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:

Dynamic nonlinearities enable direction opponency in Drosophila elementary motion detectors

Abstract

Direction-selective neurons respond to visual motion in a preferred direction. They are direction-opponent if they are also inhibited by motion in the opposite direction. In flies and vertebrates, direction opponency has been observed in second-order direction-selective neurons, which achieve this opponency by subtracting signals from first-order direction-selective cells with opposite directional tunings. Here, we report direction opponency in Drosophila that emerges in first-order direction-selective neurons, the elementary motion detectors T4 and T5. This opponency persists when synaptic output from these cells is blocked, suggesting that it arises from feedforward, not feedback, computations. These observations exclude a broad class of linear-nonlinear models that have been proposed to describe direction-selective computations. However, they are consistent with models that include dynamic nonlinearities. Simulations of opponent models suggest that direction opponency in first-order motion detectors improves motion discriminability by suppressing noise generated by the local structure of natural scenes.

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: Measurements of opponency and LN models for motion detection.
Fig. 2: The first-order DS neurons T4 and T5 exhibit opponent responses.
Fig. 3: Opponency persists under changes of contrast and stimulus type.
Fig. 4: Temporal tuning of opponent suppression.
Fig. 5: T4 and T5 cells with silenced synapses continue to show opponency.
Fig. 6: Feedforward models can produce DO responses.
Fig. 7: Opponency improves velocity discriminability in EMDs.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Code availability

Code for all modeling is available at https://github.com/ClarkLabCode/OpponencyModels.

References

  1. Snowden, R. J., Treue, S., Erickson, R. G. & Andersen, R. A. The response of area MT and V1 neurons to transparent motion. J. Neurosci. 11, 2768–2785 (1991).

    Article  CAS  Google Scholar 

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

    Article  Google Scholar 

  3. Heeger, D. J., Boynton, G. M., Demb, J. B., Seidemann, E. & Newsome, W. T. Motion opponency in visual cortex. J. Neurosci. 19, 7162–7174 (1999).

    Article  CAS  Google Scholar 

  4. Qian, N. & Andersen, R. A. Transparent motion perception as detection of unbalanced motion signals. II. Physiology. J. Neurosci. 14, 7367–7380 (1994).

    Article  CAS  Google Scholar 

  5. Joesch, M., Plett, J., Borst, A. & Reiff, D. Response properties of motion-sensitive visual interneurons in the lobula plate of Drosophila melanogaster. Curr. Biol. 18, 368–374 (2008).

    Article  CAS  Google Scholar 

  6. Mauss, A. S. et al. Neural circuit to integrate opposing motions in the visual field. Cell 162, 351–362 (2015).

    Article  CAS  Google Scholar 

  7. Maisak, M. S. et al. A directional tuning map of Drosophila elementary motion detectors. Nature 500, 212–216 (2013).

    Article  CAS  Google Scholar 

  8. Krapp, H. G. & Hengstenberg, R. Estimation of self-motion by optic flow processing in single visual interneurons. Nature 384, 463–466 (1996).

    Article  CAS  Google Scholar 

  9. Hassenstein, B. & Reichardt, W. Systemtheoretische Analyse der Zeit-, Reihenfolgen-und Vorzeichenauswertung bei der Bewegungsperzeption des Rüsselkäfers Chlorophanus. Z . Nat. 11, 513–524 (1956).

    Google Scholar 

  10. Adelson, E. & Bergen, J. Spatiotemporal energy models for the perception of motion. J. Opt. Soc. Am. A 2, 284–299 (1985).

    Article  CAS  Google Scholar 

  11. Gruntman, E., Romani, S. & Reiser, M. B. Simple integration of fast excitation and offset, delayed inhibition computes directional selectivity in Drosophila. Nat. Neurosci. 21, 250–257 (2018).

    Article  CAS  Google Scholar 

  12. Wienecke, C. F., Leong, J. C. & Clandinin, T. R. Linear summation underlies direction selectivity in drosophila. Neuron 99, 680–688.e4 (2018).

    Article  CAS  Google Scholar 

  13. Leong, J. C. S., Esch, J. J., Poole, B., Ganguli, S. & Clandinin, T. R. Direction selectivity in Drosophila emerges from preferred-direction enhancement and null-direction suppression. J. Neurosci. 36, 8078–8092 (2016).

    Article  CAS  Google Scholar 

  14. Fransen, J. W. & Borghuis, B. G. Temporally diverse excitation generates direction-selective responses in ON-and OFF-type retinal starburst amacrine cells. Cell Rep. 18, 1356–1365 (2017).

    Article  CAS  Google Scholar 

  15. Jin, L. et al. Single action potentials and subthreshold electrical events imaged in neurons with a fluorescent protein voltage probe. Neuron 75, 779–785 (2012).

    Article  CAS  Google Scholar 

  16. Salazar-Gatzimas, E. et al. Direct measurement of correlation responses in Drosophila elementary motion detectors reveals fast timescale tuning. Neuron 92, 227–239 (2016).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  18. Sweeney, S. T., Broadie, K., Keane, J., Niemann, H. & O’Kane, C. J. Targeted expression of tetanus toxin light chain in Drosophila specifically eliminates synaptic transmission and causes behavioral defects. Neuron 14, 341–351 (1995).

    Article  CAS  Google Scholar 

  19. Heeger, D. J. Half-squaring in responses of cat striate cells. Vis. Neurosci. 9, 427–443 (1992).

    Article  CAS  Google Scholar 

  20. Vintch, B., Movshon, J. A. & Simoncelli, E. P. A convolutional subunit model for neuronal responses in macaque V1. J. Neurosci. 35, 14829–14841 (2015).

    Article  CAS  Google Scholar 

  21. Kato, S., Xu, Y., Cho, C. E., Abbott, L. & Bargmann, C. I. Temporal responses of C. elegans chemosensory neurons are preserved in behavioral dynamics. Neuron 81, 616–628 (2014).

    Article  CAS  Google Scholar 

  22. Rust, N. C., Mante, V., Simoncelli, E. P. & Movshon, J. A. How MT cells analyze the motion of visual patterns. Nat. Neurosci. 9, 1421–1431 (2006).

    Article  CAS  Google Scholar 

  23. Haag, J., Arenz, A., Serbe, E., Gabbiani, F. & Borst, A. Complementary mechanisms create direction selectivity in the fly. eLife 5, e17421 (2016).

    Article  Google Scholar 

  24. Fisher, Y. E., Silies, M. & Clandinin, T. R. Orientation selectivity sharpens motion detection in Drosophila. Neuron 88, 390–402 (2015).

    Article  CAS  Google Scholar 

  25. Takahashi, T., Kajikawa, Y. & Tsujimoto, T. G-Protein-coupled modulation of presynaptic calcium currents and transmitter release by a GABAB receptor. J. Neurosci. 18, 3138–3146 (1998).

    Article  CAS  Google Scholar 

  26. Takemura, S.Y. et al. The comprehensive connectome of a neural substrate for ‘ON’ motion detection in Drosophila. eLife 6, e24394 (2017).

    Article  Google Scholar 

  27. Strother, J. A. et al. The emergence of directional selectivity in the visual motion pathway of Drosophila. Neuron 94, 168–182.e110 (2017.

    Article  CAS  Google Scholar 

  28. Torre, V. & Poggio, T. A synaptic mechanism possibly underlying directional selectivity to motion. Proc. R. Soc. Lond. B 202, 409–416 (1978).

    Article  Google Scholar 

  29. Fitzgerald, J.E., . & Clark, D.A. Nonlinear circuits for naturalistic visual motion estimation. eLife 4, e09123 (2015).

    Article  Google Scholar 

  30. Meyer, H. G., Schwegmann, A., Lindemann, J. P. & Egelhaaf, M. Panoramic high dynamic range images in diverse environments. Publikationen an der Universität Bielefeld https://doi.org/10.4119/unibi/2689637 (2014).

  31. Dror, R. O., O’Carroll, D. C. & Laughlin, S. B. Accuracy of velocity estimation by Reichardt correlators. J. Opt. Soc. Am. A 18, 241–252 (2001).

    Article  CAS  Google Scholar 

  32. Brinkworth, R. S. & O’Carroll, D. C. Robust models for optic flow coding in natural scenes inspired by insect biology. PLoS Comp. Biol. 5, e1000555 (2009).

    Article  Google Scholar 

  33. Clark, D. A., Bursztyn, L., Horowitz, M. A., Schnitzer, M. J. & Clandinin, T. R. Defining the computational structure of the motion detector in Drosophila. Neuron 70, 1165–1177 (2011).

    Article  CAS  Google Scholar 

  34. Salazar-Gatzimas, E., Agrochao, M., Fitzgerald, J. E. & Clark, D. A. The neuronal basis of an illusory motion percept is explained by decorrelation of parallel motion pathways. Curr. Biol. 28, 3748–3762.e8 (2018.

    Article  CAS  Google Scholar 

  35. Zheng, S., Shi, N.-Z. & Zhang, Z. Generalized measures of correlation for asymmetry, nonlinearity, and beyond. J. Am. Stat. Assoc. 107, 1239–1252 (2012).

    Article  CAS  Google Scholar 

  36. Adelson, E. H. & Movshon, J. A. Phenomenal coherence of moving visual patterns. Nature 300, 523 (1982).

    Article  CAS  Google Scholar 

  37. Movshon, J. A. & Newsome, W. T. Visual response properties of striate cortical neurons projecting to area MT in macaque monkeys. J. Neurosci. 16, 7733–7741 (1996).

    Article  CAS  Google Scholar 

  38. Saleem, A. B., Longden, K. D., Schwyn, D. A., Krapp, H. G. & Schultz, S. R. Bimodal optomotor response to plaids in blowflies: mechanisms of component selectivity and evidence for pattern selectivity. J. Neurosci. 32, 1634–1642 (2012).

    Article  CAS  Google Scholar 

  39. Harris, R. A., O’Carroll, D. C. & Laughlin, S. B. Contrast gain reduction in fly motion adaptation. Neuron 28, 595–606 (2000).

    Article  CAS  Google Scholar 

  40. Zanker, J. M., Srinivasan, M. V. & Egelhaaf, M. Speed tuning in elementary motion detectors of the correlation type. Biol. Cybern. 80, 109–116 (1999).

    Article  CAS  Google Scholar 

  41. Borst, A. & Egelhaaf, M. Principles of visual motion detection. Trends Neurosci. 12, 297–306 (1989).

    Article  CAS  Google Scholar 

  42. Creamer, M. S., Mano, O. & Clark, D. A. Visual control of walking speed in Drosophila. Neuron 100, 1460–1473 (2018).

    Article  CAS  Google Scholar 

  43. Simoncelli, E. P. & Heeger, D. J. A model of neuronal responses in visual area MT. Vis. Res. 38, 743–761 (1998).

    Article  CAS  Google Scholar 

  44. Lien, A.D., . & Scanziani, M. Cortical direction selectivity emerges at convergence of thalamic synapses. Nature 558, 80–86 (2018).

    Article  CAS  Google Scholar 

  45. Wilson, D.E., Scholl, B. & Fitzpatrick, D. Differential tuning of excitation and inhibition shapes direction selectivity in ferret visual cortex. Nature 560, 97–101 (2018).

    Article  CAS  Google Scholar 

  46. Lee, S. & Zhou, Z. J. The synaptic mechanism of direction selectivity in distal processes of starburst amacrine cells. Neuron 51, 787–799 (2006).

    Article  CAS  Google Scholar 

  47. Clark, D. A. & Demb, J. B. Parallel computations in insect and mammalian visual motion processing. Curr. Biol. 26, R1062–R1072 (2016).

    Article  CAS  Google Scholar 

  48. Mo, C.-H. & Koch, C. Modeling reverse-phi motion-selective neurons in cortex: double synaptic-veto mechanism. Neural Comput. 15, 735–759 (2003).

    Article  Google Scholar 

  49. Serbe, E., Meier, M., Leonhardt, A. & Borst, A. Comprehensive characterization of the major presynaptic elements to the Drosophila OFF motion detector. Neuron 89, 829–841 (2016).

    Article  CAS  Google Scholar 

  50. Srinivasan, M., Laughlin, S. & Dubs, A. Predictive coding: a fresh view of inhibition in the retina. Proc. R. Soc. Lond. B 216, 427 (1982).

    Article  CAS  Google Scholar 

  51. Clark, D. A. et al. Flies and humans share a motion estimation strategy that exploits natural scene statistics. Nat. Neurosci. 17, 296–303 (2014).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  53. Pologruto, T. A., Sabatini, B. L. & Svoboda, K. ScanImage: flexible software for operating laser scanning microscopes. Biomed. Eng. Online 2, 13 (2003).

    Article  Google Scholar 

  54. Mukamel, E. A., Nimmerjahn, A. & Schnitzer, M. J. Automated analysis of cellular signals from large-scale calcium imaging data. Neuron 63, 747–760 (2009).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  56. Kleiner, M. et al. What’s new in Psychtoolbox-3. Perception 36, 1–16 (2007).

    Google Scholar 

  57. Pelli, D. G. The VideoToolbox software for visual psychophysics: transforming numbers into movies. Spat. Vis. 10, 437–442 (1997).

    Article  CAS  Google Scholar 

  58. Hollander, M., Wolfe, D. A. & Chicken, E. Nonparametric Statistical Methods Vol. 751 (Hoboken, NJ, USA, John Wiley & Sons, 2013).

  59. Stavenga, D. Angular and spectral sensitivity of fly photoreceptors. II. Dependence on facet lens F-number and rhabdomere type in Drosophila. J. Comp. Physiol. A 189, 189–202 (2003).

    CAS  Google Scholar 

  60. Arenz, A., Drews, M. S., Richter, F. G., Ammer, G. & Borst, A. The temporal tuning of the Drosophila motion detectors is determined by the dynamics of their input elements. Curr. Biol. 27, 929–944 (2017).

    Article  CAS  Google Scholar 

  61. Leonhardt, A. et al. Asymmetry of Drosophila ON and OFF motion detectors enhances real-world velocity estimation. Nat. Neurosci. 19, 706–715 (2016).

    Article  CAS  Google Scholar 

  62. Behnia, R., Clark, D. A., Carter, A. G., Clandinin, T. R. & Desplan, C. Processing properties of ON and OFF pathways for Drosophila motion detection. Nature 512, 427–430 (2014).

    Article  CAS  Google Scholar 

  63. Yang, H. H. et al. Subcellular imaging of voltage and calcium signals reveals neural processing in vivo. Cell 166, 245–257 (2016).

    Article  CAS  Google Scholar 

  64. Clark, D. A., Benichou, R., Meister, M. & da Silveira, R. A. Dynamical adaptation in photoreceptors. PLoS Comp. Biol. 9, e1003289 (2013).

    Article  Google Scholar 

  65. Borst, A., Egelhaaf, M. & Haag, J. Mechanisms of dendritic integration underlying gain control in fly motion-sensitive interneurons. J. Comput. Neurosci. 2, 5–18 (1995).

    Article  CAS  Google Scholar 

  66. Freifeld, L., Clark, D. A., Schnitzer, M. J., Horowitz, M. A. & Clandinin, T. R. GABAergic lateral interactions tune the early stages of visual processing in Drosophila. Neuron 78, 1075–1089 (2013).

    Article  CAS  Google Scholar 

  67. Efron, B. Better bootstrap confidence intervals. J. Am. Stat. Assoc. 82, 171–185 (1987).

    Article  Google Scholar 

Download references

Acknowledgements

We thank J.E. Fitzgerald for alerting us to the intuitive explanation of multiplicative opponency with one biphasic filter, as well as for helpful conceptual comments and suggestions. We thank J.B. Demb, H.H. Clark, the members of the Clark Lab and our anonymous reviewers for helpful comments on the manuscript. The Arclight construct was a gift from V. Pieribone. M.S.C. was supported by an NSF GRF. D.A.C. and this research were supported by NIH R01EY026555, NIH P30EY026878, NSF IOS1558103, a Searle Scholar Award, a Sloan Fellowship in Neuroscience, the Smith Family Foundation and the E. Matilda Ziegler Foundation.

Author information

Authors and Affiliations

Authors

Contributions

B.A.B., M.S.C. and D.A.C. designed experiments. B.A.B. acquired and analyzed data. B.A.B., M.S.C., J.A.Z.V. and D.A.C. analyzed models. B.A.B., M.S.C., J.A.Z.V. and D.A.C. wrote the paper.

Corresponding author

Correspondence to Damon A. Clark.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information: Nature Neuroscience thanks Holger Krapp and 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 Opponency in modeling and behavior.

a) Intensity plots of the sinusoidal gratings moving in the cell’s preferred direction (PD) and anti-preferred null direction (ND), and the sum of the gratings. b) Summary of classical direction opponency. In the classic view, opponent direction-selectivity depends on opposing motion stimuli producing responses with opposite signs. c) An alternative scheme for detecting opponency in the presence of signal rectification. It considers a signal to be opponent if the addition of an ND stimulus suppresses responses to PD stimuli. d) Schematic diagram of a Hassenstein-Reichardt Correlator (HRC) model of direction-selectivity. This model multiplies the instantaneous contrast at one point in space with the delayed contrast at a second point in space to create DS signals. The delay is represented by the filter in the box on the vertical arms of the model. By subtracting two of these multiplied signals with opposite directional tuning, the final signal is DO. e) Schematic of a generic DO model. f) Schematic of LN model that creates a DS signal using a spatiotemporally oriented linear filter that is convolved with the stimulus and followed by a nonlinearity. The oriented space-time filter acts to enhance the variance of stimuli with the same orientation (direction). When the nonlinearity is purely quadratic, the result is a pure motion energy model. g) The experimental apparatus used to measure behavioral responses. The fly is tethered to pin and placed on an air-suspended ball in front of a panoramic visual display. h) Behavioral optomotor responses themselves exhibit classical opponency. The turning response to clockwise (CW) rotational sinewave gratings (blue, n = 69 flies) produces a CW response that is opposite in sign to the response to counterclockwise (CCW) sinewave gratings (orange, n = 69 flies). The combination of the two (purple, n = 19 flies) produces no response. Shaded region indicates stimulus duration.

Supplementary Figure 2 Composite sinusoidal grating contrast distributions.

a) Contrast probability density function of a single sinusoidal grating. b) Contrast probability density function of a composite sinusoidal grating with components moving in opposing directions. c) Contrast probability density function of a composite sinusoidal grating with components moving in orthogonal directions.

Supplementary Figure 3 Glider stimuli generate preferred-direction enhancement and null-direction suppression.

a) Space-time intensity plots of two-point gliders at two temporal update rates (5Hz for top panels, 60 Hz for bottom), that are either uncorrelated (left column), correlated to the right (middle column), or correlated to the left (right column). Gliders consist of binary stimuli with enforced correlations over space and time1,2. Duration of plots is 1 second; azimuthal extent is 60º. b) The responses of layer 1 T5 cells in both layers to the gliders specified in (a) (n = 10 flies for 5Hz, and n = 15 flies for 60Hz). Time traces show the cells increase their responses to gliders correlated in their PD relative to uncorrelated at both update rates. However, the response to motion in their ND only decreases relative to uncorrelated at the lower update rate. Shaded region indicates stimulus duration. c) The responses of layer 2 T5 cells to the same stimuli, showing the same phenomenon as in (b). Shaded region indicates stimulus duration. d) Averaged responses of T4 cells to the gliders with 5 Hz temporal update rate (puncorr,PD = 0.0012, puncorr,ND = 0.0002, n = 13 flies). e) Averaged responses of T4 cells to the gliders with 20 Hz temporal update rate (puncorr,PD = 0.0020, puncorr,ND = 0.85, n = 10 flies). f) Averaged responses of T4 cells to the gliders with 60 Hz temporal update rate (puncorr,PD = 0.0078, puncorr,ND = 0.074, n = 9 flies). (g-i) Averaged responses of T5 cells to the gliders, as in (d-f). In T4 and T5, the lower update rates show stronger null-direction suppression relative to uncorrelated stimuli (puncorr,PD = 0.0020, puncorr,ND = 0.0020 and n = 10 flies for 5 Hz; puncorr,PD = 0.0001, puncorr,ND = 0.010 and n = 14 flies for 20 Hz; puncorr-PD = 0.0010, puncorr,ND = 0.25, n = 15 flies for 60 Hz). (* p < 0.05, ** p < 0.01, *** p < 0.001, by a paired two-sided Wilcoxon signed-rank test).

Supplementary Figure 4 Tetanus toxin expression in T4 and T5 eliminates rotational behavioral responses to motion.

Sinewave gratings of various temporal frequencies were presented to flies and mean rotational behavioral responses were recorded. When tetanus toxin was expressed in T4 and T5, the flies failed to respond to these rotational stimuli, though genetic controls did. This shows that tetanus toxin expression successfully blocked synaptic transmission from these cells. The genotype in this experiment is identical to the genotype used to measure calcium signals in T4 and T5 in Fig. 5. (+ > TNT n = 9 flies; T4T5>GCaMP n = 9 flies; T4T5>GCaMP,TNT n = 3 flies).

Supplementary Figure 5 Linear-nonlinear models for direction-selective computations.

a) Space-time intensity plots of the sinusoidal gratings moving in one direction, the other direction, and the sum of the gratings. b) Sample spatiotemporal receptive field for a linear-nonlinear (LN) model of T4. c) Sample time traces of PD (blue), ND (orange), and PD + ND (purple) sinusoidal gratings after linear filtering by the receptive field shown in (b). d) Commonly used nonlinearities in LN models of elementary detectors include (clockwise from top left) a quadratic, an exponential, a soft rectifier, and a half-quadratic. All of these nonlinearities are expansive and static. e) LN models with expansive, static nonlinearities cannot be opponent with this suite of sinusoidal inputs (see Supplementary Notes 1 and 2). f) Motion energy linear-nonlinear model consisting of a spatiotemporal linear filter with oriented positive and negative lobes followed by a static quadratic nonlinearity. g) The components of PD and ND motion are presented as (1,1) + (2,2), and (1,2) + (2,1) respectively. h) The responses of a motion energy model to PD and ND motion compared to the sum of responses to their constituent parts. This model produces both PD enhancement and ND suppression relative to the linear prediction based on a decomposition of the stimulus. That is, \(r\left( {1,1} \right) + r\left( {2,2} \right) < r(PD)\) and \(r\left( {1,2} \right) + r\left( {2,1} \right) > r(ND)\). i) A linear-nonlinear model with a compressive (sigmoidal) nonlinearity can be tuned to produce weakly opponent responses at one contrast level, but is non-opponent when contrast is halved. j) The peak responses of T4- and T5-cells to single, full-contrast moving edges are larger than the peak responses to sinusoidal gratings, suggesting that sinusoidal grating inputs do not saturate the calcium signal.

Supplementary Figure 6 Multiplicative model intuition, model response timeseries, model parameter sweeps, and linearity tests.

a) An HRC half-correlator where the temporal filter of one arm is the derivative of the other. The model can be decomposed into separate components. Flipping the sign of the filters allows for the subtraction of the two models with each tuned to opposing directions (delayed arm switched). b) Response timeseries for the rectified multiplier model in Fig. 6b (see Methods for details on timeseries computation). c) Response timeseries for the dynamic gain model in Fig. 6c. d) The effect of sweeping the parameters of the dynamic gain model on the degree of opponency (see Methods for details on parameter sweeps for this and other models). The parameter values chosen are indicated by a red circle. e) The effect of sweeping the parameters of the dynamic gain model on the response to the sum of PD and OD gratings. f) Response timeseries for the three-input conductance model in Fig. 6d. g) As in (f), but for the three-input conductance model. h) As in (g), but for the three-input conductance model. i) The PD and ND voltage responses of the three-input model in Fig. 6d compared to a prediction of the response based on a linear model and responses to counterphase gratings at 8 different spatial phases3.

Supplementary Figure 7 Joint probability distributions of elementary motion detector responses and natural scene velocity.

See Methods for an explanation of how each distribution was computed. Marginal probabilities are shown above and to the left of each joint probability distribution. Model diagrams are shown in Fig. 7. a) Hassenstein-Reichardt Correlator half-detector (‘half-HRC’). b) HRC. c) Rectified half-HRC. d) Rectified HRC.

Supplementary information

Supplementary Figs. 1–7, Supplementary Table 1, and Supplementary Notes 1–3

Reporting Summary

Supplementary Video 1

The stimuli used to probe opponency. A rightward drifting sinusoid grating is shown first (blue), followed by a leftward drifting sinusoid grating (orange). Last, the counterphase grating consists of the sum of the rightward and leftward gratings (purple).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Badwan, B.A., Creamer, M.S., Zavatone-Veth, J.A. et al. Dynamic nonlinearities enable direction opponency in Drosophila elementary motion detectors. Nat Neurosci 22, 1318–1326 (2019). https://doi.org/10.1038/s41593-019-0443-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-019-0443-y

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