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:

A faithful internal representation of walking movements in the Drosophila visual system

Abstract

The integration of sensorimotor signals to internally estimate self-movement is critical for spatial perception and motor control. However, which neural circuits accurately track body motion and how these circuits control movement remain unknown. We found that a population of Drosophila neurons that were sensitive to visual flow patterns typically generated during locomotion, the horizontal system (HS) cells, encoded unambiguous quantitative information about the fly's walking behavior independently of vision. Angular and translational velocity signals were integrated with a behavioral-state signal and generated direction-selective and speed-sensitive graded changes in the membrane potential of these non-spiking cells. The nonvisual direction selectivity of HS cells cooperated with their visual selectivity only when the visual input matched that expected from the fly's movements, thereby revealing a circuit for internally monitoring voluntary walking. Furthermore, given that HS cells promoted leg-based turning, the activity of these cells could be used to control forward walking.

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: Leg movement modulates the membrane potential dynamics of LPTCs.
Figure 2: A behavioral-state signal in HS cells precedes the onset of, and continues after, the offset of behavior.
Figure 3: The dynamics of HS cells during walking bouts in darkness correlates with the angular and forward velocities of the fly.
Figure 4: The behavioral-state signal arrives to HS cells through upstream visual circuits.
Figure 5: HS cells are tuned by a weighted sum of three movement-related components.
Figure 6: HS cells encode positive combinations of visual and angular velocities.
Figure 7: HS cells control turning in the walking fly.

Similar content being viewed by others

References

  1. Franklin, D.W. & Wolpert, D.M. Computational mechanisms of sensorimotor control. Neuron 72, 425–442 (2011).

    Article  CAS  PubMed  Google Scholar 

  2. Whitlock, J.R., Sutherland, R.J., Witter, M.P., Moser, M.-B. & Moser, E.I. Navigating from hippocampus to parietal cortex. Proc. Natl. Acad. Sci. USA 105, 14755–14762 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Lappe, M., Bremmer, F. & van den Berg, A.V. Perception of self-motion from visual flow. Trends Cogn. Sci. 3, 329–336 (1999).

    Article  CAS  PubMed  Google Scholar 

  4. Warren, W.H. Jr., Kay, B.A., Zosh, W.D., Duchon, A.P. & Sahuc, S. Optic flow is used to control human walking. Nat. Neurosci. 4, 213–216 (2001).

    Article  CAS  PubMed  Google Scholar 

  5. Pfeffer, S.E. & Wittlinger, M. Optic flow odometry operates independently of stride integration in carried ants. Science 353, 1155–1157 (2016).

    Article  CAS  PubMed  Google Scholar 

  6. Morgan, B. & Frost, B.J. Visual response characteristics of neurons in nucleus of basal optic root of pigeons. Exp. Brain Res. 42, 181–188 (1981).

    Article  CAS  PubMed  Google Scholar 

  7. Britto, L.R., Natal, C.L. & Marcondes, A.M. The accessory optic system in pigeons: receptive field properties of identified neurons. Brain Res. 206, 149–154 (1981).

    Article  CAS  PubMed  Google Scholar 

  8. Grasse, K.L. & Cynader, M.S. Electrophysiology of medial terminal nucleus of accessory optic system in the cat. J. Neurophysiol. 48, 490–504 (1982).

    Article  CAS  PubMed  Google Scholar 

  9. Hausen, K. Motion sensitive interneurons in the optomotor system of the fly. Biol. Cybern. 45, 143–156 (1982).

    Article  Google Scholar 

  10. Kano, M.-S., Kano, M. & Maekawa, K. Receptive field organization of climbing fiber afferents responding to optokinetic stimulation in the cerebellar nodulus and flocculus of the pigmented rabbit. Exp. Brain Res. 82, 499–512 (1990).

    Article  CAS  PubMed  Google Scholar 

  11. Duffy, C.J. & Wurtz, R.H. Sensitivity of MST neurons to optic flow stimuli. I. A continuum of response selectivity to large-field stimuli. J. Neurophysiol. 65, 1329–1345 (1991).

    Article  CAS  PubMed  Google Scholar 

  12. Kubo, F. et al. Functional architecture of an optic flow-responsive area that drives horizontal eye movements in zebrafish. Neuron 81, 1344–1359 (2014).

    Article  CAS  PubMed  Google Scholar 

  13. Schnell, B. et al. Processing of horizontal optic flow in three visual interneurons of the Drosophila brain. J. Neurophysiol. 103, 1646–1657 (2010).

    Article  CAS  PubMed  Google Scholar 

  14. Koenderink, J.J. Optic flow. Vision Res. 26, 161–179 (1986).

    Article  CAS  PubMed  Google Scholar 

  15. Bradley, D.C., Maxwell, M., Andersen, R.A., Banks, M.S. & Shenoy, K.V. Mechanisms of heading perception in primate visual cortex. Science 273, 1544–1547 (1996).

    Article  CAS  PubMed  Google Scholar 

  16. Gu, Y., Angelaki, D.E. & Deangelis, G.C. Neural correlates of multisensory cue integration in macaque MSTd. Nat. Neurosci. 11, 1201–1210 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Britten, K.H. & van Wezel, R.J. Electrical microstimulation of cortical area MST biases heading perception in monkeys. Nat. Neurosci. 1, 59–63 (1998).

    Article  CAS  PubMed  Google Scholar 

  18. Saleem, A.B., Ayaz, A., Jeffery, K.J., Harris, K.D. & Carandini, M. Integration of visual motion and locomotion in mouse visual cortex. Nat. Neurosci. 16, 1864–1869 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Keller, G.B., Bonhoeffer, T. & Hübener, M. Sensorimotor mismatch signals in primary visual cortex of the behaving mouse. Neuron 74, 809–815 (2012).

    Article  CAS  PubMed  Google Scholar 

  20. Roth, M.M. et al. Thalamic nuclei convey diverse contextual information to layer 1 of visual cortex. Nat. Neurosci. 19, 299–307 (2016).

    Article  CAS  PubMed  Google Scholar 

  21. Borst, A., Haag, J. & Reiff, D.F. Fly motion vision. Annu. Rev. Neurosci. 33, 49–70 (2010).

    Article  CAS  PubMed  Google Scholar 

  22. Scott, E.K., Raabe, T. & Luo, L. Structure of the vertical and horizontal system neurons of the lobula plate in Drosophila. J. Comp. Neurol. 454, 470–481 (2002).

    Article  PubMed  Google Scholar 

  23. Haikala, V., Joesch, M., Borst, A. & Mauss, A.S. Optogenetic control of fly optomotor responses. J. Neurosci. 33, 13927–13934 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Kim, A.J., Fitzgerald, J.K. & Maimon, G. Cellular evidence for efference copy in Drosophila visuomotor processing. Nat. Neurosci. 18, 1247–1255 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. 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  PubMed  PubMed Central  Google Scholar 

  26. Greenspan, R.J. & Ferveur, J.F. Courtship in Drosophila. Annu. Rev. Genet. 34, 205–232 (2000).

    Article  CAS  PubMed  Google Scholar 

  27. Bloomquist, B.T. et al. Isolation of a putative phospholipase C gene of Drosophila, norpA, and its role in phototransduction. Cell 54, 723–733 (1988).

    Article  CAS  PubMed  Google Scholar 

  28. Kabra, M., Robie, A.A., Rivera-Alba, M., Branson, S. & Branson, K. JAABA: interactive machine learning for automatic annotation of animal behavior. Nat. Methods 10, 64–67 (2013).

    Article  CAS  PubMed  Google Scholar 

  29. Kamikouchi, A. et al. The neural basis of Drosophila gravity-sensing and hearing. Nature 458, 165–171 (2009).

    Article  CAS  PubMed  Google Scholar 

  30. Fraenkel, G. & Pringle, J.W.S. Halteres of flies as gyroscopic organs of equilibrium. Nature 141, 919–920 (1938).

    Article  Google Scholar 

  31. Mendes, C.S., Bartos, I., Akay, T., Márka, S. & Mann, R.S. Quantification of gait parameters in freely walking wild type and sensory deprived Drosophila melanogaster. eLife 2, e00231 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  32. Tuthill, J.C. & Wilson, R.I. Parallel transformation of tactile signals in central circuits of Drosophila. Cell 164, 1046–1059 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Preuss, T. & Hengstenberg, R. Structure and kinematics of the prosternal organs and their influence on head position in the blowfly, Calliphora erythrocephala. J. Comp. Physiol. A. 171, 483–493 (1992).

    Article  Google Scholar 

  34. Schilstra, C. & Hateren, J.H. Blowfly flight and optic flow. I. Thorax kinematics and flight dynamics. J. Exp. Biol. 202, 1481–1490 (1999).

    Article  PubMed  Google Scholar 

  35. Blaj, G. & van Hateren, J.H. Saccadic head and thorax movements in freely walking blowflies. J. Comp. Physiol. A Neuroethol. Sens. Neural Behav. Physiol. 190, 861–868 (2004).

    Article  CAS  PubMed  Google Scholar 

  36. Schnell, B., Raghu, S.V., Nern, A. & Borst, A. Columnar cells necessary for motion responses of wide-field visual interneurons in Drosophila. J. Comp. Physiol. A Neuroethol. Sens. Neural Behav. Physiol. 198, 389–395 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  37. Haag, J. & Borst, A. Active membrane properties and signal encoding in graded potential neurons. J. Neurosci. 18, 7972–7986 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Petreanu, L. et al. Activity in motor-sensory projections reveals distributed coding in somatosensation. Nature 489, 299–303 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Pabst, M. & Kral, K. Proprioceptive contribution to distance estimation by motion parallax in a praying mantid. J. Exp. Biol. 201, 1483–1491 (1998).

    Article  PubMed  Google Scholar 

  40. Mischiati, M. et al. Internal models direct dragonfly interception steering. Nature 517, 333–338 (2015).

    Article  CAS  PubMed  Google Scholar 

  41. Kress, D. & Egelhaaf, M. Gaze characteristics of freely walking blowflies Calliphora vicina in a goal-directed task. J. Exp. Biol. 217, 3209–3220 (2014).

    PubMed  Google Scholar 

  42. Schnell, B., Weir, P.T., Roth, E., Fairhall, A.L. & Dickinson, M.H. Cellular mechanisms for integral feedback in visually guided behavior. Proc. Natl. Acad. Sci. USA 111, 5700–5705 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Huston, S.J. & Krapp, H.G. Visuomotor transformation in the fly gaze stabilization system. PLoS Biol. 6, e173 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  45. 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  PubMed  Google Scholar 

  46. 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  PubMed  Google Scholar 

  47. Muijres, F.T., Elzinga, M.J., Iwasaki, N.A. & Dickinson, M.H. Body saccades of Drosophila consist of stereotyped banked turns. J. Exp. Biol. 218, 864–875 (2015).

    Article  PubMed  Google Scholar 

  48. Schuster, S., Strauss, R. & Götz, K.G. Virtual-reality techniques resolve the visual cues used by fruit flies to evaluate object distances. Curr. Biol. 12, 1591–1594 (2002).

    Article  CAS  PubMed  Google Scholar 

  49. Pelah, A. & Barlow, H.B. Visual illusion from running. Nature 381, 282–283 (1996).

    Article  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  51. Hughes, C.L. & Thomas, J.B. A sensory feedback circuit coordinates muscle activity in Drosophila. Mol. Cell. Neurosci. 35, 383–396 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Giorgianni, M.W. & Mann, R.S. Establishment of medial fates along the proximodistal axis of the Drosophila leg through direct activation of dachshund by Distalless. Dev. Cell 20, 455–468 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Stockinger, P., Kvitsiani, D., Rotkopf, S., Tirián, L. & Dickson, B.J. Neural circuitry that governs Drosophila male courtship behavior. Cell 121, 795–807 (2005).

    Article  CAS  PubMed  Google Scholar 

  54. 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 

  55. Seelig, J.D. et al. Two-photon calcium imaging from head-fixed Drosophila during optomotor walking behavior. Nat. Methods 7, 535–540 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Wilson, R.I., Turner, G.C. & Laurent, G. Transformation of olfactory representations in the Drosophila antennal lobe. Science 303, 366–370 (2004).

    Article  CAS  PubMed  Google Scholar 

  57. Suter, B.A. et al. Ephus: multipurpose data acquisition software for neuroscience experiments. Front. Neural Circuits 4, 100 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  Google Scholar 

  59. Martin, J.R. A portrait of locomotor behavior in Drosophila determined by a video-tracking paradigm. Behav. Process. 67, 207–219 (2004).

    Article  Google Scholar 

  60. Wosnitza, A., Bockemühl, T., Dübbert, M., Scholz, H. & Büschges, A. Inter-leg coordination in the control of walking speed in Drosophila. J. Exp. Biol. 216, 480–491 (2013).

    PubMed  Google Scholar 

  61. Kabsch, W. A solution for the best rotation to relate two sets of vectors. Acta Crystallogr. 32, 922–923 (1976).

    Article  Google Scholar 

  62. Breiman, L. Random forests. Machine Learning 45, 5–32 (2001).

    Article  Google Scholar 

Download references

Acknowledgements

We thank the Champalimaud Research fly facility for stock construction and maintenance, J. Liu (Janelia Research Campus) for technical support, V. Atahlye and P. Vertechi (Champalimaud Research Programme) for analytical advice, and G. Card (Janelia Research Campus), L. Luo (Stanford University), G. Miesenböck (Centre for Neural Circuits and Behaviour) and C. Mendes (CEDOC) for sharing flies. We thank N. Rito for assisting with the freely walking flies experiments. We are grateful to S. Huston, R. Wilson, R. Costa and C. Machens for early discussions about our results. We thank L. Petreanu, M. Carey, V. Jayaraman and members of the Chiappe laboratory for useful comments on the manuscript. This work was supported by the Champalimaud Foundation, by a Bial Foundation grant 191/12 (M.E.Ch), and by a Marie Curie Career Integration grant PCIG13-GA-2013-618854 (M.E.Ch).

Author information

Authors and Affiliations

Authors

Contributions

T.F. and M.E.C. designed the study. T.F. performed the electrophysiological experiments. T.L.C. designed the analysis of behavior and constructed the model. J.P.B. performed the tethered walking behaviors. All of the authors analyzed the data, and T.F., T.L.C. and M.E.C. wrote the manuscript.

Corresponding author

Correspondence to M Eugenia Chiappe.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Visual responses of the recorded LPTCs

(a) The mean±SD (n=3 trials) of the direction-selective (DS) responses (i.e., subtracting the null direction, ND, from the preferred direction, PD) of the HSE cell shown in Fig. 1c, to a wide-field horizontal (left) or vertical (right) moving grating. The light gray background indicates the stimulus period. (b) The (time) average DS response amplitude during stimulation (grey dots), and the mean value (black lines). (c) Triple immunostaining for the GFP signal (green) driven by the R81G07Gal4 line, for the biocytin signal (red) from the recorded HS cell, and for the bruchpilot protein (brp, blue), staining for synaptic neuropile. The dash line delineates the border of the lobula plate. (d–f) Same as (a–c), but for the VS1 cell shown in Fig. 1d. (g–i) Same as (a–c), but for the HSE cell in a norpA mutant background (blind fly) shown in Fig. 1e. (j) DS visual responses from HS cells in T4/T5-Gal4 (left) and UAS-Kir (middle) control flies, or from HS cells in flies with T4/T5 cells silenced by the expression of Kir. Shown is the mean±SEM (N is indicated in each corresponding figure).

Supplementary Figure 2 Classification of tethered behavior on the ball

(a) Swing-stance periods (black or white, respectively) during a walking bout as defined by videography. Note that at steady forward walking (from 0.5 to 4 s), the fly displays tripod gate (see zoom-in, N=6 flies). (b) Examples of a subset of the first ten principal components (PCs) of the side-view fly video recordings (Fig. 1a,b). Asterisks indicate the PCs whose scores are shown in (c). (c) Time-varying scores for different PCs. A zoom-in into PC5’s scores reveals a periodic behavior of a principal component associated with leg movement. (d) A wavelet transformation of PC5 scores further indicates the bandwidth of the signal, matching the mean cycle period obtained from the swing-phase analysis. (e) The PCs scores and their wavelet transform (c,d) were used to train JAABA (top schematic) to classify walking and other possible behaviors of the fly on the ball (bottom color-coded image). (f) Example of a 1-minute trial displaying the fly’s forward (grey top trace), and angular (grey bottom trace) velocity signals during walking (pink background) and non-walking (blue background) segments. For angular velocity, CCW turns are positive sign values. For forward velocity, negative sign indicates backward walking. Arrowheads show activity segments (i.e., non-walking but non-stationary segments).

Supplementary Figure 3 Walking-specific responses during locomotion in HS cells

Cross-covariance analysis per cell between the fly’s angular velocity (or forward velocity) and the Vm of HS cells, in walking or activity bouts. The associated distributions indicate the % of activity time spent in walking (“walking ratio”). Note that the example cells included in Figure 3 are not presented here.

Supplementary Figure 4 Walking direction and speed sensitivity in HS cells

(a) Change in Vm of left-side HS cells with respect to quiescence (ΔVm, color-coded) as a function of the forward and angular velocities of each individual fly. (b) Same as in (a), but for right-side HS cells. Scale bar is the same as for the maps in (a). (c) Population ΔVm (color-coded) as a function of the forward and angular acceleration of the fly. (d) Examples of the Vm of HS cells (orange), the forward (Vf, black), and angular velocities (Va, gray) of the fly in segments of walking bouts where the fly’s angular velocity did not exceed ±35°/s (gray shadow). (e) Left, weighted grand mean±SEM of the average angular velocity of the fly for the selected segments with low angular velocity as a function of the averaged Vf of the fly. The weight is given as a function of number of segments where the fly showed |Va| < 35°/s (mean = 240, range: 50-499 bouts). Right, weighted grand mean±SEM of the average Vm as a function of the average Vf of the fly. (f) Mean correlation coefficient between the fly’s Vf and the Vm of HS cells per cell, and between Vf and Va per cell. Significant correlations are red, non-significant correlations are blue (traces were shuffled 20 times by concatenated walking bouts, Wilcoxon’s signed-rank test).

Supplementary Figure 5 Movement-related modulations in VS cells

(a) Vm of a VS1 cell triggered at the onset of postural adjustments in walking (left), or activity (right) bouts. n = number of bouts. Lighter traces show triggered data from individual bouts. (b) Left, the mean Vm of the population of VS cells (mean±SEM) triggered at the onset of walking (black) or activity (gray) bouts. Right, the mean Vm triggered at the offset of walking (black) or activity (gray) bouts. n = total number of bouts. For walking onset bouts, N=10 cells. For activity onset analysis, N=6 cells. For walking offset analysis, N=7 cells. For activity offset analysis, N=4 cells. (c) Distribution of the 10% rise time of the Vm (see Methods) relative to the onset of walking bouts. Arrowhead indicates the median value. (d) ΔVm (color-coded) as a function of the forward and angular velocities of each individual fly. (e) ΔVm (color-coded) is plotted as a function of the forward and angular velocities of the fly for the population of recorded VS cells, right and left cells were pooled together (see Methods). Scale bar is the same as for the maps in (d).

Supplementary Figure 6 Walking-specific signals are independent of vision, are not coupled to the movement of antennae or halteres, and do not depend on the synaptic activity of leg mechanosensory cells.

(a) Walking velocity tuning for the recorded HS cell shown in Fig. 1e. PDmotor and NDmotor are the cell’s preferred and null turning direction of the fly (see main text). (b) Walking velocity tuning map for the population of HS cells recorded from blind (norpA) flies (N=9 cells). (c) Walking velocity tuning maps for the population of HS cells recorded from flies with antennae and halteres waxed (N=11 cells, right and left-side HS cells were pooled, see Methods). (d) Left, walking velocity tuning maps for the population of HS cells recorded from flies with leg mechanosensory cells expressing tetanus toxin to disrupt their synaptic activity (N=8 cells). Right, walking velocity maps of parental control flies with functional synaptic activity in leg mechanosensory cells (N=7 cells). (e) Top, paths of freely walking experimental (red, flies with leg mechanosensory cells expressing tetanus toxin, TNT) or control flies (black). Bottom, zoomed-in view to highlight the straightness of the paths of the walking flies. (f) Probability distributions (mean±SD) of the straightness of walking paths in experimental (red) and control (black) flies (see Methods). (g) Mean path straightness for experimental (red) and control (black) flies. For straightness analysis: 13 control flies, mean±SD: 0.87±0.03; 15 experimental flies, mean±SD: 0.71±0.07, ***,P< 0.0001, Z=4.38, Wilcoxon’s rank-sum test. (h) Probability distributions (mean±SD) of walking speed in experimental (red) and control (black) flies. For walking velocity analysis: 13 control flies, mean±SD: 11.5±1.9 mm/s; 15 experimental flies, mean±SD: 9.6±1.5 mm/s, P<0.01, Z=2.63, Wilcoxon’s rank-sum test.

Supplementary Figure 7 Correlation between head and body movements during tethered walking in darkness

(a) Top: Schematic of head-tracking set-up. Middle: example frame from a top-view camera. Bottom: example frame from a side-view camera. Contrast has been enhanced for clarity. (b) Example traces from one trial comparing the head yaw movement and the body’s turn. Top, angular velocity of the fly. Middle, head yaw angle. Bottom, head yaw velocity. Arrows indicate head-yaw angle offsets. (c) Cross-covariance between angular velocity and head angle for all walking bouts in darkness for a single fly (n=390 bouts from 126 trials). Thick black trace, mean; shadow, SD. The magenta line in all cross-covariance plots indicates the bootstrapped 95% confidence level of the analysis (see Methods). Light gray trace, example from (b). (d) Grand mean cross-covariance. The head precedes the body by 50 ms on average (mean±SEM =48.3±24.0 ms, N=9 flies). (e) Instantaneous head yaw position as a function of the angular and the forward velocities of the fly. (f) Idem as in (c) but the head yaw velocity was compared with the angular velocity of the fly. (g) Same as in (d) but with the head yaw velocity. (h) Instantaneous head yaw velocity as a function of the forward and angular velocities of the fly. (i) Cross-covariance analysis between the head pitch angle and the forward velocity of the fly. Grand mean±SEM (341±143ms, N=6 flies).

Supplementary Figure 8 Model performance analysis

(a) Predicted velocity tuning maps for different single-component models. (b) Mean cross-correlation coefficients per cell between observed and predicted Vm for different models (black, matched pairs; red, mismatched pairs; see Methods, lines: mean values, **: P<0.001, Z>3.78, N=19 cells, Wilcoxon signed-rank test). Right-most column: cross-correlation coefficients between the observed and the predicted yaw head angle. (c) Performance of the BS and the Va+Vf+BS models in each cell. Indicated are the example cells shown in Fig. 3a, and in (e). (d) Performance of the Va+Vf+BS and the Va+BS models for each cell. Mean (±SD) correlation coefficients between the predicted and observed HS dynamics (see Methods). Red and blue: the difference in the magnitude of correlation coefficients between the two models cannot be explained by the reshuffling procedure (traces were shuffled 20 times by concatenated walking bouts, Wilcoxon’s signed-rank test, P<0.005), whereas the gray pairs can (P >0.02, Wilcoxon‘s signed-rank test). (e) Example of a cell with the lowest three-compartment model performance. Note that the HS cells’ dynamics are still well described by the three-component model. (f) Top, the predicted velocity-tuning map for the example cell in Figure 5c scaled in mm/s units in both axes. θ: the angle between the forward velocity axis and the membrane potential change (ΔVm) gradient (see Methods). Bottom, distribution of θ for each cell; the mean value is indicated in black. The θ value for the observed population velocity map (N=19 cells) is indicated in red. (g) The observed (black) and predicted (magenta) head yaw angles, estimated from the three-component model fitting the head yaw angle instead of the Vm. (h) Power spectrum analysis of the observed, the predicted, and the difference between the two for Vm (left), or yaw head angle (right). Note that the largest difference for the head angle prediction is on the DC component, i.e., the offset of the head position (arrows in (g)).

Supplementary Figure 9 Decoding walking velocities from the bilateral activity of HS cells

(a) Recordings from HS cells on one side of the brain were combined with predicted HS dynamics of the other right side using the walking behavior of the fly as input for the BS+Va+Vf model (Figure 5). The dynamics of the bilateral HS cells are labeled as modeled + recorded. (b) The mean angular (left) or forward (middle) velocities of the fly and normalized data points (right) were plotted as a function of the negative (VmLeft-VmRight) or positive (VmLeft+VmRight) combination of the bilateral dynamics of modeled + recorded cells. (c) Using the bilateral modeled + recorded dynamics, or the unilateral recorded cell dynamics, we applied a linear decoder (see Methods) to predict the angular velocity (Va) of the fly. Left, top two plots, example of decoding using bilateral HS cells’ dynamics. Bottom two plots, examples of decoding using the recorded cell dynamics only, i.e., the unilateral HS cells’ activity. Red trace, estimated; black trace observed. Right, covariance coefficients (Cov Coef) between the predicted and observed velocities using bilateral or unilateral HS cells’ dynamics. Walking bouts and the cell dynamics were reshuffled 20 times to obtain a population of Cov Coefs per fly and the mean value was used. Red, Cov Coefs between matched behavior and Vm (bilateral, mean±SD=0.85±0.06; unilateral, mean±SD=0.48±0.14,); gray, Cov Coefs between mismatched behavior and Vm. The difference in the magnitude of Cov Coefs between bilateral and unilateral decoding is significant (P<0.0005, Z=3.82, Wilcoxon signed-rank test, N=19 cells). (d) Decoding of the forward velocity (Vf) of the fly. Same as in (c) Cov Coefs between matched behavior and Vm for bilateral model: mean±SD=0.50±0.13; for the unilateral model: mean±SD=0.29±0.12. The difference in the magnitude of Cov Coefs between bilateral and unilateral decoding is significant (Wilcoxon’s signed-rank test P< 0.005, N=19 cells). (e) Transfer functions for the decoding of angular velocity (Va). Gray, individual fly’s filters, black, mean filter. Note that the magnitude of the mean filters obtained with the recorded data is similar to those obtained with the modeled data. (f) Same for the decoding of forward velocity (Vf).

Supplementary Figure 10 Visuomotor interactions in HS cells

(a) Top: Velocity-tuning maps under replay visual stimulation per cell. Bottom, visual and angular velocity maps per cell. (b) Walking velocity tuning map across the population of right-side HS cells under replay trials (N=13). (c) Estimate of the visual modulation of HS cells’ activity under replay conditions in quiescent segments: the visual stimulus is scaled by the velocity tuning curve of HS cells (see Methods) to obtain an effective visual stimulus (red trace). This effective stimulus is convolved with a response kernel (see Methods) to estimate the visual-induced activity in HS cells (blue trace). For comparison, the observed visual responses under identical conditions is shown (black trace). The delay of the kernel was obtained by cross-correlation analysis between HS cell responses and visual stimuli in quiescent segments. (d) Distribution of θ (Fig. 6) for each cell for fits with R2>0.7 (top, 9/13 cells, mean±SD=37±9°). Black: mean value; red, θ for the population map shown in Figure 6b. (e) Predictions from ideal random forest decoders (see Methods) of the sum of the visual velocity (Vv) and the fly‘s angular velocity (Va, red traces), or the difference between the two (blue traces). The input signals for the decoder were either the estimated Vm dynamics from the visual stimulus and the three-component walking of the fly (top row), or the estimated Vm dynamics from Vv and the fly’s Va (bottom). c=correlation coefficient between the predicted and the observed values of the sum or difference between Vv and Va. (f) Cross-correlation coefficients between the predicted and the observed velocities for each recorded cell, for the two different inputs to the ideal decoders. Bars: the mean values across cells. Color code as in (e).

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–10 (PDF 2528 kb)

Supplementary Methods Checklist (PDF 434 kb)

Simultaneous recordings of physiology and behavior during an isolated walking bout.

Left, side view of a walking fly in darkness. Right top: the angular velocity (“rotational speed”, blue), and forward velocity (“forward speed”, red) of the walking fly. Right bottom, the change of the membrane potential of an HS cell relative to quiescence during ongoing walking. (MP4 2155 kb)

Detection of postural adjustments before the onset of tethered behavior.

Left, original side view of a fly. Right, thresholded pixel-change side view of the fly. Movie is slow-down 5X. Red square indicates the onset of postural adjustments associated with the start of walking (see Methods). (MP4 394 kb)

Head movement during tethered walking in darkness.

Top left: top view of the head to detect head yaw movements. Bottom left: the head yaw angle (magenta) and angular velocity (red) of the walking fly. Top right: side view of the fly to detect head pitch movements. Bottom right: the head pitch position (green) and forward velocity (blue) of the walking fly. (MP4 3690 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fujiwara, T., Cruz, T., Bohnslav, J. et al. A faithful internal representation of walking movements in the Drosophila visual system. Nat Neurosci 20, 72–81 (2017). https://doi.org/10.1038/nn.4435

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.4435

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