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:

State-dependent decoupling of sensory and motor circuits underlies behavioral flexibility in Drosophila

Abstract

An approaching predator and self-motion toward an object can generate similar looming patterns on the retina, but these situations demand different rapid responses. How central circuits flexibly process visual cues to activate appropriate, fast motor pathways remains unclear. Here we identify two descending neuron (DN) types that control landing and contribute to visuomotor flexibility in Drosophila. For each, silencing impairs visually evoked landing, activation drives landing, and spike rate determines leg extension amplitude. Critically, visual responses of both DNs are severely attenuated during non-flight periods, effectively decoupling visual stimuli from the landing motor pathway when landing is inappropriate. The flight-dependence mechanism differs between DN types. Octopamine exposure mimics flight effects in one, whereas the other probably receives neuronal feedback from flight motor circuits. Thus, this sensorimotor flexibility arises from distinct mechanisms for gating action-specific descending pathways, such that sensory and motor networks are coupled or decoupled according to the behavioral state.

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: Two identified DN types contribute to visually evoked landing responses in Drosophila.
Fig. 2: Landing DNs respond to visual stimuli and control leg extension amplitude.
Fig. 3: Visual responses of landing DNs are gated by flight.
Fig. 4: Decoupling of visual and motor circuits is achieved by different mechanisms in the two landing-DN types.

Similar content being viewed by others

Data availability

The data and analysis code that support the findings of this study are available from the corresponding author upon reasonable request. The leg tracking software was deposited to https://github.com/kristinbranson/APT.

References

  1. Wu, M. et al. Visual projection neurons in the Drosophila lobula link feature detection to distinct behavioral programs. eLife 5, e21022 (2016).

    Article  Google Scholar 

  2. Panser, K. et al. Automatic segmentation of Drosophila neural compartments using GAL4 expression data reveals novel visual pathways. Curr. Biol. 26, 1943–1954 (2016).

    Article  CAS  Google Scholar 

  3. Klapoetke, N. C. et al. Ultra-selective looming detection from radial motion opponency. Nature 551, 237–241 (2017).

    Article  CAS  Google Scholar 

  4. von Reyn, C. R. et al. Feature integration drives probabilistic behavior in the Drosophila escape response. Neuron 94, 1190–1204 (2017).

    Article  Google Scholar 

  5. Ache, J. M. et al. Neural basis for looming size and velocity encoding in the Drosophila giant fiber escape pathway. Curr. Biol. 29, 1073–1081.e4 (2019).

    Article  CAS  Google Scholar 

  6. von Reyn, C. R. et al. A spike-timing mechanism for action selection. Nat. Neurosci. 17, 962–970 (2014).

    Article  Google Scholar 

  7. Tammero, L. F. & Dickinson, M. H. Collision-avoidance and landing responses are mediated by separate pathways in the fruit fly, Drosophila melanogaster. J. Exp. Biol. 205, 2785–2798 (2002).

    PubMed  Google Scholar 

  8. van Breugel, F. & Dickinson, M. H. The visual control of landing and obstacle avoidance in the fruit fly Drosophila melanogaster. J. Exp. Biol. 215, 1783–1798 (2012).

    Article  Google Scholar 

  9. Branson, K., Robie, A. A., Bender, J., Perona, P. & Dickinson, M. H. High-throughput ethomics in large groups of Drosophila. Nat. Methods 6, 451–457 (2009).

    Article  CAS  Google Scholar 

  10. Kim, L. H. et al. Integration of descending command systems for the generation of context-specific locomotor behaviors. Front. Neurosci. 11, 581 (2017).

    Article  Google Scholar 

  11. Muijres, F. T., Elzinga, M. J., Melis, J. M. & Dickinson, M. H. Flies evade looming targets by executing rapid visually directed banked turns. Science 344, 172–177 (2014).

    Article  CAS  Google Scholar 

  12. Yilmaz, M. & Meister, M. Rapid Innate defensive responses of mice to looming visual stimuli. Curr. Biol. 23, 2011–2015 (2013).

    Article  CAS  Google Scholar 

  13. Card, G. & Dickinson, M. H. Visually mediated motor planning in the escape response of Drosophila. Curr. Biol. 18, 1300–1307 (2008).

    Article  CAS  Google Scholar 

  14. Borst, A. Time course of the houseflies’ landing response. Biol. Cybern. 54, 379–383 (1986).

    Article  Google Scholar 

  15. Borst, A. & Bahde, S. What kind of movement detector is triggering the landing response of the housefly? Biol. Cybern. 55, 65–69 (1986).

    Article  Google Scholar 

  16. Bacon, J. P. & Strausfeld, N. J. The dipteran ‘Giant fibre’ pathway: neurons and signals. J. Comp. Physiol. A 158, 529–548 (1986).

    Article  Google Scholar 

  17. Namiki, S., Dickinson, M. H., Wong, A. M., Korff, W. & Card, G. M. The functional organization of descending sensory-motor pathways in Drosophila. eLife 7, e34272 (2018).

    Article  Google Scholar 

  18. Goodman, L. J. The landing responses of insects. J. Exp. Biol. 37, 854–878 (1960).

    Google Scholar 

  19. 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  Google Scholar 

  20. Borst, A. & Bahde, S. Visual information processing in the fly’s landing system. J. Comp. Physiol. A 163, 167–173 (1988).

    Article  Google Scholar 

  21. May, M. L. & Hoy, R. R. Leg–induced steering in flying crickets. J. Exp. Biol. 151, 485–488 (1990).

    Google Scholar 

  22. Nolen, T. G. & Hoy, R. R. Initiation of behavior by single neurons: the role of behavioral context. Science 226, 992–994 (1984).

    Article  CAS  Google Scholar 

  23. Huston, S. J. & Krapp, H. G. Nonlinear integration of visual and haltere inputs in fly neck motor neurons. J. Neurosci. 29, 13097–13105 (2009).

    Article  CAS  Google Scholar 

  24. Bacon, J. P., Thompson, K. S. & Stern, M. Identified octopaminergic neurons provide an arousal mechanism in the locust brain. J. Neurophysiol. 74, 2739–2743 (1995).

    Article  CAS  Google Scholar 

  25. Suver, M. P., Mamiya, A. & Dickinson, M. H. Octopamine neurons mediate flight-induced modulation of visual processing in Drosophila. Curr. Biol. 22, 2294–2302 (2012).

    Article  CAS  Google Scholar 

  26. Tuthill, J. C., Nern, A., Rubin, G. M. & Reiser, M. B. Wide-field feedback neurons dynamically tune early visual processing. Neuron 82, 887–895 (2014).

    Article  CAS  Google Scholar 

  27. Riemensperger, T. et al. Behavioral consequences of dopamine deficiency in the Drosophila central nervous system. Proc. Natl Acad. Sci. USA 108, 834–839 (2011).

    Article  CAS  Google Scholar 

  28. Mamiya, A., Straw, A. D., Tomasson, E. & Dickinson, M. H. Active and passive antennal movements during visually guided steering in flying Drosophila. J. Neurosci. 31, 6900–6914 (2011).

    Article  CAS  Google Scholar 

  29. Ramirez, J. M. Interneurons in the suboesophageal ganglion of the locust associated with flight initiation. J. Comp. Physiol. A 162, 669–685 (1988).

    Article  Google Scholar 

  30. Maimon, G. Modulation of visual physiology by behavioral state in monkeys, mice, and flies. Curr. Opin. Neurobiol. 21, 559–564 (2011).

    Article  CAS  Google Scholar 

  31. Suver, M. P., Huda, A., Iwasaki, N., Safarik, S. & Dickinson, M. H. An array of descending visual interneurons encoding self-motion in Drosophila. J. Neurosci. 36, 11768–11780 (2016).

    Article  CAS  Google Scholar 

  32. Zorovic, M. & Hedwig, B. Processing of species-specific auditory patterns in the cricket brain by ascending, local, and descending neurons during standing and walking. J. Neurophysiol. 105, 2181–2194 (2011).

    Article  CAS  Google Scholar 

  33. Hoy, R., Nolen, T. & Brodfuehrer, P. The neuroethology of acoustic startle and escape in flying insects. J. Exp. Biol. 146, 287–306 (1989).

    CAS  PubMed  Google Scholar 

  34. Staudacher, E. & Schildberger, K. Gating of sensory responses of descending brain neurones during walking in crickets. J. Exp. Biol. 201, 559–572 (1998).

    Google Scholar 

  35. Haag, J., Wertz, A. & Borst, A. Central gating of fly optomotor response. Proc. Natl Acad. Sci. USA 107, 20104–20109 (2010).

    Article  CAS  Google Scholar 

  36. Kozlov, A. K., Kardamakis, A. A., Hellgren Kotaleski, J. & Grillner, S. Gating of steering signals through phasic modulation of reticulospinal neurons during locomotion. Proc. Natl Acad. Sci. USA 111, 3591–3596 (2014).

    Article  CAS  Google Scholar 

  37. Grillner, S., Hellgren, J., Ménard, A., Saitoh, K. & Wikström, M. A. Mechanisms for selection of basic motor programs—roles for the striatum and pallidum. Trends Neurosci. 28, 364–370 (2005).

    Article  CAS  Google Scholar 

  38. Maimon, G., Straw, A. D. & Dickinson, M. H. Active flight increases the gain of visual motion processing in Drosophila. Nat. Neurosci. 13, 393–399 (2010).

    Article  CAS  Google Scholar 

  39. Chiappe, M. E., Seelig, J. D., Reiser, M. B. & Jayaraman, V. Walking modulates speed sensitivity in Drosophila motion vision. Curr. Biol. 20, 1470–1475 (2010).

    Article  CAS  Google Scholar 

  40. Coen, P., Xie, M., Clemens, J. & Murthy, M. Sensorimotor transformations underlying variability in song intensity during Drosophila courtship. Neuron 89, 629–644 (2016).

    Article  CAS  Google Scholar 

  41. Ribeiro, I. M. A. et al. Visual projection neurons mediating directed courtship in Drosophila. Cell 174, 607–621.e18 (2018).

    Article  CAS  Google Scholar 

  42. Fotowat, H., Fayyazuddin, A., Bellen, H. J. & Gabbiani, F. A novel neuronal pathway for visually guided escape in Drosophila melanogaster. J. Neurophysiol. 102, 875–885 (2009).

    Article  Google Scholar 

  43. Sanes, J. R. & Zipursky, S. L. Design principles of insect and vertebrate visual systems. Neuron 66, 15–36 (2010).

    Article  CAS  Google Scholar 

  44. Tovote, P. et al. Midbrain circuits for defensive behaviour. Nature 534, 206–212 (2016).

    Article  CAS  Google Scholar 

  45. Pfeiffer, B. D., Truman, J. W. & Rubin, G. M. Using translational enhancers to increase transgene expression in Drosophila. Proc. Natl Acad. Sci. USA 109, 6626–6631 (2012).

    Article  CAS  Google Scholar 

  46. Klapoetke, N. C. et al. Independent optical excitation of distinct neural populations. Nat. Methods 11, 338–346 (2014).

    Article  CAS  Google Scholar 

  47. Tuthill, J. C., Nern, A., Holtz, S. L., Rubin, G. M. & Reiser, M. B. Contributions of the 12 neuron classes in the fly lamina to motion vision. Neuron 79, 128–140 (2013).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  49. Nern, A., Pfeiffer, B. D. & Rubin, G. M. Optimized tools for multicolor stochastic labeling reveal diverse stereotyped cell arrangements in the fly visual system. Proc. Natl Acad. Sci. USA 112, E2967–E2976 (2015).

    Article  CAS  Google Scholar 

  50. Gouwens, N. W. & Wilson, R. I. Signal propagation in Drosophila central neurons. J. Neurosci. 29, 6239–6249 (2009).

    Article  CAS  Google Scholar 

  51. Lindsay, T., Sustar, A. & Dickinson, M. The function and organization of the motor mystem controlling flight maneuvers in flies. Curr. Biol. 27, 345–358 (2017).

    Article  CAS  Google Scholar 

  52. Reiser, M. B. & Dickinson, M. H. A modular display system for insect behavioral neuroscience. J. Neurosci. Methods 167, 127–139 (2008).

    Article  Google Scholar 

  53. Williamson, W. R., Peek, M. Y., Breads, P., Coop, B. & Card, G. M. Tools for rapid high-resolution behavioral phenotyping of automatically isolated Drosophila. Cell Rep. 25, 1636–1649.e5 (2018).

    Article  CAS  Google Scholar 

  54. Brainard, D. H. The psychophysics toolbox. Spat. Vis. 10, 433–436 (1997).

    Article  CAS  Google Scholar 

  55. Gabbiani, F., Krapp, H. G. & Laurent, G. Computation of object approach by a wide-field, motion-sensitive neuron. J. Neurosci. 19, 1122–1141 (1999).

    Article  CAS  Google Scholar 

  56. Ian, O. A multifunctional role for octopamine in locust flight. Annu. Rev. Entomol. 38, 227–250 (1993).

    Article  Google Scholar 

  57. Burgos-Artizzu, X. P., Perona, P. & Dollar, P. Robust face landmark estimation under occlusion. In Proc. IEEE International Conference on Computer Vision 1513–1520 (IEEE, 2013).

  58. Ozuysal, M., Calonder, M., Lepetit, V. & Fua, P. Fast keypoint recognition using random ferns. IEEE Trans. Pattern Anal. Mach. Intell. 32, 448–461 (2010).

    Article  Google Scholar 

Download references

Acknowledgements

We thank K. von Reyn, M. Peek and S. Huston for sharing their rigs and providing software; M. Dickinson and T. Lindsay for building the Kinefly setup; M. Sumathipala, D. Parikh and P. Breads for labeling fly videos; and W. Korff and the Janelia Descending Interneuron Project for help with the DN split-GAL4 screen. We are grateful to G. Zheng and the Janelia Fly Core for fly husbandry, and to E. Gruntman and M. Reiser for sharing LED-arena code and advice. V. Jayaraman, K. Longden, E. Gruntman, C. Dallmann and Card laboratory members provided helpful comments on the manuscript. This study was supported by the Howard Hughes Medical Institute.

Author information

Authors and Affiliations

Authors

Contributions

J.M.A. conceived the project, designed experiments, conducted patch-clamp recordings, conducted behavioral and optogenetic activation experiments, wrote analysis code, analyzed the data, generated the figures and wrote the manuscript. S.N. designed experiments, conducted and analyzed behavioral optogenetic activation experiments, generated split-GAL4 driver lines and contributed anatomy figures. A.L. and K.B. developed tracking software. G.M.C. conceived the project, designed experiments, wrote the manuscript and provided supervision.

Corresponding author

Correspondence to Gwyneth M. Card.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Journal peer review information: Nature Neuroscience thanks Rachel Wilson and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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

Integrated supplementary information

Supplementary Figure 1 DNp07 and DNp10 morphology and split-Gal4 line expression patterns.

(a) Central nervous system expression driven by the DNp10-split-GAL4 line (SS01608) was limited to only the left and right copies of DNp10. Insets in a and b show optical slices through the optic glomerulus region of the brain, indicated by gray boxes. Both DNs have dendrites close to optic glomeruli of the LPLC3/4 types. Scale bars, 50 µm. (b) Central nervous system expression pattern for DNp07-split-GAL4-2 (SS01549), used in activation and silencing experiments, which expressed strongly in DNp07, and had weaker expression in a second DN (DNp13). (c) Central nervous system expression pattern for DNp07-split-GAL4-1 (SS02276), used in activation and silencing experiments, which had off-target expression in the abdominal ganglia, but did not express in other DNs. Panels a, b, c, CsChrimson-mVenus expression; panels d, e, and insets in a, b, GFP expression. (d) Unilateral MultiColor FlpOut of DNp10, DNp07, and DNp13. (e) Saggital view of DNp10, DNp07, and DNp13. Note that DNp13 does not innervate the leg neuropil areas (dashed circles). For all panels, similar results were obtained from N = 4 different flies.

Supplementary Figure 2 Movements elicited by DNp07 and DNp10 activation resemble visually evoked landing responses.

(a) Video frames from a visually evoked landing response in a tethered fly (inset in c), captured at 1000 Hz. Stimulus was a unilateral bar moving front to back at 1000 °/s. Cyan line and circle in right panel indicates body axis orientation for c-e. (b) Still image of a free flying fly landing spontaneously on the ground, rotated to match the body orientation of tethered flies. (c) Example leg tip trajectories from four different flies viewing the same stimulus as in a. Leg tip positions for ipsilateral middle and hind legs and both front legs are shown every 5 ms for 200 frames. Dashed arrows (Fly 1) indicate the vector length used to quantify the leg extension amplitude in the graphs on right and main Fig. 2. Gray shading indicates visual stimulus. All legs moved in-phase with each other. Hind leg movements were more variable than front and middle leg movements. (d) Example leg tip trajectories elicited by DNp07-split-GAL4-1 activation (100% intensity), shown every 5 ms for 50 frames. Orange shading indicates optogenetic stimulus. (e) Same as d for DNp10 activation (75% intensity). Leg trajectories and body attitude were rotated so that all body axes are oriented horizontally.

Supplementary Figure 3 DNp07 and DNp10 drive landing responses with different middle and hind leg postures upon optogenetic activation.

(a) Optogenetic activation of DNp07 (DNp07-split-GAL4-2, magenta) or DNp10 (blue) extends front, middle, and hind legs, shown as leg tip position change from start to peak during a 300-ms light pulse. Light colors show fly means, dark arrows population average. BL, body length. (b) Boxplots show median and quartiles of leg extension angles during DNp07 and DNp10 activation relative to initial position. N = 6 flies, N = 12 legs, n = 4-5 trials each. Data from left legs was reflected and included. ***P = 5.9x10-4, *P = 0.0304 (two-sided Wilcoxon rank sum test). Note that the two outliers (black) in b have minute amplitudes in a.

Supplementary Figure 4 DNp07 and DNp10 silencing experiments.

(a) Tethered flies were shown visual stimuli, including a 500°/s fast, 10°-wide, dark, front-to back moving bar (the same stimulus we used for electrophysiological experiments), frontal looming (r/v = 80), and bilateral frontal expansion (500°/s). (b) Still video frames indicate ROI (orange circles) used for automatic detection of front leg extension during the landing response. (c) Example trace of normalized mean pixel intensity of the front leg ROI, smoothed over a 50-ms time window. Sequence shows leg movement responses (orange) to three subsequent visual stimuli (stimulus time courses in gray). ROI pixel intensity traces were used to calculate time points of landing responses by applying a constant threshold (3x SD) for each fly (dashed line). (d) Takeoff was not impaired in DNp07-split-GAL4-1 or DNp10-silenced flies. Bars, show mean + SD takeoff fraction; dots, individual fly means. NGFP = 9, NDNp07 = 11, NDNp10 = 10 different flies. Flight starts in this case were without ground contact. (e) Examples of ROI intensity traces from N = 3 flies each for three genotypes showing landing responses (black or colored) or no landing responses (gray) to n = 18 presentations of looming stimuli (below). For each genotype, an example of a fly with a low, intermediate, and high response rate is shown. Dotted lines indicate an approximation of the onset time of ROI intensity changes, when the leg moved into the ROI, for GFP controls. DNp07 examples are using the DNp07-split-GAL4-1 line.

Supplementary Figure 5 DNp07 and DNp10 have similar visual tuning. We visually stimulated the right eye while recording the ipsi- (right) or contralateral (left) copy of each DN.

(a-c) DN responses (mean ± SD, sample sizes defined in b; N, number of recordings from different flies) to different speeds of front-to-back motion of bars (a), small squares (b), or lateral looming stimuli presented at 45° azimuth, and 45° elevation (c) (which, unlike frontal looming stimuli, do not strongly drive landing responses). (d) Directional tuning of landing DNs, dots show the mean ± SD peak firing rate in 20-ms bins during stimulation with 100°/s fast moving bars. N = 6 ipsilateral, N = 4 contralateral recordings per DN. Inset shows directions of motion, u, up, d, down, b, back-to-front, f, front-to-back. (e) Both DNs responded more strongly to dark compared to bright objects. The contrast selectivity was calculated from stimulation with bright, b, and dark, d, edges moving front-to-back at 500°/s (example responses from DNp07). +1 indicates response to dark edges only, -1 bright edges only. N = 11 DNp10, N = 4 DNp07 recordings from different flies. Panels a-e all from flight trials only. (f) DNp07 recordings from a setup with no leg tracking, permitting bilateral visual stimulation with dark-edge front-to-back motion (insets), providing an expansion cue that mimics frontal object approach. Right, mean number of spikes from bilateral stimulation was comparable to the linear sum of unilateral responses. All responses in f from non-flight trials. N, number of recordings from different flies; n, number of trials per fly.

Supplementary Figure 6 Timing and magnitude of leg movements during flight were correlated with landing DN activity.

(a) DNp07 spike times, example intracellular recording, spike rate, and middle leg extension amplitude vector for five examples of visually-evoked landing responses (see Supplementary Figure 2c) aligned to the timing of the first spike. One non-spiking trial (lightest gray) was aligned to peak leg extension. (b) Same plot details as in a, data from DNp10. (c) Additional example time traces of the DNp10 spike rate (in 100-ms bins) and the middle leg extension vector during different visual stimuli. (d) A data set of 41 trials from the same fly during one DNp07 recording was separated and pooled into three groups according to peak leg extension amplitude: 0-30th percentile (left), 30-70th percentile (middle), and >70th percentile (right). Traces show mean leg extension vector (black) and spike rates (magenta) aligned to movement onset. (e) Plot details as in d, for 39 trials during one DNp10 recording.

Supplementary Figure 7 Optogenetic control of landing DN spike rate.

(a-b) Whole cell DNp07 and DNp10 recordings during 50-ms optogenetic stimulation (orange) at different light intensities, indicated as percent of LED driver power. Spike time rasters (black) are shown for five subsequent trials at each intensity. Inset on lower left shows the first five spikes for three subsequent trials of DNp07 activation. Automatic spike detection was accurate, even during large depolarizations resulting in high frequency spiking. (c-d) DN spike rate depended linearly on LED intensity. Gray lines, spike rate mean ± SD (shading) per fly; colored lines, linear fits to population data. N, number of recordings from different flies. (e) Light intensities at different LED power settings, measured at the position of the fly’s head.

Supplementary Figure 8 Dopamine does not mediate the flight-dependent increase of DNp10 visual responses.

(a) Mean ± SD spike rate of DNp10 during presentation of front-to-back moving bars of different velocities in non-flying flies. N, number of recordings from different flies. (b) Visual responses of the same flies to the same set of stimuli after application of 10 mM DA. N, number of recordings from different flies. (c) Example responses to 500°/s fast bar during non-flight (black) non-flight + DA (orange), and flight after DA application (gray). All trials from the same fly.

Supplementary Figure 9 Landing DNs are multimodal and respond to mechanosensory stimulation via air puffs.

(a). Spike rate (upper traces) and membrane potential of DNp10 during 1-s air puff stimulation (cyan bar). Different colors indicate different manipulations. Ablation of the ipsi- (-i. ant, dark orange) or contralateral (-c. ant, light orange) antenna halved the depolarization of DNp10, and ablation of both antennae eliminated spiking in DNp10. In contrast, DNp07 was inhibited by air puffs. Ablation of both antennae (no ant., black.) reduced the hyperpolarization of DNp07 strongly and additional ablation of the legs (no ant. and no legs, gray) nearly eliminated the mechanosensory response. All traces in a are population means. (b) Quantification of ablation experiments. Plots show trial means (small dots) and grand mean (large dots) of membrane potential calculated over a 900-ms window starting 100 ms after air puff onset for each condition. ***P = 3x10-9, t = 9.1, df = 24 (two-sided t-test). **P = 0.0017 , t = -3.6, df = 22 (two-sided t-test). Sample sizes given in a. All data from non-flying flies. Air puffs resulted in DNp10 excitation independent of whether or not they elicited a flight bout.

Supplementary Figure 10 Schematic model for state-dependent decoupling of sensory and motor networks.

In non-flying flies (left), visual responses of landing DNs (DNp07, magenta, and DNp10, blue) are gated out. The activity of sensory networks activated by visual stimuli such as looming cues can therefore not drive landing motor patterns in the VNC. Other descending pathways, like those driving takeoff (black), are not gated out and hence responsive. In this case, looming visual stimuli are channeled into the escape takeoff pathway in non-flight, and the fly takes off in response to a looming stimulus. In flight (right), landing DNs are gated in via octopaminergic neuromodulation (OA, orange) or by intrinsic feedback from flight motor circuits (FB, green), such that visual responses drive landing DNs and, as a result, landing motor patterns in the VNC. Though not directly measured here, descending neurons for escape may be gated out during flight, since in this state visual stimuli do not evoke escape-related behaviors, such as rapid middle leg extension (see Videos M1-2 and M8-10), and instead frontal visual looming stimuli activate the landing pathway during flight.

Supplementary Figure 11 Optogenetic excitability of landing DNs is similar in flight and non-flight.

(a) Optogenetic activation of DNp07 elicited similar responses independent of whether flies were flying or not. (b) Boxplots show median and quartiles of spike numbers per trial during flight and non-flight bouts for N = 3 flies and n trials per fly and condition; dashes, outliers. On average, about one additional spike was elicited during non-flight bouts in DNp07. (c-d) Same plot details as (a-b), data from DNp10, N = 4 flies and n trials per fly and condition. The same number of spikes was elicited during optogenetic DNp10 activation, independent of whether or not the fly was flying. Dotted line in c shows hyperpolarized non-flight resting potential. P-values are from two-sided Wilcoxon Signed Rank test, sample sizes (n, number of trials per condition) are indicated above.

Supplementary information

Supplementary Figures 1-11.

Reporting Summary

Supplementary Video 1

Example of DNp10 > GFP control fly responses to repeated presentation of frontal visual looming stimuli. Video was captured at 100 Hz. Each trial sequence is 1.5s long and includes the last second of looming expansion and 500 ms after the stimulus stopped moving (see also Supplementary Fig. 2d). Video is slowed-down 6x. All trials were scored as landing trials

41593_2019_413_MOESM4_ESM.mp4

Supplementary Video 2 Example of DNp07 > GFP control fly responses to repeated presentation of frontal visual looming stimuli. Video was captured at 100 Hz. Each trial sequence is 1.5s long and includes the last second of looming expansion and 500 ms after the stimulus stopped moving (see also Supplementary Fig. 2d). Video is slowed-down 6x. All trials were scored as landing trials

41593_2019_413_MOESM5_ESM.mp4

Supplementary Video 3 Spontaneous landing approach in free flight. Flies landing spontaneously extend their leg to the typical landing pose. Video was captured at 600 Hz and slowed down 200x

41593_2019_413_MOESM6_ESM.mp4

Supplementary Video 4 Optogenetic activation of DNp07 reliably drives landing responses. Montage shows 20 trials of the same fly responding to 300-ms long activation of DNp07 (DNp07-split-GAL4-2 > UAS-CsChrimson) at 100% intensity. Video was captured at 125 Hz and slowed down to 1/5 real time. White square indicates optogenetic stimulation LED on

41593_2019_413_MOESM7_ESM.mp4

Supplementary Video 5 Optogenetic activation of DNp10 reliably drives landing responses. Montage shows 20 trials of the same fly responding to 300-ms long activation of DNp10 (DNp10-split-GAL4 > UAS-CsChrimson) at 100% intensity. Video was captured at 125 Hz and slowed down to 1/5 real time. White square indicates optogenetic stimulation LED on

41593_2019_413_MOESM8_ESM.avi

Supplementary Video 6 DNp07 activation drives landing responses in flight and similar leg movements in non-flight bouts. Montage shows six trials of the same fly responding to DNp07 activation at 100% intensity. The time stamp shows time relative to stimulus onset, the white dot in the top right indicates timing of Chrimson LED. The top row shows flight trials; the bottom row shows non-flight trials. The frame rate was 1, 000 Hz. Similar videos were acquired for N = 4 flies. Note that DN activation initiated flight in some trials, which had much longer latencies than leg movements. Trials with ground contact resulted in similar leg movements without flight starts (data not shown). Genotype used was DNp07-split-GAL4-1 > UAS-CsChrimson

41593_2019_413_MOESM9_ESM.avi

Supplementary Video 7 DNp10 activation drives landing responses in flight and similar leg movements in non-flight bouts. Montage of six trials of the same fly responding to DNp10 activation at 100% intensity. Details as for Supplementary Video 6. Genotype used was DNp10-split-GAL4 > UAS-CsChrimson

41593_2019_413_MOESM10_ESM.mp4

Supplementary Video 8 Example DNp07-silenced fly responses to repeated presentation of frontal visual looming stimuli. Video was captured at 100 Hz. Each trial sequence is 1.5s long and includes the last second of looming expansion and 500 ms after the stimulus stopped moving (see also Supplementary Fig. 2d). Video is slowed-down 6x. None of the trials was scored as a landing trial. Genotype used was DNp07-split-GAL4-1 > UAS-Kir2.1

41593_2019_413_MOESM11_ESM.mp4

Supplementary Video 9 Example DNp10-silenced fly responses to repeated presentation of frontal visual looming stimuli. Video was captured at 100 Hz. Each trial sequence is 1.5s long and includes the last second of looming expansion and 500 ms after the stimulus stopped moving (see also Supplementary Fig. 2d). Video is slowed-down 6x. Trial 10 was scored as a landing trial. Several other trials show partial leg extensions. Genotype used was DNp10-split-GAL4 > UAS-Kir2.1

41593_2019_413_MOESM12_ESM.mp4

Supplementary Video 10 Example DNp10-silenced fly responses to repeated presentation of frontal visual looming stimuli. Video was captured at 100 Hz. Each trial sequence is 1.5 s long and includes the last second of looming expansion and 500 ms after the stimulus stopped moving (see also Supplementary Fig. 2d). Video is slowed-down 6x. Trials 1, 3, 4, 5, 9 and 10 were scored as landing trials. Several other trials show partial leg extensions. Genotype used was DNp10-split-GAL4 > UAS-Kir2.1

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ache, J.M., Namiki, S., Lee, A. et al. State-dependent decoupling of sensory and motor circuits underlies behavioral flexibility in Drosophila. Nat Neurosci 22, 1132–1139 (2019). https://doi.org/10.1038/s41593-019-0413-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-019-0413-4

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