27
Accepted manuscripts are peer-reviewed but have not been through the copyediting, formatting, or proofreading process. Copyright © 2018 Nicholas et al. This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International license, which permits unrestricted use, distribution and reproduction in any medium provided that the original work is properly attributed. This Accepted Manuscript has not been copyedited and formatted. The final version may differ from this version. Research Articles: Systems/Circuits Integration of Small- and Wide-Field Visual Features in Target-Selective Descending Neurons of both Predatory and Non-Predatory Dipterans Sarah Nicholas 1 , Jack Supple 3 , Richard Leibbrandt 1 , Paloma T Gonzalez-Bellido 2,3 and Karin Nordström 1 1 Centre for Neuroscience, Flinders University, GPO Box 2100, 5001 Adelaide, SA, Australia 2 Department of Ecology, Evolution, and Behavior, University of Minnesota, St. Paul, MN 55108, USA 3 Department of Physiology, Development and Neuroscience, University of Cambridge, CB2 3EG, UK https://doi.org/10.1523/JNEUROSCI.1695-18.2018 Received: 9 July 2018 Revised: 18 October 2018 Accepted: 21 October 2018 Published: 29 October 2018 Author contributions: S.J.N., J.S., R.L., P.T.G.-B., and K.N. designed research; S.J.N., J.S., and P.T.G.-B. performed research; S.J.N., J.S., P.T.G.-B., and K.N. analyzed data; S.J.N., J.S., R.L., P.T.G.-B., and K.N. edited the paper; S.J.N., J.S., P.T.G.-B., and K.N. wrote the paper; K.N. wrote the first draft of the paper. Conflict of Interest: The authors declare no competing financial interests. This work was funded by the Air Force Office of Scientific Research (FA9550-15-1-0188 to P.T. Gonzalez- Bellido and K. Nordström) and the Australian Research Council (DP170100008 to P.T. Gonzalez-Bellido and K. Nordström and DP180100144 to K. Nordström). We thank Vanshika Sinh for constructive feedback on the manuscript, Malin Thyselius for the Eristalis pictograms and The Adelaide Botanic Gardens for continuous support. Corresponding author: Karin Nordström, [email protected] Cite as: J. Neurosci 2018; 10.1523/JNEUROSCI.1695-18.2018 Alerts: Sign up at www.jneurosci.org/alerts to receive customized email alerts when the fully formatted version of this article is published.

Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

Accepted manuscripts are peer-reviewed but have not been through the copyediting, formatting, or proofreadingprocess.

Copyright © 2018 Nicholas et al.This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International license, whichpermits unrestricted use, distribution and reproduction in any medium provided that the original work is properly attributed.

This Accepted Manuscript has not been copyedited and formatted. The final version may differ from this version.

Research Articles: Systems/Circuits

Integration of Small- and Wide-Field Visual Features in Target-SelectiveDescending Neurons of both Predatory and Non-Predatory Dipterans

Sarah Nicholas1, Jack Supple3, Richard Leibbrandt1, Paloma T Gonzalez-Bellido2,3 and Karin Nordström1

1Centre for Neuroscience, Flinders University, GPO Box 2100, 5001 Adelaide, SA, Australia2Department of Ecology, Evolution, and Behavior, University of Minnesota, St. Paul, MN 55108, USA3Department of Physiology, Development and Neuroscience, University of Cambridge, CB2 3EG, UK

https://doi.org/10.1523/JNEUROSCI.1695-18.2018

Received: 9 July 2018

Revised: 18 October 2018

Accepted: 21 October 2018

Published: 29 October 2018

Author contributions: S.J.N., J.S., R.L., P.T.G.-B., and K.N. designed research; S.J.N., J.S., and P.T.G.-B.performed research; S.J.N., J.S., P.T.G.-B., and K.N. analyzed data; S.J.N., J.S., R.L., P.T.G.-B., and K.N.edited the paper; S.J.N., J.S., P.T.G.-B., and K.N. wrote the paper; K.N. wrote the first draft of the paper.

Conflict of Interest: The authors declare no competing financial interests.

This work was funded by the Air Force Office of Scientific Research (FA9550-15-1-0188 to P.T. Gonzalez-Bellido and K. Nordström) and the Australian Research Council (DP170100008 to P.T. Gonzalez-Bellido andK. Nordström and DP180100144 to K. Nordström). We thank Vanshika Sinh for constructive feedback on themanuscript, Malin Thyselius for the Eristalis pictograms and The Adelaide Botanic Gardens for continuoussupport.

Corresponding author: Karin Nordström, [email protected]

Cite as: J. Neurosci 2018; 10.1523/JNEUROSCI.1695-18.2018

Alerts: Sign up at www.jneurosci.org/alerts to receive customized email alerts when the fully formatted versionof this article is published.

Page 2: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

1

Integration of Small- and Wide-Field Visual Features in Target-Selective Descending Neurons of 1

both Predatory and Non-Predatory Dipterans 2

Sarah Nicholas*1, Jack Supple*3, Richard Leibbrandt1, Paloma T Gonzalez-Bellido#2,3, Karin 3

Nordström#1 4

1Centre for Neuroscience, Flinders University, GPO Box 2100, 5001 Adelaide, SA, Australia; 5

2Department of Ecology, Evolution, and Behavior, University of Minnesota, St. Paul, MN 55108, USA; 6

3Department of Physiology, Development and Neuroscience, University of Cambridge, CB2 3EG, UK 7

Corresponding author: Karin Nordström, [email protected] 8

*These authors contributed equally 9

#These authors contributed equally 10

Number of pages: 21 pages 11

Number of figures, tables, multimedia and 3D models (separately): 4 figures, 1 table 12

Number of words for Abstract, Introduction, and Discussion (separately) 13

Abstract: 245 14

Introduction: 641 15

Discussion: 1363 16

Conflict of Interest 17

The authors declare no conflict of interest. 18

Acknowledgements 19

This work was funded by the Air Force Office of Scientific Research (FA9550-15-1-0188 to P.T. 20

Gonzalez-Bellido and K. Nordström) and the Australian Research Council (DP170100008 to P.T. 21

Gonzalez-Bellido and K. Nordström and DP180100144 to K. Nordström). We thank Vanshika Sinh for 22

constructive feedback on the manuscript, Malin Thyselius for the Eristalis pictograms and The Adelaide 23

Botanic Gardens for continuous support. 24

Keywords 25

Motion vision, target detection, target tracking, optic flow, dipteran TSDN, electrophysiology, neural 26

tuning, background clutter 27

Page 3: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

2

Abstract 28

For many animals, target motion carries high ecological significance as this may be generated by a 29

predator, prey or potential mate. Indeed, animals whose survival depends on early target detection are 30

often equipped with a sharply tuned visual system, yielding robust performance in challenging 31

conditions. For example, many fast-flying insects use visual cues for identifying targets, such as prey 32

(e.g. predatory dragonflies and robberflies) or conspecifics (e.g. non-predatory hoverflies), and can 33

often do so against self-generated background optic flow. Supporting these behaviors, the optic lobes 34

of insects that pursue targets harbor neurons that respond robustly to the motion of small moving 35

objects, even when displayed against syn-directional background clutter. However, in diptera, the 36

encoding of target information by the descending neurons, which are more directly involved in 37

generating the behavioral output, has received less attention. We characterized target selective 38

neurons by recording in the ventral nerve cord of male and female predatory Holcocephala fusca 39

robberflies and of male non-predatory Eristalis tenax hoverflies. We show that both species have 40

dipteran target-selective descending neurons (dTSDNs) that only respond to target motion if the 41

background is stationary or moving slowly, moves in the opposite direction, or has un-naturalistic 42

spatial characteristics. The response to the target is suppressed when background and target move at 43

similar velocities, which is strikingly different to the response of target neurons in the optic lobes. As the 44

neurons we recorded from are pre-motor, our findings affect our interpretation of the neurophysiology 45

underlying target-tracking behaviors. 46

47

Significance Statement 48

Many animals use sensory cues to detect moving targets that may represent predators, prey or 49

conspecifics. For example, birds of prey show superb sensitivity to the motion of small prey, and 50

intercept these at high speeds. In a similar manner, predatory insects visually track moving prey, often 51

against cluttered backgrounds. Accompanying this behavior, the brains of insects that pursue targets 52

contain neurons that respond exclusively to target motion. We here show that dipteran insects also 53

have target selective descending neurons in the part of their nervous system that corresponds to the 54

vertebrate spinal cord. Surprisingly, and in contrast to the neurons in the brain, these pre-motor 55

neurons are inhibited by background patterns moving in the same direction as the target. 56

57

Page 4: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

3

Introduction 58

Target detection and tracking serve important biological functions for animals to efficiently avoid 59

predators, find prey or identify conspecifics. Target detection can be performed by different senses. 60

Bats find prey with echolocation (Falk et al., 2014), squid embryos avoid predators with their lateral line 61

system (York et al., 2016), and insects visually identify conspecifics (Land and Collett, 1974). Often, 62

targets need to be visualized against self-generated optic flow, which is a difficult computational task 63

(Yang et al., 2012; Held et al., 2016), especially in conditions where both local luminance and relative 64

contrast may change rapidly (Mohamed et al., 2014; Ma et al., 2015). Nonetheless, many insects 65

appear to have solved this efficiently, as evidenced by their high-speed pursuits of targets, which is 66

particularly impressive considering that insects carry low-spatial resolution compound eyes and small 67

brains (Land, 1997). 68

Notably, many insects that pursue targets display local adaptations. For example, the eyes of target-69

pursuing insects often have faster photoreceptors (Weckström and Laughlin, 1995; Burton and 70

Laughlin, 2003; Gonzalez-Bellido et al., 2011) and a region with increased spatial resolution, a fovea, in 71

which they try to place the image of the target during aerial pursuits (Collett, 1980; Olberg et al., 2007; 72

Wardill et al., 2017). The optic lobes harbor target sensitive neurons with receptive fields that often 73

collocate with the optical fovea (Strausfeld, 1980; Barnett et al., 2007), such as small target motion 74

detector (STMD) neurons (O'Carroll, 1993; Nordström et al., 2006). STMDs could relay signals to 75

target-selective descending neurons (TSDNs; Olberg, 1981; Olberg, 1986; Gronenberg and Strausfeld, 76

1990; Namiki et al., 2018) whose receptive fields also colocate with the optical fovea (Gonzalez-Bellido 77

et al., 2013). 78

In the lobula, some hoverfly and dragonfly STMDs show remarkably robust responses to targets 79

moving in visual clutter (Nordström and O'Carroll, 2009). In fact, the STMD response to target motion is 80

unaffected by the addition of background motion with the same velocity (i.e. syn-directional motion, 81

without relative velocity differences; Nordström et al., 2006; Wiederman and O'Carroll, 2011). This can 82

be modelled (Wiederman et al., 2008) using the associated unique spatio-temporal profile, where the 83

target’s dark contrast edge (OFF) is immediately followed by a bright contrast change (ON). Models 84

incorporating temporal correlation of half-wave rectified ON and OFF signals from the same point in 85

space, together with fast adaptation and strong spatial inhibition, lead to robust responses to target 86

motion in visual clutter (Wiederman et al., 2008). To date it is largely unclear how TSDNs respond to 87

targets moving in clutter (except one data trace, see Olberg, 1986). This is important as TSDNs are 88

closer to the motor output (Gonzalez-Bellido et al., 2013), and may thus be more pertinent for 89

interpreting the behavioral relevance of neurophysiological recordings. Besides the TSDNs, there are 90

Page 5: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

4

many other descending neurons (Namiki et al., 2018), such as those that respond to background 91

motion (Wertz et al., 2009b; Wertz et al., 2009a). These project to areas of the thoracic ganglia with 92

motor neurons of the neck, wings and halteres (Suver et al., 2016). 93

As the ventral nerve cord forms an information bottleneck between the central processing of visually 94

driven information and peripheral transformation into motor output (Namiki et al., 2018), we here 95

quantified the response to moving targets and visual clutter. For this purpose, we characterized 96

dTSDNs in the predatory robberfly Holcocephala fusca and the non-predatory hoverfly Eristalis tenax, 97

the latter of which pursue conspecifics. As the two animals pursue targets for different reasons, and are 98

evolutionarily distant, any similarities could potentially inform us about general mechanisms underlying 99

dipteran target detection. In Eristalis we also identified neurons sensitive to visual clutter. We found that 100

Holcocephala and Eristalis dTSDNs did not respond to targets moving across background clutter, 101

unless the background moved in the opposite direction, moved slowly, or was highly un-naturalistic, 102

suggesting that dTSDNs receive inhibitory input from pre-synaptic neurons tuned to wide-field optic 103

flow. 104

Materials and methods 105

Animals and electrophysiology 106

38 male Eristalis tenax hoverflies were reared from eggs laid by wild-caught hoverflies and housed as 107

described (Nicholas et al., 2018). 16 female and 3 male adult Holcocephala fusca robberflies were wild 108

caught and recorded from on the day of capture. Prior to recording, the animal was immobilized dorsal 109

side down, and a small hole cut at the antero-ventral thoracic surface to expose the ventral nerve cord. 110

For Eristalis extracellular recordings, a sharp polyimide-insulated tungsten electrode (0.1 MOhm, 111

Microprobes, Gaithersburg, USA) was inserted into the nerve cord, with mechanical support given to 112

the cord by a small wire hook. The animal was grounded via a silver wire inserted into the ventral 113

cavity, which also served as the recording reference. To prevent drying up of the exposed ventral cavity 114

a small amount of petroleum jelly and mineral oil mix (1:1 ratio) was applied. Extracellular signals were 115

amplified at 1000x gain and filtered through a 10 – 3000 Hz bandwidth filter on a DAM50 differential 116

amplifier (World Precision Instruments), filtered through a HumBug (Quest Scientific, North Vancouver, 117

Canada), digitized via a Powerlab 4/30 (ADInstruments, Sydney, Australia), and acquired at 10 kHz 118

with LabChart 7 Pro software (ADInstruments). 119

For Holcocephala extracellular recordings, a sharp glass-insulated tungsten electrode (2-4 MOhm, 120

Microelectrodes Ltd., Cambridge, UK) was inserted into the nerve cord, with mechanical support given 121

to the cord by a small hook fashioned from a hypodermic needle. The animal was grounded by a 122

Page 6: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

5

saline-filled microelectrode inserted into the ventral cavity, which also served as the recording 123

reference. Fly saline was prepared as follows (Gengs et al., 2002): 138 mM NaCl, 2 mM KCl, 1.8 mM 124

CaCl2, 4 mM MgCl2, 5 mM TES, pH 7.15. Hydration of the ventral cavity was maintained by continual 125

capillary action from an additional saline-filled microelectrode. Extracellular signals were amplified at 126

500x gain and filtered through a 300 – 3000 Hz analogue bandpass filter on an NPI-BA amplifier (NPI 127

Electronic, Germany), filtered through a HumBug (Digitimer, UK), digitized on a micro1401 DAQ 128

(Cambridge Electronic Design Limited, UK), and acquired at 25 kHz with Spike2 software (Cambridge 129

Electronic Design Limited, UK). 130

Visual stimuli 131

For Eristalis experiments, visual stimuli were displayed on an Asus LCD screen (Asus, Taipei, Taiwan) 132

with a spatial resolution of 2560 x 1440 pixels running at 165 Hz, using the Psychophysics toolbox in 133

Matlab (Mathworks, 2017). Eristalis males were placed at a distance of 7 cm, giving a projected screen 134

size of 154 x 137 degrees. For Holcocephala experiments, visual stimuli were projected onto a 17.3 x 135

9.6 cm white screen using a DepthQ 360 projector (Cambridge Research Systems) with a spatial 136

resolution of 1280 x 720 pixels running at 360 Hz, using StimulateOpenGL software (version 20160216, 137

Janelia Research Campus, https://github.com/cculianu/StimulateOpenGL_II). Holcocephala animals 138

were placed at a distance of 7 cm, giving a projected screen size of 102 x 70 degrees. 139

For all dTSDN experiments, we first mapped the receptive field (Nordström et al., 2006; Gonzalez-140

Bellido et al., 2013) and confirmed that the neuron responded selectively to small targets and not to 141

wide-field optic flow (Fig. 1) or to looming stimuli. For all target-background experiments the targets 142

moved either horizontally (Eristalis) or vertically (Holcocephala) across the center of each neuron’s 143

receptive field. Only data from neurons showing both a robust and consistent response to the target 144

moving over a gray background throughout the recording were included in this study. For size-tuning 145

experiments in Eristalis we moved a black target with a fixed width (3 degrees) horizontally across a 146

white background at an average velocity of 180 degrees per second (since we used flat screens, the 147

projected angular velocity varied between their central and peripheral parts). We varied the vertical 148

extent of the target from the smallest we could display on the screen (0.2 degrees), to bars that covered 149

its entire height (137 degrees). For size-tuning experiments in Holcocephala a small black square target 150

appeared in a random location anywhere on the screen, remained stationary for 150 ms and then 151

moved in a random direction for 100 ms, in a total of 2400 random trajectories (as in Gonzalez-Bellido 152

et al., 2013). 153

Page 7: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

6

For target-background experiments in both Eristalis and Holcocephala we used an artificially generated 154

naturalistic background pattern with a slope constant of the amplitude spectrum and RMS-contrast 155

close to those of natural scenes (Dyakova and Nordström, 2017). To generate this background pattern, 156

we used the fact that the spatial statistics of an image can be quantified by doing a Fast Fourier 157

transform and plotting the rotationally averaged amplitude as a function of spatial frequency (Tolhurst et 158

al., 1992). Displayed in a log-log graph the amplitude is then inversely proportional to the spatial 159

frequency raised to the power alpha (Dyakova and Nordström, 2017). In natural scenes, the alpha, also 160

referred to as the slope constant, is close to 1.1 (Tolhurst et al., 1992; Dyakova and Nordström, 2017), 161

which we used here. We set the mean luminance of the background to the mean of the screen, and 162

always covered the entire screen. In Eristalis the target (3x3 degrees) moved horizontally across the 163

screen at an average velocity of 180 degrees per second. In Holcocephala the target (2x2 degrees) 164

moved vertically across the screen at an average velocity of 160 degrees per second. Unless otherwise 165

indicated, the target and the background moved at the same velocity. 166

In Eristalis, we defined optic flow sensitive neurons based on their receptive field properties and 167

response to a high-contrast sinusoidal grating moving in 8 different directions (wavelength of 7 degrees 168

drifting at 5 Hz, Fig. 1Biv), using blowfly data as comparison (Wertz et al., 2009b; Wertz et al., 2009a). 169

We displayed the same artificially generated naturalistic background pattern as in the target-170

background experiments described above, but without the target. The pattern moved horizontally 171

across the screen, unless otherwise indicated. 172

Holcocephala recording time was limited due to the animals having to be used on the day of capture, 173

and only being available for a maximum of 8 weeks of the year. Therefore, we were able to explore 174

more stimulus parameters in Eristalis. We recorded from 27 dTSDNs and 29 neurons sensitive to wide-175

field motion in 38 male Eristalis and 38 dTSDNs in 3 male and 16 female Holcocephala. In all 176

experiments, in both species, the stimulus trial order was randomized. 177

Experimental Design and Statistical Analysis 178

Spike sorting of Eristalis extracellular data (Fig. 1Ai, Bi) was done using LabChart 7 Pro with the Spike 179

Histogram Add-On (ADInstruments, Sydney, Australia), which uses the action potential amplitude and 180

width to identify responses from individual neurons (insets, Fig. 1Ai, Bi). In addition, we quantified the 181

interspike intervals (Fig. 1Aii, Bii) from the resulting spike trains. All further data analysis was done in 182

Matlab. Holcocephala extracellular data (Fig. 1Ci) was sorted in Spike2 (Cambridge Electronic Design 183

Limited, UK), which uses principal component analysis (PCA) on the waveform shape (inset, Fig. 1Ci), 184

followed by manual clustering. 185

Page 8: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

7

For all experiments in Eristalis TSDNs, we quantified the mean spike frequency for the time that the 186

target traversed each neuron’s receptive field (bar under data in Fig. 1Aiii). For optic flow sensitive 187

neurons, we quantified the mean spike frequency for the entire stimulus duration (bar under data in Fig. 188

1Biii). The dTSDN neurons were not spontaneously active (Fig. 1A), but the optic flow sensitive 189

neurons sometimes had a spontaneous rate, which is indicated in each graph. All experiments were 190

repeated 6-18 times in each animal, where we varied the precise target location slightly between trials 191

to avoid habituation. The data from repetitions within a neuron were averaged, with the graphs in the 192

paper showing variation across neurons. 193

Holcocephala and Eristalis responses to differently sized targets were normalized to the maximum 194

response of each neuron. To separate target induced responses from background induced responses 195

in Holcocephala, the response to the same time window when only a background pattern was shown 196

was subtracted from the response to target only motion (bar under data in Fig. 1Ciii). For Holcocephala 197

target-background experiments, each condition consisted of three repetitions of the stimuli. Due to the 198

fast habituation of the neurons in Holcocephala, only the response from the first presentation of the 199

stimuli from each animal was used for further analysis. 200

Statistical analysis was performed in GraphPad Prism (version 7.0d, GraphPad Software Inc, USA), 201

after ensuring that the data were normally distributed, with details of the test and significance given in 202

each figure legend. P-values below 0.05 were used to refute the null hypothesis. 203

Results 204

Descending neuron identification 205

We performed extracellular recordings in the ventral nerve cord of Holcocephala fusca robberflies and 206

Eristalis tenax hoverflies (Fig. 1, top row). We identified individual neurons based on the waveform 207

(including amplitude and width of each action potential, Fig. 1, insets top row). We defined descending 208

visual neurons as dTSDNs by their peak response to small objects subtending a few degrees of the 209

visual field, with no response to elongated bars (Eristalis data, Fig. 1Ai, iv), to larger objects 210

(Holcocephala data, Fig. 1Ci, iv), to looming or to wide-field stimuli (Eristalis data, Fig. 1Biv). The 211

selective response to the motion of small targets in these dTSDNs (Fig. 1) is similar to the response 212

properties of previously described STMDs found in the lobula of Eristalis hoverflies (Nordström et al., 213

2006). These results (Fig. 1A, C) are in accordance with the previously proposed notion that TSDNs 214

may be downstream of the lobula STMD neurons (Barnett et al., 2007; Nordström and O'Carroll, 2009). 215

Whether these neurons are directly or indirectly connected remains to be clarified. 216

Page 9: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

8

In Eristalis, we identified a second group of wide-field sensitive descending visual neurons, which 217

respond to sinusoidal gratings in a direction selective manner (Fig. 1B). We predominantly recorded 218

from two types of wide-field sensitive neurons, which responded preferentially to motion up and to the 219

right across the animal’s visual field (225 degrees, light purple, Fig. 1Biv) and to motion down across 220

the visual field (90 degrees, dark purple, Fig. 1Biv), respectively. The direction tuning (Fig. 1Biv) follows 221

the typical sinusoidal shape seen in similar descending neurons previously described in blowflies and 222

Drosophila, which receive direct input from optic flow-tuned lobula plate tangential cells (LPTCs, Wertz 223

et al., 2009b; Suver et al., 2016). 224

Dipteran TSDNs do not respond to targets moving in the same direction as background clutter 225

In the Eristalis lobula, some STMDs respond robustly to targets moving in visual clutter, even when 226

there are no velocity differences between target and background (Nordström et al., 2006). To 227

investigate whether this property is also present at the dTSDN level, we presented Eristalis and 228

Holcocephala with a small, high-contrast target (3x3 degrees and 2x2 degrees, respectively) moving 229

across a background pattern. The background was artificially generated to have naturalistic spatial 230

statistics and RMS contrast (Dyakova and Nordström, 2017). The target was presented moving 231

horizontally (Eristalis) or vertically (Holcocephala) across the screen, with the background pattern 232

moving in the same direction. As controls, we recorded the response to the target moving over a mean 233

luminance background (open symbols, Fig. 2A), or over the cluttered background presented stationarily 234

(gray symbols, Fig. 2A). The responses to these two controls were not significantly different from each 235

other (Fig. 2A). 236

We found a consistent trend in both species: the presence of background movement substantially 237

reduced the dTSDN responses (Fig. 2A). This effect became significant when the background moved at 238

velocities > 10 degrees per second (green data, Fig. 2A). When the target and the background moved 239

at the same velocity (gray arrow, Fig. 2A), the response to the target’s motion had completely 240

disappeared in both species (Eristalis = green; Holcocephala = red, Fig. 2A). The dTSDN responses 241

were also absent when the background moved two or three times faster than the target (Fig. 2A, data 242

points to the right of gray arrow), or at half the velocity in Eristalis (Fig. 2A, green data point to the left of 243

the gray arrow). It thus seems as if the dTSDNs are unresponsive to targets presented against syn-244

directional background motion, with or without relative velocity differences. 245

We next recorded from optic flow sensitive neurons in the Eristalis descending nerve cord. We found 246

that the response of the wide-field neurons increased with the velocity of the background pattern (Fig. 247

2B). We also found that as the response of the Eristalis wide-field neurons to background velocity 248

Page 10: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

9

increased, the response of the dTSDNs to targets moving over backgrounds with different velocities 249

decreased (Fig. 2C, and see Table 1 for different curve fits to the data). 250

Dipteran TSDNs respond stronger to target motion when the background moves in the opposite 251

direction 252

Next, we tested whether the direction of target and background motion were important in their 253

interactions. We presented the target and background moving at the same speed (as indicated with a 254

gray arrow, Fig. 2A), but tested a variety of background directions (in steps of 45 degrees, closed 255

symbols, Fig. 2D). For comparison, we recorded the response to targets moving over a uniform mean 256

luminance background (open symbols, Fig. 2D). We found that the response to target motion depended 257

on the direction of background motion. As shown above, the response was completely suppressed 258

when the target and the background moved in the same direction (i.e. 0 degrees relative direction 259

difference, Fig. 2D). The dTSDN response to the target’s motion increased when the background 260

moved in the opposite direction to the target (i.e. 180 degrees relative direction difference, Fig. 2D). 261

However, even when the background moved in the opposite direction to the target, the response was 262

significantly lowered to 46% in Eristalis, and 24% in Holcocephala compared to the control condition 263

(no background, Fig. 2D), showing remarkable consistency across the two species. Importantly, the 264

Eristalis response was strongly suppressed even when the background moved at a relative direction of 265

45 degrees (green data, Fig. 2D). 266

In Eristalis we next recorded the response of descending wide-field sensitive neurons to the 267

background pattern moving in different directions (Fig. 2E). In each recording we defined 0 as the 268

neuron’s preferred direction (see Fig. 1Biv for the underlying direction preferences of the two most 269

commonly encountered wide-field sensitive neurons). Again, we found that the response of the Eristalis 270

TSDNs to targets moving over backgrounds in different directions decreased as the response of the 271

wide-field neurons to different directions of background motion increased (Fig. 2F, Table 1). 272

Local mechanisms do no explain dTSDN response suppression from background motion 273

In some STMDs in the Eristalis brain a moving target is detected against background motion even in 274

the absence of relative movement (Nordström et al., 2006), but this ability seems to be gone in the 275

dTSDNs (Fig. 2). Thus, our dTSDN and wide-field descending neuron results (Fig. 2) led us to 276

postulate that the response to target motion is actively suppressed by the wide-field system. The 277

suppression appears to be weaker at low velocities (Fig. 2A, C) and when the background moves in a 278

different direction to the target (Fig. 2D, F). 279

Page 11: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

10

One possible alternative explanation is that the response suppression is caused by a reduced local 280

relative contrast associated with the target moving over the naturalistic background, compared to when 281

it moves over a uniform mean luminance background (open symbols, Fig. 2A, D). To investigate this 282

possibility, we placed a gray mean luminance patch over the background, centered on the target’s 283

trajectory. The patch ensured that the local contrast surrounding the target’s trajectory was equal to the 284

control condition (open symbols, Fig. 2A, D). For this experiment, the target and the background moved 285

at the same velocity. We found that a small cover did not increase the dTSDN response (the smallest 286

cover was 3 times higher than the target, closed symbols, Fig. 3A). It was not until we covered the 287

majority of the background texture (127 degrees height, 77% of the total area) that the dTSDN 288

response to target motion increased significantly, but even then it was significantly lower than control 289

conditions (Fig. 3A). This was surprising because in all these cases the local contrast was the same as 290

in the control condition (open symbol, Fig. 3A). 291

We next recorded the response of wide-field sensitive neurons to the background pattern when it was 292

covered by a similar gray mean luminance “strip” of varying heights. We found that the response of the 293

wide-field neurons decreased linearly with the height of the cover (Fig. 3B). In addition, we found that 294

the response of the Eristalis dTSDNs to different cover heights decreased as the response of the wide-295

field neurons increased when the background texture was covered (Fig. 3C, Table 1). 296

In order to test whether the extent of the background pattern matters, we followed up by inverting the 297

plain and background areas: the strip surrounding the target now contained the original background 298

pattern, and the screen outside of the patch had uniform luminance (i.e. no pattern). We varied the size, 299

and thus the percentage of screen covered, with a moving, patterned background. We found that the 300

dTSDN response to target motion decreased as the height of the patch with the moving background 301

pattern increased. However, for the drop in dTSDN responses to reach significance, the patterned 302

patch had to cover 70% of the screen height (while the projected target only covered 1% of the height; 303

Fig. 3D). By recording from wide-field sensitive neurons we found that the wide-field response 304

increased with the height of the background pattern (Fig. 3E), and that the response of the Eristalis 305

TSDNs decreased as the response of the wide-field neurons increased (Fig. 3F, Table 1). 306

Taken together, these experiments (Fig. 3) suggest that it is not local contrast differences between the 307

target and the background that result in the dTSDNs being silent when the target and background move 308

in the same direction. Rather, it seems as if optic flow information inhibits the dTSDNs. 309

310

Page 12: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

11

Dipteran TSDN response suppression is strongest when the background is most naturalistic 311

The background pattern used in our experiments was artificially generated to have natural image 312

statistics with respect to its contrast and amplitude spectrum (see Dyakova and Nordström, 2017). 313

Natural images have amplitude spectra slope constants of nearly 1 (Tolhurst et al., 1992), to which both 314

peripheral and central sensory neurons are tuned (e.g. van Hateren, 1992; Song and Juusola, 2014; 315

Dyakova et al., 2015). As a final experiment, we varied two background properties: the contrast of the 316

pattern and the slope constant of its amplitude spectrum. The target and the background moved at the 317

same velocity (i.e. there was no relative motion between the two). We found that the dTSDN responses 318

to target motion were higher when the contrast of the background was lower (Fig. 4A). In addition, 319

when the background had medium contrast (0.4), the most naturalistic pattern (slope constant, alpha, of 320

1.1), resulted in a near absence of dTSDN responses (dashed line, grey symbol, Fig. 4A). This was 321

surprising because with the same contrast level, the slope constant, alpha, of 1.8 gave responses 322

similar to controls and the slope constant alpha of 0.5 reached at least half of the control responses. 323

This finding is of importance because it suggests that the suppression of dTSDN responses, shown in 324

this study, is strongest when the target is presented against more naturalistic backgrounds. 325

We recorded the response of Eristalis wide-field neurons to the same background patterns, and found 326

that the response increased with contrast (Fig. 4B). Furthermore, at medium background contrast (0.4) 327

the strongest response was generated by the most naturalistic background pattern (dashed line, grey 328

symbol, alpha of 1.1). Again, the dTSDN response decreased as the response of the optic flow neurons 329

increased when we varied the contrast and amplitude spectrum of the background pattern (Fig. 4C, 330

Table 1). Thus, our findings are consistent with the hypothesis that wide-field sensitive neurons are 331

involved in suppressing the dTSDN responses to target motion. 332

Discussion 333

We have shown that Eristalis and Holcocephala have dipteran target-selective descending neurons, 334

dTSDNs (Fig. 1), with similar size tuning as STMDs in the dragonfly and hoverfly lobula (O'Carroll, 335

1993; Nordström et al., 2006) and dragonfly TSDNs (Olberg, 1986; Gonzalez-Bellido et al., 2013). 336

Dipteran TSDNs thus show physiological homology to dragonfly TSDNs, but whether they are also 337

morphological homologues remains to be studied. Dipteran TSDNs are unresponsive to targets moving 338

in visual clutter (patterned background), unless the background is moving slowly (Fig. 2A), in the 339

opposite direction (Fig. 2D), or has un-naturalistic spatial characteristics (Fig. 4A). In Eristalis we also 340

recorded from neurons that respond to wide-field optic flow, with response properties similar to 341

descending neurons described in blowflies and Drosophila (Wertz et al., 2009b; Suver et al., 2016). In 342

Page 13: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

12

all tested conditions, we found that background stimuli that increased the activity in the wide-field 343

descending neurons led to the dTSDN responses decreasing (Fig. 2C, F, Fig. 3C, F, Fig. 4C, Table 1). 344

Our findings are important as dTSDNs act as a bottleneck between sensory processing and behavioral 345

motor output. Our results demonstrate that dTSDN response suppression induced by background 346

motion is i) not explained by local effects (Fig. 3), and ii) likely caused by pre-synaptic neurons tuned to 347

wide-field optic flow. 348

Target selective neurons 349

We defined Eristalis and Holcocephala neurons as dTSDNs based on their sharp size selectivity, with 350

no response to full-screen bars, large objects, or sinusoidal gratings that drive optic flow sensitive 351

neurons strongly (Fig. 1). Dragonfly TSDNs descend from the brain (Olberg, 1986) and project to the 352

sub-esophageal ganglion and all three thoracic ganglia (Gonzalez-Bellido et al., 2013) where they likely 353

connect with motor neurons of the fore- and the hind-wings. In dipterans, descending visual neurons 354

project to the three thoracic ganglia where they may control neck, leg and/or wing motor neurons 355

(Namiki et al., 2018). The dTSDNs described here are thus likely to provide input to the motor neurons 356

involved in target-pursuit behaviors, by e.g. aligning the head (and thus the fovea) to the image of the 357

target, and/or by rapidly changing flight course as is necessary during high-speed target pursuits 358

(Collett and Land, 1975; Collett and Land, 1978; Wardill et al., 2017). 359

Some target neurons (STMDs) in the lobula respond robustly to the motion of small targets against a 360

cluttered background, even without relative velocity differences (Nordström et al., 2006; Wiederman 361

and O'Carroll, 2011). It was therefore surprising to find that the dTSDN responses, i.e. the responses of 362

the presumed downstream targets of STMDs (Nordström and O'Carroll, 2009), were so strongly 363

affected under similar stimulus conditions (Fig. 2A, D). Since the background inhibition was observed in 364

both Eristalis and Holcocephala (Fig. 2A, D), whose responses were recorded at different times, by 365

different teams and with different instrumentation, this finding is clearly not a species- or experimental 366

specific oddity. Importantly, the target size, as well as the angular velocities of the target and the 367

background were within the range of those experienced during target pursuits in Eristalis (Collett and 368

Land, 1978) and Holcocephala (Wardill et al., 2017). 369

In hoverflies and other non-predatory flies, target detection is primarily used for conspecific 370

identification or territorial interactions (Wellington and Fitzpatrick, 1981), whereas predatory robberflies 371

would use these neurons for detection of suitable prey (Wardill et al., 2017). However, both species 372

intercept their targets, with Holcocephala recently shown to use proportional navigation, a strategy that 373

is likely shared by other dipterans (Fabian et al., 2018). Despite the marked ecological differences, the 374

Page 14: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

13

size tuning (Fig. 1Aiv, Civ) and responses to targets in clutter (Fig. 2A, D) were remarkably similar, 375

suggesting that task-specific constraints resulted in the evolution of target selective neurons with similar 376

response properties across species. 377

Background suppression 378

We found that in the presence of a moving background dTSDNs only responded to the target when the 379

background velocity was low (Fig. 2A) or had un-naturalistic spatial characteristics (Fig. 4A), i.e. under 380

conditions where the background clutter did not drive optic flow sensitive neurons strongly (Fig. 2B, Fig. 381

4B). The relationship between the responses of dTSDNs and the neurons tuned to wide-field optic flow 382

(Fig. 2C, F, Fig. 3C, F, Fig. 4C, Table 1) suggests that activation of the wide-field pathway results in 383

inhibition of the target tracking pathway. At this stage we are not able to pinpoint the type of inhibition 384

as the relationship between the responses of wide-field neurons and dTSDNs could be described by 385

either linear or non-linear functions (Table 1). Since some STMDs in the lobula respond strongly under 386

similar target-background conditions (Nordström et al., 2006; Wiederman and O'Carroll, 2011), it is 387

possible that the inhibition takes place post-synaptic to these STMDs, but pre-synaptic to the dTSDNs. 388

The observed inhibition could be implemented directly by the LPTCs since information from the lobula 389

plate (wide-field motion) and the lobula (target motion) interacts extensively in the protocerebrum 390

(Namiki et al., 2018). LPTCs have wide-reaching output synapses in the posterior slope (Suver et al., 391

2016), where they may synapse with dTSDNs. At least in dragonflies, TSDN input dendrites branch in 392

this area (Olberg, 1986). 393

Suppression from the wide-field system was observed when the target moved over a moving 394

background partially covered by a patch of uniform luminosity (Fig. 3A). Thus, it is unlikely that local 395

contrast mechanisms related to target detection underpin our finding. As some STMDs respond under 396

similar target-background conditions (Nordström et al., 2006), it is more likely caused by the 397

suppression of responses to an already detected target. 398

Why go through all the trouble of extracting a clean target motion signal in the STMDs (Nordström et 399

al., 2006) and then not transform it into a pre-motor command in the dTSDNs (Fig. 2-4)? One possibility 400

is that during actual target pursuits, the target and the background only rarely move in the same 401

direction. Indeed, during pursuits in the hoverfly Syritta pipiens the target is actively foveated, but when 402

the target image leaves the fovea the hoverfly performs saccadic tracking to reduce the target position 403

error (Collett, 1980). During perfect foveation there would be no remaining target motion, only 404

background motion. One way to investigate this hypothesis would be to replay reconstructed target 405

pursuits (Wardill et al., 2017). Furthermore, in highly textured environments Holcocephala and Eristalis 406

Page 15: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

14

use behavioral adaptations to eliminate the influence of background motion. For example, 407

Holcocephala attempt to visualize the target against the clear sky (Wardill et al., 2017), whereas 408

hoverflies often detect the target from a hovering stance, resulting in minimal background motion. In 409

dragonflies, the prey detection distance is decreased by about 1/3 when prey is visualized against close 410

compared to distant background vegetation, respectively (Switzer and Eason, 2000), whereas blowfly 411

target tracking performance is unaffected by background motion (Trischler et al., 2010). 412

Another possibility is that our results are affected by our animals being immobilized, as visual 413

responses are highly affected by the fly’s activity state (Maimon et al., 2010; Jung et al., 2011; Kim et 414

al., 2015). Indeed, many descending neurons are multi-modal (Tsubouchi et al., 2017), and even if 415

TSDNs are predominantly visual (Olberg, 1986), they may be strongly affected by input from other 416

modalities (Huston and Krapp, 2009; Kim et al., 2015; Fujiwara et al., 2017). Furthermore, during high-417

speed pursuit the fly suppresses stabilizing optomotor responses, which would otherwise counteract 418

voluntary turns towards the target (Collett, 1980; Pal, 2015). Indeed, in flying Drosophila the prediction 419

of the expected re-afferent signal is quantitatively subtracted from some LPTCs in anticipation of 420

voluntary turns (Kim et al., 2015). Assuming this is also the case in our model system, LPTCs would not 421

be able to inhibit the dTSDNs as observed in our data (Fig. 2-4). In blowflies an artificial background 422

rotation during conspecific pursuit does not alter the quality of chasing performance (Trischler et al., 423

2010), indicating that target tracking is unaffected by efference copy signals. Investigating these issues 424

will require recording from dTSDNs under actual pursuit with telemetric recordings as is already 425

possible in larger insects (Fotowat et al., 2011; Harrison et al., 2011). This will be informative as 426

Drosophila LPTCs and bee optic lobe neurons respond differently to the same visual stimulus 427

experienced in open or closed loop (Fujiwara et al., 2017)(Paulk et al., 2014). 428

Page 16: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

15

References 429

Barnett PD, Nordström K, O'Carroll DC (2007) Retinotopic organization of small-field-target-detecting 430

neurons in the insect visual system. Curr Biol 17:569-578. 431

Burton BG, Laughlin SB (2003) Neural images of pursuit targets in the photoreceptor arrays of male 432

and female houseflies Musca domestica. J Exp Biol 206:3963-3977. 433

Collett TS (1980) Angular tracking and the optomotor response. An analysis of visual reflex interaction 434

in a hoverfly. J Comp Physiol A 140:145-158. 435

Collett TS, Land MF (1975) Visual control of flight behaviour in the hoverfly, Syritta pipiens L. J Comp 436

Physiol A 99:1-66. 437

Collett TS, Land MF (1978) How hoverflies compute interception courses. J Comp Physiol A 125:191-438

204. 439

Dyakova O, Nordström K (2017) Image statistics and their processing in insect vision. Curr Opin Insect 440

Sci 24:7-14. 441

Dyakova O, Lee Y-J, Longden KD, Kiselev VG, Nordström K (2015) A higher order visual neuron tuned 442

to the spatial amplitude spectra of natural scenes. Nat Commun 6:8522. 443

Fabian ST, Sumner ME, Wardill TJ, Rossoni S, Gonzalez-Bellido PT (2018) Interception by two 444

predatory fly species is explained by a proportional navigation feedback controller. Journal of The 445

Royal Society Interface 15. 446

Falk B, Jakobsen L, Surlykke A, Moss CF (2014) Bats coordinate sonar and flight behavior as they 447

forage in open and cluttered environments. J Exp Biol 217:4356-4364. 448

Fotowat H, Harrison RR, Gabbiani F (2011) Multiplexing of motor information in the discharge of a 449

collision detecting neuron during escape behaviors. Neuron 69:147-158. 450

Fujiwara T, Cruz TL, Bohnslav JP, Chiappe ME (2017) A faithful internal representation of walking 451

movements in the Drosophila visual system. Nat Neurosci 20:72-81. 452

Gengs C, Leung HT, Skingsley DR, Iovchev MI, Yin Z, Semenov EP, Burg MG, Hardie RC, Pak WL 453

(2002) The target of Drosophila photoreceptor synaptic transmission is a histamine-gated chloride 454

channel encoded by ort (hclA). J Biol Chem 277:42113-42120. 455

Gonzalez-Bellido PT, Wardill TJ, Juusola M (2011) Compound eyes and retinal information processing 456

in miniature dipteran species match their specific ecological demands. Proc Natl Acad Sci U S A 457

108:4224-4229. 458

Gonzalez-Bellido PT, Peng H, Yang J, Georgopoulos AP, Olberg RM (2013) Eight pairs of descending 459

neurons in the dragonfly give wing motor centers accurate population vector of prey direction. Proc 460

Natl Acad Sci USA 110:676-701. 461

Page 17: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

16

Gronenberg W, Strausfeld NJ (1990) Descending neurons supplying the neck and flight motor of 462

Diptera: physiological and anatomical characteristics. J Comp Neurol 302:973-991. 463

Harrison RR, Fotowat H, Chan R, Kier RJ, Olberg R, Leonardo A, Gabbiani F (2011) Wireless 464

Neural/EMG Telemetry Systems for Small Freely Moving Animals. IEEE transactions on biomedical 465

circuits and systems 5:103-111. 466

Held D, Thrun S, Savarese S (2016) Learning to Track at 100 FPS with Deep Regression Networks. In: 467

Computer Vision – ECCV 2016 (Leibe B, Matas J, Sebe N, Welling M, eds), pp 749-765. Cham: 468

Springer International Publishing. 469

Huston SJ, Krapp HG (2009) Nonlinear integration of visual and haltere inputs in fly neck motor 470

neurons. J Neurosci 29:12097-13105. 471

Jung SN, Borst A, Haag J (2011) Flight activity alters velocity tuning of fly motion-sensitive neurons. J 472

Neurosci 31:9231-9237. 473

Kim AJ, Fitzgerald JK, Maimon G (2015) Cellular evidence for efference copy in Drosophila visuomotor 474

processing. Nat Neurosci 18:1247-1255. 475

Land MF (1997) Visual acuity in insects. Annu Rev Entomol 42:147-177. 476

Land MF, Collett TS (1974) Chasing behaviour of houseflies (Fannia canicularis). J Comp Physiol A 477

89:331-357. 478

Ma C, Huang JB, Yang X, Yang MH (2015) Hierarchical Convolutional Features for Visual Tracking. In: 479

2015 IEEE International Conference on Computer Vision (ICCV), pp 3074-3082. 480

Maimon G, Straw AD, Dickinson MH (2010) Active flight increases the gain of visual motion processing 481

in Drosophila. Nat Neurosci 13:393-399. 482

Mohamed MA, Böddeker C, Mertsching B (2014) Real-time moving objects tracking for mobile-robots 483

using motion information. In: 2014 IEEE International Symposium on Safety, Security, and Rescue 484

Robotics (2014), pp 1-6. 485

Namiki S, Dickinson MH, Wong AM, Korff W, Card GM (2018) The functional organization of 486

descending sensory-motor pathways in Drosophila. eLife 7. 487

Nicholas S, Thyselius M, Holden M, Nordström K (2018) Rearing and Long-Term Maintenance of 488

Eristalis tenax Hoverflies for Research Studies. JoVE:e57711. 489

Nordström K, O'Carroll DC (2009) Feature detection and the hypercomplex property in insects. Trends 490

Neurosci 32:383-391. 491

Nordström K, Barnett PD, O'Carroll DC (2006) Insect detection of small targets moving in visual clutter. 492

PLoS Biol 4:378-386. 493

O'Carroll D (1993) Feature-detecting neurons in dragonflies. Nature 362:541-543. 494

Page 18: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

17

Olberg RM (1981) Object- and self-movement detectors in the ventral nerve cord of the dragonfly. J 495

Comp Physiol A 141:327-334. 496

Olberg RM (1986) Identified target-selective visual interneurons descending from the dragonfly brain. J 497

Comp Physiol A 159:827-840. 498

Olberg RM, Seaman RC, Coats MI, Henry AF (2007) Eye movements and target fixation during 499

dragonfly prey-interception flights. J Comp Physiol A 193:685-693. 500

Pal S (2015) Dynamics of aerial target pursuit. Eur Phys J Special Topics 224:3295-3309. 501

Paulk AC, Stacey JA, Pearson TW, Taylor GJ, Moore RJ, Srinivasan MV, van Swinderen B (2014) 502

Selective attention in the honeybee optic lobes precedes behavioral choices. Proc Natl Acad Sci U 503

S A 111:5006-5011. 504

Song Z, Juusola M (2014) Refractory sampling links efficiency and costs of sensory encoding to 505

stimulus statistics. J Neurosci 34:7216-7237. 506

Strausfeld NJ (1980) Male and female visual neurones in dipterous insects. Nature 283:381-383. 507

Suver MP, Huda A, Iwasaki N, Safarik S, Dickinson MH (2016) An array of descending visual 508

interneurons encoding self-motion in Drosophila. J Neurosci 36:11768. 509

Switzer PV, Eason PK (2000) Proximate constraints on intruder detection in the dragonfly Perithemis 510

tenera (Odonata: Libellulidae): Effects of angle of approach and background. Annals of the 511

Entomological Society of America 93:333-339. 512

Tolhurst D, Tadmor Y, Chao T (1992) Amplitude spectra of natural images. Ophthal Physiol Opt 513

12:229-232. 514

Trischler C, Kern R, Egelhaaf M (2010) Chasing behavior and optomotor following in free-flying male 515

blowflies: Flight performance and interactions of the underlying control systems. Front Behav 516

Neurosci 4:20. 517

Tsubouchi A, Yano T, Yokoyama TK, Murtin C, Otsuna H, Ito K (2017) Topological and modality-518

specific representation of somatosensory information in the fly brain. Science 358:615-623. 519

van Hateren JH (1992) Real and optimal neural images in early vision. Nature 360:68-70. 520

Wardill T, Fabian S, Pettigrew A, Stavenga DG, Nordström K, Gonzalez-Bellido P (2017) A novel 521

interception strategy in a miniature robber fly with extreme visual acuity. Curr Biol 27:854-859. 522

Weckström M, Laughlin SB (1995) Visual ecology and voltage-gated ion channels in insect 523

photoreceptors. Trends Neurosci 18:17-21. 524

Wellington W, Fitzpatrick S (1981) Territoriality in the drone fly, Eristalis tenax (Diptera, Syrphidae). 525

Can Entomol 113:695-704. 526

Wertz A, Haag J, Borst A (2009a) Local and global motion preferences in descending neurons of the 527

fly. J Comp Physiol A 195:1107-1120. 528

Page 19: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

18

Wertz A, Gaub B, Plett J, Haag J, Borst A (2009b) Robust coding of ego-motion in descending neurons 529

of the fly. J Neurosci 29:14993-15000. 530

Wiederman SD, O'Carroll DC (2011) Discrimination of features in natural scenes by a dragonfly neuron. 531

J Neurosci 31:7141-7144. 532

Wiederman SD, Shoemaker PA, O'Carroll DC (2008) A model for the detection of moving targets in 533

visual clutter inspired by insect physiology. PLoS ONE 3:e2784. 534

Yang J-B, Shi M, Yi Q-M (2012) A New Method for Motion Target Detection by Background Subtraction 535

and Update. Physics Procedia 33:1768-1775. 536

York CA, Bartol IK, Krueger PS (2016) Multiple sensory modalities used by squid in successful predator 537

evasion throughout ontogeny. J Exp Biol 219:2870-2879. 538

539

Page 20: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

19

Figure Legends 540

Figure 1. Eristalis and Holcocephala have dipteran target selective descending neurons (dTSDNs) 541

A. i) An example of a dTSDN response recorded in the Eristalis ventral nerve cord during stimulation 542

with a small target drifted horizontally across its receptive field. The inset diagram shows the mean 543

(thick line) and individual (thin lines) waveforms of 30 action potentials isolated from this example. ii) 544

The resulting spike train as a function of time. The inset graph shows the interspike intervals (ISI). iii) 545

Histogram of spike rate within 40 ms bins and mean response during target motion across the dTSDN 546

receptive field (bar under data). iv) The response to different target heights, where the target width was 547

fixed at 3 degrees. When the bar subtended more than 10 degrees of the visual field, the dTSDN 548

response was strongly suppressed. The data are normalized to each neurons’ maximum response. 549

N=27. B. i) Example of wide-field neuron response recorded from the ventral nerve cord in Eristalis 550

during stimulation with a background pattern drifting horizontally. The inset diagram shows the mean 551

(thick) and individual (thin lines) waveforms of 30 action potentials. ii) The resulting spike train as a 552

function of time, and the interspike interval (ISI, inset). iii) The spike rate in 40 ms bins and the mean 553

response during the peri-stimulus duration (bar under data). iv) The response to high-contrast 554

sinusoidal gratings moving in 8 different directions (wavelength 7 degrees, 5 Hz) of two different types 555

of wide-field sensitive neurons, here referred to as type 1 (dark purple, N=8) and type 2 (light purple, 556

N=14), and of dTSDNs (green, N = 27). C. i) An example response of a Holcocephala TSDN, and the 557

waveform (inset). ii) The resulting spike train and interspike interval (ISI). iii) The spike rate within 10 ms 558

bins and mean response for the stimulus duration (bar under data). iv) The response across neurons to 559

square targets of varying size (side length indicated on x-axis). The target appeared at a random 560

position on the screen, remained stationary for 150 ms and then moved in a random direction for 100 561

ms. N=12. The data are normalized to each neurons’ maximum response. In all panels the data are 562

displayed as mean sem. 563

Figure 2. Syn-directional background pattern motion strongly suppresses the dTSDN response to target 564

motion 565

A. The dTSDN response to a small target drifted across its receptive field. The target was presented 566

against a naturalistic background pattern moving in the same direction as the target. The green data 567

show the response of Eristalis TSDNs (N=8) and the red data the response of Holcocephala TSDNs 568

(N=38). The star (*) indicates p<0.05, one-way ANOVA. B. The response of Eristalis wide-field sensitive 569

neurons to the naturalistic background pattern moving at different velocities (N=8). Significant 570

differences between response and spontaneous rate are indicated (two-way ANOVA followed by 571

Page 21: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

20

Sidak’s multiple comparisons). C. The Eristalis TSDN response as a function of the response of the 572

wide-field sensitive neurons when using background patterns moving at different velocities. D. The 573

dTSDN response to a small target drifted across its receptive field. The target was presented against a 574

naturalistic background pattern moving at the same speed as the target, but in different directions. The 575

green data show the response of Eristalis TSDNs (N=5) and the red data the response of Holcocephala 576

TSDNs (N=38). Different letters above (Eristalis) or below (Holcocephala) the data points indicate that 577

they are significantly different from each other (one-way ANOVA followed by Tukey’s multiple 578

comparison’s test, tested separately for each species). E. The response of Eristalis wide-field sensitive 579

neurons to the naturalistic background pattern moving in different directions (N=8). Significant 580

differences between response and spontaneous rate are indicated (two-way ANOVA followed by 581

Sidak’s multiple comparisons). F. The response of the Eristalis TSDNs as a function of the response of 582

the wide-field sensitive neurons when the background pattern moved in different directions. In all 583

panels the data are displayed as mean sem. 584

Figure 3. A small amount of background pattern motion suppresses the dTSDN response to target 585

motion 586

A. The Eristalis TSDN response to a small target moving across a background pattern, both moving at 587

the same velocity. The background pattern was covered by a stationary mean luminance strip of 588

different heights, centered on the target’s trajectory. Different letters indicate that the data points are 589

significantly different from each other (one-way ANOVA followed by Tukey’s multiple comparison’s test, 590

N=7). B. The response of Eristalis wide-field sensitive neurons to the background pattern, covered by a 591

stationary mean luminance strip of different heights, centered on each neuron’s receptive field. N=7. 592

Significant differences between response and spontaneous rate are indicated (two-way ANOVA 593

followed by Sidak’s multiple comparisons). C. The response of the Eristalis TSDNs as a function of the 594

response of the wide-field sensitive neurons when the background pattern was covered by a gray strip. 595

D. The Eristalis TSDN response to a small target moving across a background pattern, at the same 596

velocity. The vertical extent of the background pattern was varied, and centered on the target’s 597

trajectory. Different letters indicate that the data points are significantly different from each other (one-598

way ANOVA followed by Tukey’s multiple comparison’s test, N=7). E. The response of Eristalis wide-599

field sensitive neurons to different background pattern heights. N=11. Significant differences between 600

response and spontaneous rate are indicated (two-way ANOVA followed by Sidak’s multiple 601

comparisons). F. The response of the Eristalis TSDNs as a function of the response of the wide-field 602

sensitive neurons when the vertical extent of the background pattern was varied. In all panels the data 603

are displayed as mean sem. 604

Page 22: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

21

Figure 4. Target motion is least visible when the background pattern is most naturalistic 605

A. The Eristalis TSDN response to a small target moving across a background pattern, at the same 606

velocity. We varied the contrast of the background pattern (x-axis), and the slope constant of its 607

amplitude spectrum (alpha, as color coded). A two-way ANOVA showed a significant effect of alpha 608

(p=0.0046), contrast (p<0.0001) and a significant interaction between contrast and alpha (p=0.0002). 609

N=5. B. The response of Eristalis wide-field sensitive neurons to background patterns with varied 610

contrast (x-axis) and slope constant of its amplitude spectrum (alpha, as color coded). A two-way 611

ANOVA showed a significant effect of alpha (p<0.0001), contrast (p<0.0001) and a significant 612

interaction between contrast and alpha (p<0.0001). N=7. C. The response of the Eristalis TSDNs as a 613

function of the response of the wide-field sensitive neurons when we varied background pattern 614

contrast and slope constant. In all panels the data are displayed as mean sem. 615

Page 23: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their
Page 24: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their
Page 25: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their
Page 26: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their
Page 27: Integration of Small- and Wide-Field Visual Features in Target … · 2018-10-29 · Bats find prey with echolocation (Falk et al., 2014) , squid embryos avoid predators with their

1

Table 1. Fit parameters for the data shown in Fig. 2C, F, 3C, F, 4C.

For each data set we show two fit parameters, which can be used to determine how well the data fit a particular function. The coefficient of determination (R2) quantifies the goodness of the fit on a scale from 0 to 1 (perfect), and the standard deviation of the residuals (Sy.x) is similar to the root mean square error (RMSE) but the number of parameters fit by the regression are taken into account. In all conditions, we used a least squares, unconstrained fit, with a maximum of 1000 iterations.

f(x)=k+c*x f(x)=k+c/x f(x)=k+c/x2 f(x)=k*e-c*x Background

variable Coefficient of determination

(R2)

Standard deviation

of the residuals

(Sy.x)

R2 Sy.x R2 Sy.x R2 Sy.x

Velocity (Fig. 2C)

0.8411 10.46 0.7888 12.06 0.6426 15.69 0.9347 7.244

Direction (Fig. 2F)

0.7879 9.062 0.8469 7.697 0.7351 10.13 0.8859 7.281

Cover (Fig. 3C)

0.9259 1.115 0.888 1.37 0.8281 1.698 0.936 1.196

Height (Fig. 3F)

0.7753 10.44 0.9247 6.044 0.9358 5.581 0.9306 6.701

Contrast/alpha (Fig. 4C)

0.9132 8.899 0.8197 12.2 0.6151 18.73 0.9762 5.027