome Biol. 13, R21 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Park, J.H. et al. Differential regulation of circadian pacemaker output by separate clock genes in Drosophila. Proc. Natl. Acad. Sci. USA 97, 3608–3613 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Martinek, S., Inonog, S., Manoukian, A.S. & Young, M.W. A role for the segment polarity gene shaggy/GSK-3 in the Drosophila circadian clock. Cell 105, 769–779 (2001).

    Article  CAS  PubMed  Google Scholar 

  52. Plautz, J.D., Kaneko, M., Hall, J.C. & Kay, S.A. Independent photoreceptive circadian clocks throughout Drosophila. Science 278, 1632–1635 (1997).

    Article  CAS  PubMed  Google Scholar 

  53. Calleja, M., Moreno, E., Pelaz, S. & Morata, G. Visualization of gene expression in living adult Drosophila. Science 274, 252–255 (1996).

    Article  CAS  PubMed  Google Scholar 

  54. Lin, D.M. & Goodman, C.S. Ectopic and increased expression of Fasciclin II alters motoneuron growth cone guidance. Neuron 13, 507–523 (1994).

    Article  CAS  PubMed  Google Scholar 

  55. Dietzl, G. et al. A genome-wide transgenic RNAi library for conditional gene inactivation in Drosophila. Nature 448, 151–156 (2007).

    Article  CAS  PubMed  Google Scholar 

  56. Pfeiffer, B.D. et al. Refinement of tools for targeted gene expression in Drosophila. Genetics 186, 735–755 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Zhang, L. et al. DN1p circadian neurons coordinate acute light and PDF inputs to produce robust daily behavior in Drosophila. Curr. Biol. 20, 591–599 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Siegmund, T. & Korge, G. Innervation of the ring gland of Drosophila melanogaster. J. Comp. Neurol. 431, 481–491 (2001).

    Article  CAS  PubMed  Google Scholar 

  59. Grima, B., Chélot, E., Xia, R. & Rouyer, F. Morning and evening peaks of activity rely on different clock neurons of the Drosophila brain. Nature 431, 869–873 (2004).

    Article  CAS  PubMed  Google Scholar 

  60. Yao, Z., Macara, A.M., Lelito, K.R., Minosyan, T.Y. & Shafer, O.T. Analysis of functional neuronal connectivity in the Drosophila brain. J. Neurophysiol. 108, 684–696 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Konopka, R.J. & Benzer, S. Clock mutants of Drosophila melanogaster. Proc. Natl. Acad. Sci. USA 68, 2112–2116 (1971).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Wang, J.W., Beck, E.S. & McCabe, B.D. A modular toolset for recombination transgenesis and neurogenetic analysis of Drosophila. PLoS One 7, e42102 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Jenett, A. et al. A GAL4-driver line resource for Drosophila neurobiology. Cell Rep. 2, 991–1001 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Rideout, E.J., Dornan, A.J., Neville, M.C., Eadie, S. & Goodwin, S.F. Control of sexual differentiation and behavior by the doublesex gene in Drosophila melanogaster. Nat. Neurosci. 13, 458–466 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Collins, B. et al. Differentially timed extracellular signals synchronize pacemaker neuron clocks. PLoS Biol. 12, e1001959 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  66. Cyran, S.A. et al. The Double-time protein kinase regulates the subcellular localization of the Drosophila clock protein Period. J. Neurosci. 25, 5430–5437 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Jang, A.R., Moravcevic, K., Saez, L., Young, M.W. & Sehgal, A. Drosophila TIM binds importin α1, and acts as an adapter to transport PER to the nucleus. PLoS Genet. 11, e1005205 (2015).

    Article  CAS  Google Scholar 

  68. Glossop, N.R. et al. VRILLE feeds back to control circadian transcription of Clock in the Drosophila circadian oscillator. Neuron 37, 249–261 (2003).

    Article  CAS  PubMed  Google Scholar 

  69. Ruben, M., Drapeau, M.D., Mizrak, D. & Blau, J. A mechanism for circadian control of pacemaker neuron excitability. J. Biol. Rhythms 27, 353–364 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Stewart, B.A., Atwood, H.L., Renger, J.J., Wang, J. & Wu, C.F. Improved stability of Drosophila larval neuromuscular preparations in haemolymph-like physiological solutions. J. Comp. Physiol. A 175, 179–191 (1994).

    Article  CAS  PubMed  Google Scholar 

  71. McCarthy, E.V. et al. Synchronized bilateral synaptic inputs to Drosophila melanogaster neuropeptidergic rest/arousal neurons. J. Neurosci. 31, 8181–8193 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank B. Al-Anzi (California Institute of Technology), D. Clark (Yale University), P. Hardin (Texas A & M University), P. Herrero (Universidad Autónoma de Madrid), M. Rosbash (Brandeis University), F. Rouyer (Institut des Neurosciences Paris-Saclay), A. Sehgal (University of Pennsylvania School of Medicine), O. Shafer (University of Michigan), S. Sweeney (University of York), the Developmental Studies Hybridoma Bank and the Bloomington stock center for flies and antibodies. We thank the TRiP stock center at Harvard Medical School (NIH/NIGMS R01-GM084947) for transgenic RNAi fly stocks. We thank C. Desplan for sharing the two-photon microscope and perfusion chamber and O. Shafer for advice on calcium imaging. We thank R. Behnia, C. Desplan, C. Hackley, E. Meekhof, A. Petsakou, H. Piggins and Z. Zhu for discussions and comments on the manuscript. This investigation was conducted in facilities constructed with support from Research Facilities Improvement Grant Number C06 RR-15518-01 from the National Center for Research Resources, US National Institutes of Health (NIH). Imaging was performed at the NYU Center for Genomics & Systems Biology. This work was supported by EMBO ALTF 249-2009 (M.C.), the Charles H. Revson foundation (M.C.), EMBO ALTF 680-2009 (C.B.), HSFPO LT000077/2010-l (C.B.), NIH grant R01 EY017916 (to C. Desplan), NIH grant GM063911 (J.B.) and the NYU Abu Dhabi Research Institute (G1205).

Author information

Authors and Affiliations

Authors

Contributions

M.C. and B.C. performed the RNAi screen. M.C. performed all other experiments and analyses except the immunostaining in Supplementary Figure 3, which was done by C.B. M.C. and J.B. wrote the manuscript, with comments from B.C. and C.B.

Corresponding author

Correspondence to Justin Blau.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Characterization of Lk and Lkr mutants.

(a) Representative actograms in DD for three Lkr mutant alleles (Lkrc003, LkrMimic1 and LkrMimic2) as heterozygotes over a wild type chromosome (top row) or as hemizygotes over a deficiency (Df(3L)BSC371, bottom row) that removes Lkr. All flies are in a w1118 background. See Supplementary Table 2 for details and sample sizes. (b-c) Quantification of LK peptide levels in Lkc275 and elav > Dcr-2 + LkRNAi mutants by immunostaining. (b) Images are z-projections, color-coded to reflect pixel intensity (scale on right). LHLK cell bodies are indicated by arrowheads and shown in insets acquired at higher laser power to reveal faint LK staining in flies expressing LkRNAi. Scale bars: 20μm (c) Quantification of immunostaining. LK levels are expressed as a fraction of control (y w) levels and are indicated above each bar. Error bars indicate SEM. n=16 cell bodies from 8 brains for each sample from 1 experiment. (d) Quantification of Lkr mRNA reduction in Lkrc003 and elav > Dcr-2 + LkrRNAi mutants by quantitative real-time RT-PCR. Lkr mRNA levels are expressed as a fraction of control (y w) levels and indicated above each bar. Calmodulin was used to normalize input RNA levels. Error bars indicate SEM. n=6 replicates for each sample (see Methods).

Source data

Supplementary Figure 2 LK and LK-R are not clock neurons and the molecular clock is intact in LkRNAi mutants.

(a) LHLK neurons have no detectable GFP when UAS-nlsGFP was expressed via per-Gal4. This is despite broad GFP expression in the brain, including clock neurons identified by immunostaining (left panel, LNvs labeled by PDF) or anatomically (middle panel, DN1s and DN3s). per-Gal4 is absent from LHLKs (middle panel) and other LK neurons (SELKs, right panel) labeled by LK immunostaining. The middle and right panels are single confocal sections. Scale bar 100μm (left panel), 20μm in middle and right panels. n=8 brains examined. (b) Functional test for the presence of a molecular clock in LK and LK-R neurons. Locomotor activity rhythm strength (power) of flies over 10 days in DD after entrainment to LD cycles. Asterisks indicate significant difference (p<0.05) with the control by KS test, ns: non-significant. Expressing a cycle dominant negative construct (UAS-cycΔ) in clock neurons with tim-Gal4 (2nd bar) causes severe arrhythmicity compared to controls (UAS-cycΔ / +, 1st bar, p<0.0001, D=0.878). In contrast, UAS-cycΔ has no effect when expressed with Lk-Gal4 (3rd bar, p=0.449, D=0.238) or LkrR65C07-Gal4 (4th bar, p=0.735, D=0.186). See Supplementary Table 2 for details and sample sizes. (c-e) The molecular clock is intact in Lk mutants. (c) PDF, TIMELESS (TIM) and VRILLE (VRI) immunostaining in control (left, y w) and Lk mutant (right, elav > Dcr-2 + LkRNAi) brains dissected at 4 time points on the 2nd and 3rd days in DD. PDF staining marks s-LNvs. The brightness and contrast have been adjusted for TIM and VRI to reveal faint staining in all images. Scale bars 20μm. (d-e) Quantification of non-adjusted images reveal that TIM (d) and VRI (e) oscillations are similar in control and LkRNAi mutants. Statistics by Kruskal-Wallis ANOVA (3d.f each): y w TIM, p<0.0001, H=149.8; y w VRI, p<0.0001, H=89.6; LkRNAi TIM, p<0.0001, H=175.2; LkRNAi VRI, p<0.0001, H=175.1. Error bars indicate SEM. Sample sizes (indicated as n=x;y where x is the number of neurons and y the number of brains): y w CT12 n=60;10, y w CT18 n=73;10, y w CT24 n=64;10, y w CT6 n=60;9, LkRNAi CT12 n=61;10, LkRNAi CT18 n=68;10, LkRNAi CT24 n=65;10, LkRNAi CT6 n=75;10.

Source data

Supplementary Figure 3 Characterization of the LkrR65C07-Gal4 line.

(a-e) LkrR65C07-Gal4 labels neurons with a very similar anatomy to Lkr > FLEXAMP clones that also likely connect to LHLKs. (a) LkrR65C07-Gal4 driving CD8::GFP co-stained for LK labels neurons with their cell bodies in the lateral horn (LH) region and PI. Lateral horn LK-RR65C07 neurons project towards the FSB as observed in Fig. 2g with Lkr > FLEXAMP. (b) Single confocal section of the dashed square in a show overlap of LK-RR65C07 projections with LHLK arborizations and numerous potential sites of contact. (c) LK-RR65C07 neurons innervate several layers of the FSB. (d) LkrR65C07-Gal4 is also expressed in EB neurons. (e) LkrR65C07-Gal4 driving Syt::GFP co-stained for LK shows that LK-RR65C07 neuron synaptic terminals are found predominantly in the FSB. (f) Single confocal sections showing LK-RR65C07 projections in the FSB intermingle with neurons with previously characterized functions in locomotion and sleep. LkrR65C07-LexA driving CD8::GFP crossed to Gal4 lines driving CD8::RFP (121Y-Gal4, c5-Gal4 and c584-Gal4). Scale bars: 20μm, except in a and e: 100μm. n≥8 brains examined for each experiment.

Supplementary Figure 4 Calcium-imaging control experiments

All statistics by Kolmogorov-Smirnov (KS) test as in Fig. 3. Each experiment is from ≥2 independent replicates. (a, left) Perfusing 2.5 mM ATP significantly increases intracellular calcium in s-LNvs (n=19;4) and l-LNvs (n=26;4) of Pdf > P2X2 + GCaMP6S brains compared to 0.1% DMSO vehicle perfusion (the vehicle control is shown for s-LNvs, veh, n=22;3). p<0.0001, D=0.909 for ATP perfusion vs. veh in s-LNvs; p<0.0001, D=0.969 for ATP vs veh (n=32;3) in l-LNvs. (a, middle) No increase in intracellular calcium in LHLK neurons from Lk > GCaMP6S; Pdf-LexA > P2X2 brains with ATP perfusion (n=12;3) or vehicle (n=16;4). No statistical difference for ATP vs. veh (p=0.948, D=0.187). (a, right) Dot plots show distribution of Maximum ΔF/F0 in left and middle line graphs. (b, left) Dose response of LHLK neurons to decreasing concentrations of CCh. From top to bottom: 100 μM CCh, n=23;11; 50 μM CCh, n=23;7; 10 μM CCh, n=61;17; vehicle, n=12;4. (b, middle) LHLK response to 10 μM CCh appears to be direct since it was unaffected by 20 min pre-incubation with 2 μM TTX (n=28;7) compared to controls (veh, n=22;6, p=0.326, D=0.259). (b, right) Quantification of line graph data. For CCh dose responses, asterisks indicate a significant difference from the vehicle control by KS test (100 μM: p<0.0001, D=0.937; 50 μM: p<0.0001, D=0.937; 10 μM: p<0.0001, D=0.711). (c, left) Perfusing 10 μM PDF increases calcium gradually in s-LNvs (n=18;6) but not in l-LNvs (n=17;6). The vehicle control shown is for l-LNvs, veh, n=15;6). (c, middle) Perfusing 10 μM PDF has no effect on LHLK neurons (n=10;9, veh: n=14;7). (c, right) Quantification of line graph data. Statistics: For s-LNvs veh vs PDF: p=0.0057, D=0.567 (veh, n=15;6). For l-LNvs veh vs PDF: p=0.571, D=0.263. For LHLKs veh vs PDF: p=0.060, D=0.5. (d, left) PDF pre-incubation decreases LHLK response to 10 μM CCh as in Fig. 3c (20min 20μM PDF pre-incubation, n=38;19, veh: n=28;14, p=0.016, D=0.372). (d, middle) PDF's inhibitory effect is transient since it disappears after 15 min washout. (n=38;19, veh: n=28;14, p=0.312, D=0.231). (d, right) Quantification of line graph data. (e) Perfusing 100 μM LK (n=15;5) does not increase calcium levels in LK-R neurons in the lateral horn region in Lkr > GCaMP6S brains compared to vehicle control (n=4;2, p=0.985, D=0.233), quantified in dot plot. (f) 10 min pre-incubation with 100 μM LK does not change l-LNv responses to 100 μM CCh (n=21;7; veh: n=15;6, p=0.916, D=0.190), quantified in dot plot. (g) LK-R neurons labeled by the LkrR65C07-Gal4 line with cell bodies in the lateral horn and projecting to the FSB are inhibited by 100 μM LK pre-incubation as LK-R neurons in Fig. 3e (response to 100 μM CCh, p<0.0001, D=0.562; n=80;4 each). (h) LK-RR65C07 neuron baseline GCaMP6S levels are rhythmic in DD (p<0.0001, H=41.41, 2d.f. by Kruskal-Wallis one-way ANOVA; CT0 n=300;12, CT11 n=318;12, CT23 n=272;12).

Source data

Supplementary Figure 5 LK and LK-RR65C07 neuron manipulations affect sleep and locomotor activity levels.

(a-d) Sleep and locomotor activity for the data shown in Fig. 6b. (e-f) Sleep for the data shown in Fig. 6c. Sample sizes and replicates are as in Fig. 6b-c, asterisks: experimental flies significantly different from both parental controls (p<0.05) by KS test as in Fig. 6, error bars in c-d: SEM. (a-b) Sleep profiles (average minutes sleeping per 30 min bin) over 3 days in DD (days indicated at bottom) with a 24hr TrpA1 activation on day 2 (red shaded area, 28°C) as in Fig. 6b-c. Control flies decreased sleep in response to heat. (a) In contrast, Lk > TrpA1 flies increased sleep while (b) LkrR65C07 > TrpA1 flies decreased sleep only slightly further than control flies, probably due to the already low sleep levels at this time of day at 28°C. The first 12hr of heat pulse are magnified in the inset. (c-d) Waking activity (average beam crossing per awake minute) calculated for 3 consecutive days over 6hr time windows. (c) Activating LK neurons caused flies to be significantly less active than control flies while awake. (d) Activating LK-RR65C07 neurons caused flies to be significantly more active than controls while awake during the first 6hr time window. Waking activity recovered to normal levels on subsequent days in both experiments although it was more variable in d than c. (e-f) Sleep profiles of flies with (e) LK or (f) LK-RR65C07 neurons inhibited with shits. (e) LK inhibition had almost no effect on sleep, whereas (f) LK-RR65C07 inhibition significantly increased sleep throughout subjective day and night.

Supplementary Figure 6 Lk>TrpA1 and LkrR65C07>TrpA1 activation phenotypes are modulated by light and require neuronal TrpA1 expression.

(a) TrpA1 activation / shits inhibition experiments performed in DD and LD. The first 12hr of heat pulse are shown for TrpA1 activation of (left) LK neurons, (middle) LK-RR65C07 neurons and (right) shits inhibition of LK-RR65C07 neurons. Upper panels show experiments performed in DD (same data as in Fig. 6b-c), lower panels show the same experiments performed in LD (from 2 independent replicates). Heat pulses start at CT0 or ZT0. Asterisks: p<0.05 by KS test as in Fig. 6b-c, error bars: SEM. (left) Light suppresses/delays the decreased locomotor activity phenotype induced by LK neuron activation for most of the day. However the phenotype is apparent during the last ~4hr of the day (~ZT8-12). (middle) Similarly, light delays the increase in locomotor activity induced by LK-RR65C07 neuronal activation seen during the first 6hr in DD. The phenotype becomes apparent during the second half of the day (~ZT7-10). (right) Light delays the phenotype caused by LK-RR65C07 neuron inhibition with Shits as with activation experiments. Decreased locomotor activity becomes apparent during the last ~5hr of the day (~ZT7-12). Sample sizes for DD experiments are the same as in Fig 6b-c. Sample sizes for LD experiments: Lk > + n=61; TrpA1 / + n=62; Lk > TrpA1 n=60; LkrR65C07 > + n=63; LkrR65C07 > TrpA1 n=62; Shits / + n=57; LkrR65C07 > Shits n=52. (b) Restricting UAS-TrpA1 expression to non-neuronal cells using elav-Gal80 overrides the locomotor activity phenotypes caused by acutely activating LK and LK-RR65C07 neurons. Population average locomotor activity for 3 days in DD (days indicated at bottom, grey and black bars indicate subjective day and night) with 24hr of TrpA1 activation on day 2 (red shaded area, 28°C). The changes in locomotor activity caused by activating LK (left) and LK-RR65C07 (right) neurons are rescued by including an elav-Gal80 transgene. Thus the locomotor activity phenotype is not due to stimulating diuresis via non-neuronal LK-R-expressing stellate cells in the Malphigian tubules. Sample sizes: (left) UAS-TrpA1 / + n=62, elav-Gal80 /+; UAS-TrpA1 / + n=51, Lk > TrpA1 n=63, Lk; elav-Gal80 > TrpA1 n=51, (right) UAS-TrpA1 / + n=94, elav-Gal80 /+; UAS-TrpA1 / + n=75, LkrR65C07 > TrpA1 n=93, LkrR65C07 / elav-Gal80 > TrpA1 n=89. Data are from 2 (left) or 3 (right) independent experiments.

Source data

Supplementary Figure 7 Model for propagation of clock electrical rhythms through output circuits to drive circadian behavior.

Electrical rhythms originate in the clock network by interplay of the intracellular molecular clocks and communication between different clock neuron groups. For simplicity, only Morning clock neurons (s-LNvs and DN1ps; M cells, blue) are shown here. M cell activity peaks around dawn and is low around dusk (blue curve). Clock neuron activity rhythms produce 24hr oscillations in the strength of clock outputs that propagate into output circuits. LHLK neuronal activity (orange) is in antiphase with M cells since M cells inhibit LHLKs. Following the same principle, neuronal activity rhythms propagate to the next layer of the LK circuit, the LK-R neurons (brown), and the phase is reversed again since LK inhibits LK-R neurons. DH44 (green) activity is also in antiphase with M cell and the relevant subset of DH44-Receptor neurons (DH44-R) have not yet been identified. Additional output circuits include DH31-Receptor neurons (DH31-R) and PDFR-expressing neurons in the EB. Ultimately, the rhythmic activity of output circuits transfers to brain areas controlling locomotor activity and sleep, generating behavioral rhythms.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–7 and Supplementary Tables 1 and 2 (PDF 1193 kb)

Supplementary Methods Checklist (PDF 460 kb)

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cavey, M., Collins, B., Bertet, C. et al. Circadian rhythms in neuronal activity propagate through output circuits. Nat Neurosci 19, 587–595 (2016). https://doi.org/10.1038/nn.4263

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.4263

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