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.

  • Letter
  • Published:

A potassium channel β-subunit couples mitochondrial electron transport to sleep

Abstract

The essential but enigmatic functions of sleep1,2 must be reflected in molecular changes sensed by the brain’s sleep-control systems. In the fruitfly Drosophila, about two dozen sleep-inducing neurons3 with projections to the dorsal fan-shaped body (dFB) adjust their electrical output to sleep need4, via the antagonistic regulation of two potassium conductances: the leak channel Sandman imposes silence during waking, whereas increased A-type currents through Shaker support tonic firing during sleep5. Here we show that oxidative byproducts of mitochondrial electron transport6,7 regulate the activity of dFB neurons through a nicotinamide adenine dinucleotide phosphate (NADPH) cofactor bound to the oxidoreductase domain8,9 of Shaker’s KVβ subunit, Hyperkinetic10,11. Sleep loss elevates mitochondrial reactive oxygen species in dFB neurons, which register this rise by converting Hyperkinetic to the NADP+-bound form. The oxidation of the cofactor slows the inactivation of the A-type current and boosts the frequency of action potentials, thereby promoting sleep. Energy metabolism, oxidative stress, and sleep—three processes implicated independently in lifespan, ageing, and degenerative disease6,12,13,14—are thus mechanistically connected. KVβ substrates8,15,16 or inhibitors that alter the ratio of bound NADPH to NADP+ (and hence the record of sleep debt or waking time) represent prototypes of potential sleep-regulatory drugs.

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: Hyperkinetic senses redox changes linked to sleep history.
Fig. 2: dFB-restricted perturbations of redox chemistry alter sleep.
Fig. 3: Optogenetically controlled ROS production in dFB neurons induces sleep.
Fig. 4: Changes in redox chemistry alter the spiking activity of dFB neurons via IA.

Similar content being viewed by others

Code availability

Custom instrument control and analysis code used in this study is available from the corresponding author upon reasonable request.

Data availability

The datasets generated during this study are available from the corresponding author upon reasonable request.

References

  1. Rechtschaffen, A. Current perspectives on the function of sleep. Perspect. Biol. Med. 41, 359–390 (1998).

    Article  CAS  Google Scholar 

  2. Mignot, E. Why we sleep: the temporal organization of recovery. PLoS Biol. 6, e106 (2008).

    Article  Google Scholar 

  3. Donlea, J. M., Thimgan, M. S., Suzuki, Y., Gottschalk, L. & Shaw, P. J. Inducing sleep by remote control facilitates memory consolidation in Drosophila. Science 332, 1571–1576 (2011).

    Article  ADS  CAS  Google Scholar 

  4. Donlea, J. M., Pimentel, D. & Miesenböck, G. Neuronal machinery of sleep homeostasis in Drosophila. Neuron 81, 860–872 (2014).

    Article  CAS  Google Scholar 

  5. Pimentel, D. et al. Operation of a homeostatic sleep switch. Nature 536, 333–337 (2016).

    Article  ADS  CAS  Google Scholar 

  6. Balaban, R. S., Nemoto, S. & Finkel, T. Mitochondria, oxidants, and aging. Cell 120, 483–495 (2005).

    Article  CAS  Google Scholar 

  7. Murphy, M. P. How mitochondria produce reactive oxygen species. Biochem. J. 417, 1–13 (2009).

    Article  CAS  Google Scholar 

  8. Gulbis, J. M., Mann, S. & MacKinnon, R. Structure of a voltage-dependent K+ channel β subunit. Cell 97, 943–952 (1999).

    Article  CAS  Google Scholar 

  9. Long, S. B., Campbell, E. B. & Mackinnon, R. Crystal structure of a mammalian voltage-dependent Shaker family K+ channel. Science 309, 897–903 (2005).

    Article  ADS  CAS  Google Scholar 

  10. Chouinard, S. W., Wilson, G. F., Schlimgen, A. K. & Ganetzky, B. A potassium channel beta subunit related to the aldo-keto reductase superfamily is encoded by the Drosophila Hyperkinetic locus. Proc. Natl Acad. Sci. USA 92, 6763–6767 (1995).

    Article  ADS  CAS  Google Scholar 

  11. Bushey, D., Huber, R., Tononi, G. & Cirelli, C. Drosophila Hyperkinetic mutants have reduced sleep and impaired memory. J. Neurosci. 27, 5384–5393 (2007).

    Article  CAS  Google Scholar 

  12. Orr, W. C. & Sohal, R. S. Extension of life-span by overexpression of superoxide dismutase and catalase in Drosophila melanogaster. Science 263, 1128–1130 (1994).

    Article  ADS  CAS  Google Scholar 

  13. Cirelli, C. et al. Reduced sleep in Drosophila Shaker mutants. Nature 434, 1087–1092 (2005).

    Article  ADS  CAS  Google Scholar 

  14. Hill, V. M. et al. A bidirectional relationship between sleep and oxidative stress in Drosophila. PLoS Biol. 16, e2005206 (2018).

    Article  Google Scholar 

  15. Weng, J., Cao, Y., Moss, N. & Zhou, M. Modulation of voltage-dependent Shaker family potassium channels by an aldo-keto reductase. J. Biol. Chem. 281, 15194–15200 (2006).

    Article  CAS  Google Scholar 

  16. Tipparaju, S. M., Barski, O. A., Srivastava, S. & Bhatnagar, A. Catalytic mechanism and substrate specificity of the β-subunit of the voltage-gated potassium channel. Biochemistry 47, 8840–8854 (2008).

    Article  CAS  Google Scholar 

  17. Sherin, J. E., Shiromani, P. J., McCarley, R. W. & Saper, C. B. Activation of ventrolateral preoptic neurons during sleep. Science 271, 216–219 (1996).

    Article  ADS  CAS  Google Scholar 

  18. Connor, J. A. & Stevens, C. F. Prediction of repetitive firing behaviour from voltage clamp data on an isolated neurone soma. J. Physiol. (Lond.) 213, 31–53 (1971).

    Article  CAS  Google Scholar 

  19. McCormack, T. & McCormack, K. Shaker K+ channel β subunits belong to an NAD(P)H-dependent oxidoreductase superfamily. Cell 79, 1133–1135 (1994).

    Article  CAS  Google Scholar 

  20. Pan, Y., Weng, J., Cao, Y., Bhosle, R. C. & Zhou, M. Functional coupling between the Kv1.1 channel and aldoketoreductase Kvβ1. J. Biol. Chem. 283, 8634–8642 (2008).

    Article  CAS  Google Scholar 

  21. Fogle, K. J. et al. CRYPTOCHROME-mediated phototransduction by modulation of the potassium ion channel β-subunit redox sensor. Proc. Natl Acad. Sci. USA 112, 2245–2250 (2015).

    Article  ADS  CAS  Google Scholar 

  22. Laker, R. C. et al. A novel MitoTimer reporter gene for mitochondrial content, structure, stress, and damage in vivo. J. Biol. Chem. 289, 12005–12015 (2014).

    Article  CAS  Google Scholar 

  23. Maxwell, D. P., Wang, Y. & McIntosh, L. The alternative oxidase lowers mitochondrial reactive oxygen production in plant cells. Proc. Natl Acad. Sci. USA 96, 8271–8276 (1999).

    Article  ADS  CAS  Google Scholar 

  24. Fernandez-Ayala, D. J. M. et al. Expression of the Ciona intestinalis alternative oxidase (AOX) in Drosophila complements defects in mitochondrial oxidative phosphorylation. Cell Metab. 9, 449–460 (2009).

    Article  CAS  Google Scholar 

  25. Wiedau-Pazos, M. et al. Altered reactivity of superoxide dismutase in familial amyotrophic lateral sclerosis. Science 271, 515–518 (1996).

    Article  ADS  CAS  Google Scholar 

  26. Yim, M. B. et al. A gain-of-function of an amyotrophic lateral sclerosis-associated Cu,Zn-superoxide dismutase mutant: An enhancement of free radical formation due to a decrease in K m for hydrogen peroxide. Proc. Natl Acad. Sci. USA 93, 5709–5714 (1996).

    Article  ADS  CAS  Google Scholar 

  27. Artiushin, G. & Sehgal, A. The Drosophila circuitry of sleep–wake regulation. Curr. Opin. Neurobiol. 44, 243–250 (2017).

    Article  CAS  Google Scholar 

  28. Shu, X. et al. A genetically encoded tag for correlated light and electron microscopy of intact cells, tissues, and organisms. PLoS Biol. 9, e1001041 (2011).

    Article  CAS  Google Scholar 

  29. Ng, J. et al. Genetically targeted 3D visualisation of Drosophila neurons under electron microscopy and X-ray microscopy using miniSOG. Sci. Rep. 6, 38863 (2016).

    Article  ADS  CAS  Google Scholar 

  30. Yao, W. D. & Wu, C. F. Auxiliary Hyperkinetic β subunit of K+ channels: regulation of firing properties and K+ currents in Drosophila neurons. J. Neurophysiol. 81, 2472–2484 (1999).

    Article  CAS  Google Scholar 

  31. Donlea, J. M. et al. Recurrent circuitry for balancing sleep need and sleep. Neuron 97, 378–389.e4 (2018).

    Article  CAS  Google Scholar 

  32. Kaplan, W. D. & Trout, W. E., III. The behavior of four neurological mutants of Drosophila. Genetics 61, 399–409 (1969).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Stern, M. & Ganetzky, B. Altered synaptic transmission in Drosophila Hyperkinetic mutants. J. Neurogenet. 5, 215–228 (1989).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  35. Kaneko, M., Park, J. H., Cheng, Y., Hardin, P. E. & Hall, J. C. Disruption of synaptic transmission or clock-gene-product oscillations in circadian pacemaker cells of Drosophila cause abnormal behavioral rhythms. J. Neurobiol. 43, 207–233 (2000).

    Article  CAS  Google Scholar 

  36. Renn, S. C. P., Park, J. H., Rosbash, M., Hall, J. C. & Taghert, P. H. A pdf neuropeptide gene mutation and ablation of PDF neurons each cause severe abnormalities of behavioral circadian rhythms in Drosophila. Cell 99, 791–802 (1999).

    Article  CAS  Google Scholar 

  37. Tanaka, N. K., Tanimoto, H. & Ito, K. Neuronal assemblies of the Drosophila mushroom body. J. Comp. Neurol. 508, 711–755 (2008).

    Article  Google Scholar 

  38. Stocker, R. F., Heimbeck, G., Gendre, N. & de Belle, J. S. Neuroblast ablation in Drosophila P[GAL4] lines reveals origins of olfactory interneurons. J. Neurobiol. 32, 443–456 (1997).

    Article  CAS  Google Scholar 

  39. Lee, T. & Luo, L. Mosaic analysis with a repressible cell marker for studies of gene function in neuronal morphogenesis. Neuron 22, 451–461 (1999).

    Article  CAS  Google Scholar 

  40. Parkes, T. L. et al. Extension of Drosophila lifespan by overexpression of human SOD1 in motorneurons. Nat. Genet. 19, 171–174 (1998).

    Article  CAS  Google Scholar 

  41. Watson, M. R., Lagow, R. D., Xu, K., Zhang, B. & Bonini, N. M. A. A Drosophila model for amyotrophic lateral sclerosis reveals motor neuron damage by human SOD1. J. Biol. Chem. 283, 24972–24981 (2008).

    Article  CAS  Google Scholar 

  42. Anderson, P. R., Kirby, K., Hilliker, A. J. & Phillips, J. P. RNAi-mediated suppression of the mitochondrial iron chaperone, frataxin, in Drosophila. Hum. Mol. Genet. 14, 3397–3405 (2005).

    Article  CAS  Google Scholar 

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

    Article  ADS  CAS  Google Scholar 

  44. Shaw, P. J., Cirelli, C., Greenspan, R. J. & Tononi, G. Correlates of sleep and waking in Drosophila melanogaster. Science 287, 1834–1837 (2000).

    Article  ADS  CAS  Google Scholar 

  45. Hendricks, J. C. et al. Rest in Drosophila is a sleep-like state. Neuron 25, 129–138 (2000).

    Article  CAS  Google Scholar 

  46. Shaw, P. J., Tononi, G., Greenspan, R. J. & Robinson, D. F. Stress response genes protect against lethal effects of sleep deprivation in Drosophila. Nature 417, 287–291 (2002).

    Article  ADS  CAS  Google Scholar 

  47. van Alphen, B., Yap, M. H. W., Kirszenblat, L., Kottler, B. & van Swinderen, B. A dynamic deep sleep stage in Drosophila. J. Neurosci. 33, 6917–6927 (2013).

    Article  Google Scholar 

  48. Connor, J. A. & Stevens, C. F. Voltage clamp studies of a transient outward membrane current in gastropod neural somata. J. Physiol. (Lond.) 213, 21–30 (1971).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank D. Pimentel for electrophysiology advice and C. Chintaluri and M. Murphy for discussions. N. Bonini, B. Dickson, B. Ganetzky, J. Hall, T. Holmes, K. Ito, H. Jacobs, L. Luo, J. Ng, J. Phillips, F. Rouyer, G. Rubin, R. Stocker, P. Taghert, Z. Yan, the Bloomington Stock Center, and the Vienna Drosophila Resource Center provided flies. This work was supported by grants (to G.M.) from the Wellcome Trust and the Gatsby Charitable Foundation. A.K. held postdoctoral fellowships from the Swiss National Science Foundation and EMBO; S.M.S. was a Commonwealth Scholar.

Reviewer information

Nature thanks James Hodge, Yasuo Mori, Michael Palladino and the other anonymous reviewer(s) for their contribution to the peer review of this work.

Author information

Authors and Affiliations

Authors

Contributions

G.M., S.M.S. and A.K. designed the study and analysed the data. A.K. performed electrophysiological recordings and carried out imaging experiments, molecular manipulations, and behavioural analyses with S.M.S. C.B.T. developed instrumentation and code. G.M. directed the research and wrote the paper.

Corresponding author

Correspondence to Gero Miesenböck.

Ethics declarations

Competing interests

A patent application has been filed by G.M., A.K., S.M.S. and Oxford University Innovation Ltd. on the basis of work described in this study.

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 Chronic or acute dFB-restricted perturbations of redox chemistry have no effect on waking locomotor activity or arousability.

a, Locomotor counts per waking minute of flies expressing R23E10-GAL4-driven SOD1 or pro-oxidant SOD1(A4V) do not differ from their respective parental controls (genotype effect: P > 0.2612, Kruskal–Wallis ANOVA with Dunn’s post-hoc test). b, The arousability of flies expressing R23E10-GAL4-driven SOD1 (left) or pro-oxidant SOD1(A4V) (right) does not differ from their respective parental controls (grey colours as in a) (genotype effects: P > 0.2487, vibrational force effects: P < 0.0001, vibrational force × genotype interactions: P > 0.9857, two-way ANOVA). Data are means ± s.e.m. of six trials per genotype (n = 16–32 flies each). c, Locomotor counts per waking minute of flies expressing R23E10-GAL4-driven miniSOG, with or without RNAi transgenes targeting KV channel subunits, and parental controls, in a custom video-tracking system31. Activity was monitored for 10 min before the photo-oxidation of miniSOG and then for a 30-min interval that included an initial 9-min exposure to blue light (genotype effect: P = 0.0827, illumination effect: P = 0.8059, illumination × genotype interaction: P = 0.3086, two-way repeated-measures ANOVA). Data are means ± s.e.m. n, number of flies (a, c) or trials (b). For statistical details see Supplementary Table 2.

Extended Data Fig. 2 Chronic perturbations of redox chemistry in cryptochrome- or pigment dispersing factor (PDF)-expressing clock neurons, Kenyon cells, or olfactory projection neurons have no impact on sleep.

a, Sleep in flies expressing cry-GAL4-driven SOD1 or SOD1(A4V) in clock neurons and parental controls. Kruskal–Wallis ANOVA with Dunn’s post-hoc test failed to detect significant differences of experimental flies from both of their respective parental controls (P > 0.1426). b, Sleep in flies expressing pdf-GAL4-driven SOD1 or SOD1(A4V) in clock neurons and parental controls. Kruskal–Wallis ANOVA with Dunn’s post-hoc test failed to detect significant differences of experimental flies from both of their respective parental controls (P > 0.1732). c, Sleep in flies expressing OK107-GAL4-driven SOD1 or SOD1(A4V) in KCs and parental controls. One-way ANOVA with Holm-Šídák’s post-hoc test failed to detect significant differences of experimental flies from both of their respective parental controls (P > 0.0603). d, Sleep in flies expressing GH146-GAL4-driven SOD1 or SOD1(A4V) in olfactory projection neurons and parental controls. Kruskal–Wallis ANOVA with Dunn’s post-hoc test failed to detect significant differences of experimental flies from both of their respective parental controls (P > 0.6901). Data are means ± s.e.m. n, number of flies. For statistical details see Supplementary Table 2.

Extended Data Fig. 3 Chronic dFB-restricted manipulations of cryptochrome have no impact on sleep.

Sleep in flies expressing two different R23E10-GAL4-driven cry-targeting RNAi transgenes and parental controls. One-way ANOVA with Holm-Šídák’s post-hoc test failed to detect significant differences of experimental flies from both of their respective parental controls (P > 0.1718). Data are means ± s.e.m. n, number of flies. For statistical details see Supplementary Table 2.

Extended Data Fig. 4 Blue illumination of miniSOG-negative dFB neurons has no effect on their electrical activity.

ae, dFB neurons expressing R23E10-GAL4-driven CD8::GFP, before and after a 9-min exposure to blue light. Example voltage responses to current steps (a, sample sizes in b): illumination increases the input resistance (b, Rm; P = 0.0098, paired t-test) but not the membrane time constant (b, τm; P = 0.0723, paired t-test) and leaves unchanged the current–spike frequency function (c, left; current × genotype interaction; P = 0.9982, two-way repeated-measures ANOVA) and interspike interval distribution (c, right; P = 0.0947, Kolmogorov–Smirnov test). Example IA (normalized to peak) evoked by voltage steps to +40 mV (d, sample sizes in e): illumination leaves unchanged the IA amplitude (e; P = 0.8040, Wilcoxon test) and both inactivation time constants (e, τfast: P = 0.6387, τslow: P = 0.2958, Wilcoxon tests). *P < 0.05. Data are means ± s.e.m. n, number of cells. For statistical details see Supplementary Table 2.

Extended Data Table 1 Parameters of IA inactivation

Supplementary information

Supplementary Tables

This file contains Supplementary Tables 1-2.

Reporting Summary

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kempf, A., Song, S.M., Talbot, C.B. et al. A potassium channel β-subunit couples mitochondrial electron transport to sleep. Nature 568, 230–234 (2019). https://doi.org/10.1038/s41586-019-1034-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-019-1034-5

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