Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After...

29
1 Type-specific dendritic integration in mouse retinal ganglion cells 1 Yanli Ran 1,2* , Ziwei Huang 1,2* , Tom Baden 1,4 , Harald Baayen 5 , Philipp Berens 1-3,6,# , Katrin Franke 1,3,# , 2 Thomas Euler 1-3,#,@ 3 1 Institute for Ophthalmic Research, University of Tübingen, Tübingen, Germany 4 2 Centre for Integrative Neuroscience, University of Tübingen, Tübingen, Germany 5 3 Bernstein Centre for Computational Neuroscience, University of Tübingen, Tübingen, Germany 6 4 Sussex Neuroscience, School of Life Sciences, University of Sussex, Brighton, UK 7 5 Departement of Linguistics, University of Tübingen, Tübingen, Germany 8 6 Institute of Bioinformatics and Medical Informatics, University of Tübingen, Tübingen, Germany 9 10 11 12 * These authors contributed equally to this work 13 # Senior authors 14 @ Correspondence: [email protected] 15 . CC-BY-NC-ND 4.0 International license available under a was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made The copyright holder for this preprint (which this version posted September 3, 2019. ; https://doi.org/10.1101/753335 doi: bioRxiv preprint

Transcript of Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After...

Page 1: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

1

Type-specific dendritic integration in mouse retinal ganglion cells 1

Yanli Ran1,2*, Ziwei Huang1,2*, Tom Baden1,4, Harald Baayen5, Philipp Berens1-3,6,#, Katrin Franke1,3,#, 2

Thomas Euler1-3,#,@ 3

1Institute for Ophthalmic Research, University of Tübingen, Tübingen, Germany 4 2Centre for Integrative Neuroscience, University of Tübingen, Tübingen, Germany 5 3Bernstein Centre for Computational Neuroscience, University of Tübingen, Tübingen, Germany 6 4Sussex Neuroscience, School of Life Sciences, University of Sussex, Brighton, UK 7 5Departement of Linguistics, University of Tübingen, Tübingen, Germany 8 6Institute of Bioinformatics and Medical Informatics, University of Tübingen, Tübingen, Germany 9

10

11

12 *These authors contributed equally to this work 13 #Senior authors 14 @Correspondence: [email protected] 15

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 2: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

2

ABSTRACT 16

Neural computation relies on the integration of synaptic inputs across a neuron’s dendritic arbour. 17

However, the fundamental rules that govern dendritic integration are far from understood. In 18

particular, it is still unclear how cell type-specific differences in dendritic integration arise from general 19

features of neural morphology and membrane properties. Here, retinal ganglion cells (RGCs), which 20

relay the visual system’s first computations to the brain, represent an exquisite model. They are 21

functionally and morphologically diverse yet defined, and they allow studying dendritic integration in 22

a functionally relevant context. Here, we show how four morphologically distinct types of mouse RGC 23

with shared excitatory synaptic input (transient Off alpha, transient Off mini, sustained Off, and F-24

miniOff) exhibit distinct dendritic integration rules. Using two-photon imaging of dendritic calcium 25

signals and biophysical modelling, we demonstrate that these RGC types strongly differ in their spatio-26

temporal dendritic integration: In transient Off alpha cells, dendritic receptive fields displayed little 27

spatial overlap, indicative of a dendritic arbour that is partitioned in largely isolated regions. In contrast, 28

dendritic receptive fields in the other three RGCs overlapped greatly and were offset to the soma, 29

suggesting strong synchronization of dendritic signals likely due to backpropagation of somatic signals. 30

Also temporal correlation of dendritic signals varied extensively among these types, with transient Off 31

mini cells displaying the highest correlation across their dendritic arbour. Modelling suggests that 32

morphology alone cannot explain these differences in dendritic integration, but instead specific 33

combinations of dendritic morphology and ion channel densities are required. Together, our results 34

reveal how neurons exhibit distinct dendritic integration profiles tuned towards their type-specific 35

computations in their circuits, with the interplay between morphology and ion channel complement 36

as a key contributor. 37

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 3: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

3

INTRODUCTION 38

Across the nervous system, the output signal of a neuron is determined by how it integrates the often 39

thousands of synaptic inputs it receives across its dendritic arbour (Guo et al., 2014; Koch et al., 1982). 40

However, still little is known about how dendritic integration is shaped by differences between neuron 41

types, such as specific dendritic morphology and ion channel complement and density. To investigate 42

type-specific dendritic integration and the key factors driving it, we here use the vertebrate retina, a 43

model system with a clear input-output relationship that can be recorded in a dish (Ames and Nesbett, 44

1981). The retina decomposes the visual signal into ~40 feature-specific parallel channels (reviewed in 45

(Baden et al., 2018)), relayed to the brain by a matching number of retinal ganglion cell types (Baden 46

et al., 2016; Sanes and Masland, 2015). Retinal ganglion cells (RGCs) receive their main excitatory drive 47

from the bipolar cells (BCs), which pick up the photoreceptor signal in the outer retina. In addition, 48

RGCs (and BCs) receive inhibitory input from amacrine cells (ACs) (reviewed in (Diamond, 2017)), 49

completing the canonical RGC input circuit. Different RGC types differ in morphology (Bae et al., 2018; 50

Helmstaedter et al., 2013; Sümbül et al., 2014), synaptic connectivity (Field et al., 2010; Helmstaedter 51

et al., 2013), and expression of ion channels (Rheaume et al., 2018; Siegert et al., 2012). 52

To explain the emergence of diverse RGC functions, many previous studies have focused on the 53

selective connectivity with presynaptic neurons in the inner plexiform layer (IPL) (e.g. (Helmstaedter 54

et al., 2013; Lee et al., 2010; Yu et al., 2018)). Different RGC types arborize in specific layers of the IPL 55

and, hence, receive synaptic inputs from distinct combinations of BC and AC types (Helmstaedter et 56

al., 2013). This spatiotemporally heterogeneous input provides the basis of type-specific feature 57

extraction (Roska and Werblin, 2001). In addition, RGC dendrites may themselves perform complex 58

computations and therefore contribute to the generation of specific output channels, e.g. through 59

their dendritic geometry, and the complement, distribution and density of passive and active ion 60

channels (Branco and Hausser, 2010; Goldstein and Rall, 1974; Koch et al., 1982; Lai and Jan, 2006; 61

Stuart and Spruston, 2015). So far, dendritic processing in the retina has been studied mainly in 62

interneurons (e.g. (Chapot et al., 2017; Grimes et al., 2010; Hausselt et al., 2007; Koren et al., 2017)). 63

Despite some theoretical work in this direction (reviewed in (Guo et al., 2014)), experimental evidence 64

for type-specific dendritic computation and their biophysical mechanisms in RGCs remains limited and 65

is restricted to a few specific types (i.e. direction-selective RGCs, (Dorostkar et al., 2010; Oesch et al., 66

2005; Sivyer and Williams, 2013); On alpha RGCs, (Schwartz et al., 2012)). 67

Here, we exploit the unique structure of the IPL to isolate the contributions of type-specific synaptic 68

input profiles from intrinsic cellular mechanisms to elucidate whether RGC types sampling from a 69

similar input space use specific dendritic integration profiles to generate functionally diverse outputs. 70

To this end, we studied in mouse retina the dendritic integration properties of four Off RGC types that 71

receive excitatory input from a highly overlapping set of presynaptic neurons. To record light stimulus-72

evoked signals across the dendritic arbour of individual RGCs, we used two-photon Ca2+ imaging. We 73

found that these morphologically diverse RGC types differed strongly in their spatio-temporal dendritic 74

integration properties. A biophysical model suggests that the differential dendritic integration in these 75

RGC types arises from the type-specific combination of dendritic morphology and ion channel 76

complement. 77

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 4: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

4

RESULTS 78

Estimating local dendritic receptive fields in single retinal ganglion cells 79

To study dendritic integration in different RGC types, we recorded Ca2+ signals in response to visual 80

stimulation across the dendritic arbour of individual cells in the ex-vivo, whole-mounted mouse retina 81

using two-photon imaging. For that, we injected individual RGCs with the fluorescent Ca2+ indicator 82

dye Oregon Green BAPTA-1 (OGB-1) using sharp electrodes (Methods), resulting in completely labelled 83

individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 84

allowing us to extract morphological parameters such as dendritic arbour area, branching order and 85

asymmetry. To determine the cell’s dendritic stratification profile across the IPL relative to the ChAT 86

bands, blood vessels labelled with Sulforhodamine 101 (SR101) were used as landmarks (Fig. 1a, b; 87

Methods). 88

89

Figure 1 | Recording dendritic receptive fields (RFs) in individual retinal ganglion cells (RGCs). a, Z-projection 90 of an image stack showing an Off RGC filled with the synthetic Ca2+ indicator Oregon green BAPTA-1 (OGB-1; 91 green) and blood vessels (red) in top view (top) and as side view (bottom). Dashed white lines mark blood vessels 92

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 5: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

5

at the borders to ganglion cell layer (GCL) and inner nuclear layer (INL). b, Reconstructed morphology of cell from 93 (a). Dashed grey lines between vessel plexi indicate ChAT bands. c, Example scan fields, as indicated by blue 94 rectangles in (b), with exemplary region of interest (ROI; white) each. d, De-trended Ca2+ signals from ROIs in (c) 95 during dense noise stimulation (20x15 pixels, 30 µm/pixel, 5 Hz). e, Spatial receptive field (RF) maps from event-96 triggered averaging for left ROI in (c) at different times (δ, [s]) before an event and singular value decomposition 97 (svd; Methods) map (left). Up-sampled and smoothed RF map overlaid with the cell’s morphology (right; red 98 crosshair indicates soma position), ROI position (blue dot) and RF contour. f, Like (e) but for right ROI in (c). g, RF 99 contours of ROIs from (e, f) overlaid on the reconstructed cell morphology. h, Top- and side-view of example cell 100 with all analysed ROIs (n=15 scan fields, n=193 of 232 ROIs passed the quality test; see Methods and Suppl. Fig. 101 S1a, b), colour-coded by dendritic distance from soma. i, RF contours of ROIs from (h). Scale bars: a,b,e-i, 50 µm, 102 c, 10 µm. 103

To map dendritic receptive fields (RFs) of RGCs (Fig. 1c, d), we used a binary dense noise stimulus 104

(20x15 pixels, 30 µm/pixel) that was centred on the recording field. For each recording field (32x16 105

pixels @31.25 Hz), we extracted regions-of-interest (ROIs) along the dendrites using local image 106

correlations (Supplementary Fig. S1a; Methods). Next, we registered the position and distance of each 107

dendritic segment relative to the soma and extracted each ROI’s Ca2+ signal. To mitigate the effect of 108

low signal-to-noise ratio in some dendritic recordings, we used routinely automatic smoothness 109

determination (Methods; (Sahani and Linden, 2003)) to obtain reliable estimates of each ROI´s RF. 110

Then, we overlaid the RF contours with the cell´s morphology (Fig. 1e-g). For each cell, we recorded 111

different dendritic regions at various distances from the soma yielding between 40 and 232 ROIs per 112

cell (Fig. 1h, i; Supplementary Fig. S1c). This enabled us to systematically probe dendritic integration 113

across an RGC’s dendritic arbour and link the properties of local dendritic RFs to overall cell 114

morphology. 115

Recorded retinal ganglion cells are clustered into four morphological types 116

To compare dendritic integration profiles across RGC types with highly overlapping excitatory inputs, 117

we focussed on Off RGCs that stratify close to the Off ChAT band (Fig. 2a; Supplementary Fig. S2). We 118

recorded n=31 cells and clustered them into four morphological groups, using four morphological 119

criteria: soma size, arbour asymmetry, arbour density difference and arbour area following Bae et al. 120

(2018) (Fig. 2; Methods). One group likely corresponded to transient Off alpha (tOff alpha) RGCs, as 121

indicated by a large soma and dendritic area (for statistics, see Table 1) and their characteristic 122

stratification profile (compare to “4 ow” RGCs in the EyeWire database of reconstructed cells of the 123

mouse retina, http://museum.eyewire.org). The second group likely represented the Off “mini” alpha 124

transient type (tOff mini; (Baden et al., 2016)): Cells assigned to this group exhibited an IPL 125

stratification profile very similar to tOff alpha cells, but had smaller somata and dendritic areas. The 126

third group resembled the morphology of F-miniOff cells (Rousso et al., 2016), exhibiting an IPL 127

stratification profile peaking between the Off ChAT band and the outer IPL border and a small, highly 128

asymmetrical dendritic arbour. Finally, the forth group displayed a similar IPL stratification profile as 129

sustained Off alpha RGCs (“1wt” cells in (Bae et al., 2018)), but had smaller somata and arbour areas. 130

These cells may correspond to the Off sustained (G7) RGCs identified by Baden et al (2016). Here, we 131

refer to them as sustained Off (sOff). In the following, for simplicity, we will refer to these 132

morphological groups as RGC types. 133

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 6: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

6

134

Figure 2 | Anatomical clustering of recorded RGCs. a, Top- and side-views of four reconstructed Off RGCs, one 135 of each studied type, with IPL stratification profiles as mean (black) and for all recorded cells of that type (grey). 136 Dashed lines indicate On and Off ChAT bands. b, Cluster-dendrogram with the morphological parameters used 137 in each clustering step and the resulting RGC groups: n=17 tOff alpha, n=5 tOff mini, n=4 sOff and n=5 F-miniOff. 138 Colours indicate cluster (RGC type), dot diameter represents soma area. Inset: Illustration of arbour density 139 difference measure. Scale bar: a, 50 µm. 140

Dendritic integration profiles vary across retinal ganglion cell types 141

Dendrites can process incoming synaptic inputs on a local and a global scale, resulting in rather 142

compartmentalized and synchronized dendrites, respectively (Behabadi and Mel, 2014; Bernander et 143

al., 1991; Koch et al., 1982; Magee, 2000; Polsky et al., 2004; Williams, 2004). To investigate whether 144

the four Off RGC types differ with respect to their integration mode, we first assessed how the RF size 145

changed as a function of dendritic distance to the soma. In tOff alpha cells, local RF area systematically 146

decreased as a function of ROI distance from the soma (Fig. 3a-c; cf. Fig. 1g, i), suggesting that signals 147

in distal dendrites of tOff alpha cells are more isolated and local than those in proximal dendrites. This 148

was not the case in the three remaining RGC types, where RF size remained relatively constant across 149

different positions of the dendritic arbour (Fig. 3a-c). In fact, proximal RFs were significantly larger in 150

tOff alpha cells than in the other RGC types (Fig. 3d; for statistical analysis, see Supplementary 151

Information), which did not differ significantly in their RF size along their dendrite. Together, among 152

the recorded RGC types, dendritic signals in tOff alpha cells are the least spatially synchronized, 153

suggesting that they process dendritic input more locally than the other types. 154

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 7: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

7

155

Figure 3 | Local dendritic RF area and position varies in different RGC types. a, Top-views of the different 156 reconstructed RGC types, overlaid with ROIs that passed the RF quality test. ROI colours indicate dendritic 157 distance to soma. b, Cells from (a) with three ROIs of increasing distance from soma and corresponding RF 158 contours overlaid (red “x” indicates soma position). c, Dendritic RF area as a function of dendritic distance to 159 soma; data pooled across cells of the same type (see below). d, Comparison of RF area change with dendritic 160 distance to soma between each pair of RGC type. Red shaded areas indicate dendritic sections with significant 161 differences between types (Methods). e, Cells from (a) with arrows indicating spatial offset between ROI centre 162 and the RF contour’s geometrical centre, with arrows pointing at the latter. f, RF offset of all recorded ROIs as a 163 function of dendritic distance to soma. g, Like (d) but for RF offset changes. Data from tOff alpha (n=17\1,452\850 164 cells\total ROIs\ROIs passing the quality test), tOff mini (n=5\387\295), sOff (n=4\208\154) and F-miniOff RGCs 165 (n=5\265\126); for individual cell morphologies, see Suppl. Fig. S2. Scale bar: a, 50 µm. For statistical analysis, 166 see Suppl. Information. 167

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 8: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

8

168

Figure 4 | Dendritic RF overlap. a, Top-views of the different reconstructed RGC types with RF contours of three 169 ROIs overlaid. b, Illustration of dendritic and angular distance (Ѳ) between two ROIs (measured from the last 170 common branching node) and RF overlap (grey area) of two RF contours (ellipses). c, Hexagon maps showing the 171 dendritic RF overlap index (colour-coded) as a function of Ѳ and dendritic distance for all ROI pairs: tOff alpha 172 (n=17\40,777 cells\ROI pairs), tOff mini (n=5\13,524), sOff (n=4\3,141), and F-miniOff (n=5\2,097). d, 2D 173 comparison maps for plot area marked by the dashed red rectangles in (c) for each pair of RGC types. Colour 174 codes difference in RF overlap index, with whitened areas indicating no significant difference. For statistical 175 analysis, see Suppl. Information. 176

Synchronization of dendrites can originate from strong backpropagation of somatic spikes to the 177

dendrites (reviewed in (Stuart and Spruston, 2015)). This is not only expected to increase dendritic RF 178

size but should also shift the RF´s centre closer towards the soma. In contrast, for a more isolated 179

dendrite without backpropagation, the RF centre should roughly correspond to the respective ROI 180

position. Therefore, we next measured ROI-to-RF-centre offsets for the four RGC types (Fig. 3e, f). We 181

found that tOff alpha cells displayed small offsets that did not change much as a function of dendritic 182

distance from the soma. In contrast, the other three RGC types displayed large offsets, with the RF 183

centre strongly shifted towards the centre of the dendritic arbour, which in tOff mini and sOff cells also 184

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 9: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

9

coincided with the soma (Fig. 3e). Moreover, in tOff mini and sOff cells, offsets increased with dendritic 185

distance from the soma (Fig. 3f). In F-miniOff cells, due to their asymmetrical dendritic arbours, offsets 186

increased with dendritic distance from the arbour centre (Fig. 3f), resulting in an inverted-bell shaped 187

curve. For large dendritic distances, the offsets were significantly different between all pairs of RGC 188

types (Fig. 3g). These results confirm that dendrites of tOff mini, sOff and F-miniOff cells are more 189

synchronized than those of tOff alpha cells, possibly due to backpropagation. 190

Strongly isolated dendrites, as observed in tOff alpha cells, could allow dendritic computations at a 191

finer spatial scale than the whole cell’s RF. Such isolated dendrites are expected to be spatially more 192

independent than the better synchronized dendrites of tOff mini, sOff and F-miniOff cells. To test this 193

prediction, we determined the overlap of RFs for every ROI pair recorded in a single cell (Fig. 4a, b). 194

We then assessed how the overlap changed with dendritic distance and angle between ROIs (Fig. 4b, c; 195

Supplementary Fig. S3). We found localized and spatially independent RFs only in tOff alpha RGCs (Fig. 196

4a, c). Here, RF overlap decreased substantially with increasing dendritic and angular distance between 197

ROIs, in line with our previous results. In tOff mini cells, RFs showed partial overlap even when the 198

ROIs were located at opposite sides of the dendritic arbour (Fig. 4a, c). For sOff and F-miniOff cells, RFs 199

overlapped substantially, independent of dendritic and angular distance between ROIs. As a result, the 200

RF overlap maps significantly differed between tOff alpha and the other three RGC types, and partially 201

between tOff mini and the remaining two RGC types (Fig. 4d), supporting significant differences in 202

dendritic processing – from more local in tOff alpha to more global in F-miniOff. 203

Together, these results suggest that different RGC types with similar input profiles apply vastly 204

different dendritic integration rules. For example, the dendrites of tOff alpha cells seem to exhibit little 205

backpropagation but reasonably strong forward propagation, integrating RFs from all dendrites 206

symmetrically. This leads to larger proximal than distal RFs and distal RFs with little overlap and 207

displacement. In contrast, the other three RGC types show strong indication for backpropagation 208

across their dendritic arbour, causing distal RFs to be highly overlapping and displaced towards the 209

centre of the dendritic arbour. 210

Temporal dendritic integration varies between RGC types 211

Dendritic inputs are not only integrated across space, but also over time. To relate spatial to temporal 212

dendritic integration, we next probed the temporal synchronisation of light responses across the 213

dendritic arbour of the four RGC types. For that, we used a “chirp” stimulus that consisted of a light 214

step followed by frequency and contrast modulations (Methods) and was presented as local (100 µm 215

in diameters) and full-field (500 µm) version. Notably, F-miniOff RGCs did not show any reliable dendritic 216

chirp responses, despite the same ROIs passing our RF quality threshold (Methods). This finding is 217

consistent with earlier observations in this RGC type (cf. “x2 cell” of Extended Data Fig. 5 in (Baden et 218

al., 2016)). Therefore, we focussed the following analysis on the remaining three RGC types. 219

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 10: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

10

220 Figure 5 | Temporal correlation across dendrites. a, Exemplary response of a tOff alpha RGC to local (middle) 221 and full-field chirp (right) recorded from three ROIs indicated on the reconstructed cell (left). Values next to the 222 traces indicate linear correlation coefficient of the corresponding trace pair. b,c, Like (a), but for tOff mini (b) and 223

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 11: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

11

sOff RGC (c). d, Hexagon maps showing response correlations for local (left) and full-field chirp (right) as a 224 function of angular distance and dendritic distance between ROIs for tOff alpha (n=17\12,770\13,001 cells\pairs 225 for full-field\pairs for local), tOff mini (n=5\6,529\6,529) and sOff RGCs (n=4\2,622\2,557). Colour encodes 226 correlation. e, 2D comparison maps for inter-ROI correlation of local and full-field chirp responses for the plot 227 area marked by dashed black rectangle in (d) for each RGC type. Colour codes difference in correlation, with 228 whitened areas indicating no significant difference. f,g, Like (e), but for the comparison between cell types for 229 local chirp responses (f) and full-field chirp responses (g). For statistical analysis, see Suppl. Information. 230

We found that dendritic responses to the local chirp in tOff alpha and tOff mini RGCs were quite similar 231

but differed from those in sOff RGCs (Fig. 5a-c). In the latter, local chirp responses were more sustained 232

than those in the other two types (Supplemental Fig. S4c); this difference resonates with sOff cells 233

stratifying slightly more distally (cf. Fig. 2a) and, hence, presumably receiving more input from 234

sustained BC types (Franke et al., 2017; Roska and Werblin, 2001; Yu et al., 2018). When presented 235

with the full-field chirp, tOff alpha and tOff mini RGC responses became somewhat more distinct (i.e. 236

to the frequency modulation). This difference was not found in an earlier study (Baden et al., 2016) 237

but may be related to the fact that in the present study, light stimuli could be precisely centred on the 238

recorded cell. In addition, all three RGC types often showed On events that were much less frequent 239

for the local chirp (Fig. 5a-c; Supplemental Fig. S4b, d). Similar On-events in Off-cells have also been 240

observed in BC responses (Franke et al., 2017). In general, differences between full-field and local chirp 241

responses were more pronounced in sOff RGCs (Supplemental Fig. S4e), suggesting that stimulus size 242

had a larger effect on sOff cells compared to the other two types. This could be due to a stronger 243

inhibitory surround or connections to BCs that are more strongly influenced by surround stimulation 244

(Franke et al., 2017). 245

To analyse the temporal properties of dendritic integration in these cells, we quantified the correlation 246

of local or full-field chirp responses between ROI pairs across the dendritic arbour (Fig. 5d). In all three 247

RGC types, correlations between ROIs were higher for responses to local than to full-field chirps (Fig. 248

5d, e). The decorrelation observed for full-field chirps was especially pronounced in sOff cells (Fig. 5e). 249

In tOff alpha and sOff RGCs, correlation decreased with dendritic and angular distance (Fig. 5d). While 250

in tOff alpha cells, lower correlation coincided with smaller and less overlapping RFs, sOff RGCs 251

displayed low correlation in their distal dendrites while at the same time featuring highly overlapping 252

RFs across the whole dendritic arbour. In contrast to the other two RGCs, temporal correlation in tOff 253

mini cells was largely independent of dendritic and angular distance (Fig. 5d). In addition, correlation 254

was overall much higher, indicating that dendritic segments in tOff mini cells are temporally more 255

synchronized (cf. Fig. 5c). Because differences in temporal correlation between RGC types persisted 256

when applying a more stringent quality criterion (Supplemental Fig. S6; Methods), it is unlikely that 257

they were due to systematic differences in recording quality (i.e. signal-to-noise-ratio). 258

Taken together, our data suggest that spatio-temporal integration is tuned across the RGC dendritic 259

arbour in a highly type-specific manner (Fig. 5f, g). The studied RGC types ranged between two main 260

dendritic integration profiles: The first profile featured strongly isolated dendrites (e.g. in tOff alpha) 261

and may render the cell sensitive to fine visual stimulus structures within the cell’s RF. In contrast, the 262

second profile featured strongly synchronised dendrites with highly overlapping RFs (e.g. in tOff mini 263

RGCs) and may tune the cell towards robustly detecting a stimulus independent of its location within 264

the RF. 265

Simulation reveals mechanistic principles for type-specific dendritic integration 266

The dendritic integration properties of RGC types may be influenced by morphological features, such 267

as branching pattern, dendritic thickness and segment length, and the complement and distribution of 268

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 12: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

12

ion channels (Koch et al., 1982; Rall and Rinzel, 1973). To understand which of these properties may 269

explain the dendritic integration profiles we observed, we built a simple, morphology-inspired 270

biophysical model and focussed on the effects of the type-specific morphology and dendritic channel 271

densities in tOff alpha and tOff mini cells (Fig. 6a). 272

273

Figure 6 | Simulation of dendritic signal propagation in tOff alpha and tOff mini RGCs. a, Reconstructed cell 274 morphologies of tOff alpha and tOff mini RGC (same cells as in Figs. 3,4). b, Dendrite radius (top) and segment 275 length as functions of branch order (data from http://museum.eyewire.org; n=2 for tOff alpha (4ow); n=3 for 276 tOff mini (4i)). c, Illustration of the ball-and-stick models used for simulations in (d, e). Simulated inputs at 277 proximal (25 µm to soma) and distal (85% of the total dendrite length to soma) positions indicated as red and 278 black Gaussians, respectively. Respective read-out positions for (d) are indicated below the dendrite. The 279 thickness change of the bars (bottom) corresponds to the decay of forward (red) and backward (black) signal 280 propagation expected from our experimental data. d, Heat maps showing the signal amplitude at the two read-281 out positions indicated in (c), normalized to the amplitude at the respective input position as a function of ion 282 channel density combinations. White boxes indicate channel combinations that are consistent with our 283 experimental results. e, Normalized signal amplitude at read-out positions along the dendrite as a function of 284 dendritic distance for the channel combinations indicated by boxes in (d). Generic voltage-gated (𝐶𝑎, 𝐾, 𝑁𝑎) 285 and Ca2+ activated (𝐾,𝐶𝑎 ) conductances were modeled after Fohlmeister and Miller (1997). For details, see 286

Methods. 287

To capture the morphological differences between the cell types, we first extracted morphological 288

parameters from a published EM dataset (Bae et al., 2018). We found that for tOff alpha cells, dendritic 289

radius decreased systematically with increasing branch order; this decrease was less pronounced in 290

tOff mini cells (Fig. 6b). In addition, dendritic segment length increased with branch order for tOff alpha 291

cells, while it remained constant for tOff mini cells. Based on these differences, we built a ball-and-292

stick model for each cell type (Fig. 6c). For our simulations, we provided either a proximal or distal 293

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 13: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

13

input, with readout positions at the dendritic tip and close to the soma (Methods). Based on the 294

dendritic integration profiles of tOff alpha and tOff mini cells (cf. Figs. 3-5), we hypothesize that (i) in 295

tOff alpha cells, forward propagation (from distal to proximal dendrites) should be stronger than 296

backward propagation and (ii) that backpropagation should be strong in tOff mini cells (Fig. 6c). 297

To investigate the role of ion channel distribution on dendritic signal propagation, we systematically 298

varied the dendritic density of Ca2+-activated K+ channels (𝐾,𝐶𝑎) and voltage-gated K+, Na+ and Ca2+ 299

channels (𝐾 , 𝑁𝑎 and 𝐶𝑎). Notably, the same combination of channel densities had quite different 300

effects when applied to the two RGC morphologies (compare columns in Fig. 6d and Supplemental Fig. 301

7), highlighting how strongly the interplay between morphology and channel complement affects a 302

cell’s dendritic signal propagation. Notably, we found that distinct, cell-type specific sets of ion channel 303

densities were compatible with the experimentally derived hypotheses (Fig. 6d): For the tOff alpha cell 304

model, intermediate 𝐾,𝐶𝑎 and high 𝑁𝑎 and 𝐶𝑎 channel densities were required to generate 305

stronger forward propagation compared to backward propagation (Fig. 6d, e). For the same channel 306

densities, forward propagation in modelled tOff mini cell was so low that distal inputs were almost 307

completely extinguished before reaching the proximal dendrite. In contrast, with higher 𝐾 and lower 308

𝑁𝑎 densities, tOff mini cells showed strong backward and substantial forward propagation, in line 309

with our hypothesis (Fig. 6d, e). 310

Together, these results suggest that morphology alone does not explain the experimentally observed 311

differences between the two cell types. Instead, our model indicates that differences in dendritic 312

channel densities may be responsible for the distinct dendritic integration profiles in RGCs. 313

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 14: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

14

DISCUSSION 314

Here, we studied dendritic integration in four types of mouse Off RGC (tOff alpha, tOff mini, sOff, and 315

F-miniOff), which have their dendrites in overlapping strata of the IPL and, hence, receive highly 316

overlapping sets of synaptic input. Recordings of local, light-evoked dendritic activity and compart-317

mental modelling revealed surprising differences between the cells’ spatio-temporal dendritic inte-318

gration. What could these distinct integration rules be good for in terms of visual computations? 319

In tOff alpha RGCs (Deny et al., 2017; Krieger et al., 2017; Pang et al., 2003; Van Wyk et al., 2009), as 320

the distance from the soma increased, RF area decreased and dendritic RFs became increasingly non-321

overlapping, with minimal offset between recording site and respective RF centre. In addition, activity 322

on different dendritic branches was only moderately correlated. The more isolated, independent 323

dendritic segments in tOff alpha cells may help them to detect fine structures of visual stimuli and 324

support visual computations relying on spatial resolution below the RF of the entire cell. This is 325

reminiscent of what has been reported about On alpha cells, which possess nonlinear RFs and respond 326

to patterns that contain local structures finer than the cell’s RF centre (Schwartz et al., 2012). In 327

contrast, in tOff mini and sOff RGCs (Baden et al., 2016), RFs overlapped extensively and changed little 328

in area, while their centres were systematically shifted towards the soma. In addition, the timing of 329

responses was highly correlated across tOff mini dendrites, suggesting they may reliably detect stimuli 330

independent of their location within the RF. For sOff RGCs, the temporal correlation between the 331

activity of different dendritic branches decreased strongly for larger stimuli, suggesting that the cell’s 332

computational properties change for extended visual stimuli, likely due to lateral amacrine cell circuits 333

kicking in. F-miniOff cells (Rousso et al., 2016) were similar to tOff mini and sOff RGCs with some 334

particularities related to the high asymmetry of their dendritic arbour. Our morphologically inspired 335

biophysical model revealed that morphological difference alone cannot explain these experimentally 336

observed dendritic integration profiles; instead, distinct combinations of morphology, ion channel 337

complements and densities are required. 338

Dendritic integration rules have been studied extensively in the cortex (e.g. (Schmidt-Hieber et al., 339

2017; Tran-Van-Minh et al., 2016; Vetter et al., 2001)). In the retina, mainly interneurons have been at 340

the centre of interest: For example, it has been suggested that horizontal cells (Chapot et al., 2017) 341

and A17 amacrine cells (Grimes et al., 2010) provide locally computed feedback by confining signals 342

within single varicosities. Likewise, starburst amacrine cell dendrites compute the direction of motion 343

“dendrite-wise” by dividing their dendritic arbour into isolated sectors which contain 15-20 varicosities 344

each (Euler et al., 2002; Koren et al., 2017; Poleg-Polsky et al., 2018). In RGCs, dendritic integration has 345

been studied in direction-selective (DS) RGCs, where intrinsic properties of their dendritic arbour 346

(Schachter et al., 2010; Sivyer and Williams, 2013) and partially their asymmetry (Trenholm et al., 2011) 347

contribute to the generation of DS output. Reminiscent of our findings in tOff alpha cell, the dendritic 348

arbour of DS RGCs is functionally partitioned, with the “DS mechanism” replicated across the dendritic 349

arbour, such that local motion within the cell’s RF can cause a robust spiking response (Barlow and 350

Levick, 1965; Oesch et al., 2005). 351

We chose to focus on four types of Off RGCs because they are expected to receive similar excitatory 352

synaptic inputs. Nevertheless, due to small differences in dendritic stratification depth, they make 353

connections with partially different sets of BCs: tOff alpha cells contact mainly transient type 3a and 4 354

BCs, while sOff cells likely contact mainly the more sustained type 1 and 2 BCs (Bae et al., 2018; 355

Helmstaedter et al., 2013; Yu et al., 2018). In line with this, we found that the dendrites of tOff alpha 356

cells exhibited more transient responses than those from sOff cells. Since tOff mini RGCs co-stratify 357

with tOff alpha RGCs, they likely receive excitatory inputs from the same BC types and thus should 358

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 15: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

15

exhibit similar response properties. Indeed, these two RGC types showed similar responses to local 359

chirps. Nevertheless, they may be differentially modulated by type-specific connectivity to ACs. In line 360

with this, the two cell types showed more distinct responses to full-field chirps. In principle, the 361

interaction of excitation from BCs and inhibition from ACs may attenuate the excitatory inputs and 362

affect dendritic integration (Roska et al., 2006), raising the possibility that the observed type-specific 363

differences could at least partially result from type-specific microcircuit connectivity rather than mainly 364

from cell intrinsic properties as suggested above. While we cannot exclude this possibility, our 365

simulation results suggest a more parsimonious explanation for the observed differences in these 366

dendritic integration profiles, relying on cell-intrinsic mechanisms only. 367

Apart from contributions of the microcircuit connectivity, dendritic integration is mainly determined 368

by a combination of morphological features and passive and active membrane properties, which can 369

differ significantly between RGC types (reviewed in (Stuart and Spruston, 2015)). In some RGC types 370

like the tOff alpha, for instance, the dendritic diameter becomes smaller and dendritic segment length 371

gets longer with increasing branch order. This, in turn, results in a higher axial resistance and shorter 372

propagation distance for more distal dendritic signals. In other RGC types like tOff mini, however, 373

dendritic diameter and segment length does not systematically change with increasing branch order. 374

In addition, a variety of ion channels, including Ca2+-activated K+ channels, hyperpolarization-activated 375

cyclic nucleotide-gated (HCN) channels, and voltage-gated K+, Na+ and Ca2+ channels, have been found 376

in RGC dendrites, differing in density and dendritic locations between cell types (Van Hook et al., 2019). 377

An earlier theoretical study suggested that alpha RGCs – with their large dendritic arbours, thick and 378

short proximal but thin and long distal branches (Wässle et al., 1981) – feature independent dendritic 379

regions (Koch et al., 1982). In contrast, RGCs with constant dendritic diameter and branch length across 380

their dendritic arbour are thought to produce densely coupled dendritic regions. In these RGCs, their 381

morphology could enable more efficient dendritic backpropagation and therefore lead to the 382

synchronization of dendritic signals (Tran-Van-Minh et al., 2015). Indeed, we observed more 383

independent dendritic regions in tOff alpha cells, but more spatially synchronized dendritic regions in 384

tOff mini, sOff and F-miniOff cells. In tOff mini and tOff alpha cells, their forward and backward 385

propagation were differentially modulated by the same combinations of ion channel densities, 386

confirming that dendritic morphology is a key determinant of dendritic signal propagation efficiency. 387

However, our simulation results suggest that the dendritic integration properties of tOff alpha and tOff 388

mini RGCs could not be explained by dendritic morphology alone but require dendritic ion channels in 389

agreement with earlier simulation studies (Maturana et al., 2014). One possible reason might be that 390

for most RGCs, action potentials generated in the soma can back propagate to the dendritic arbour 391

(Velte and Masland, 1999), which needs dendritic ion channels to enable the efficient backpropagation 392

(van Rossum et al., 2003; Velte and Masland, 1999). 393

Our simulation results are based on highly simplified ball-and-stick models of RGC dendrites, as these 394

allowed us to focus on the principles of dendritic integration. Obviously, these models come with 395

several caveats and possibilities for future extensions: First, the morphological parameters we used 396

were extracted from EM data (Bae et al., 2018), where tissue can shrink due to chemical fixation, such 397

that we may have underestimated the axial conductance based on dendrite diameter. Second, it has 398

been reported that the branching pattern is an important variable for determining propagation 399

efficiency of dendritic signals, mainly because of the diameter changes at branch-points (Ferrante et 400

al., 2013). Third, the density and complement of ion channels can vary along the dendrite (Hausser et 401

al., 2000; Van Hook et al., 2019), raising the possibility that spatially varying ion channel densities 402

would allow for more refined control over dendritic computations. Finally, dendritic signalling is driven 403

by the complex interaction of excitatory and inhibitory inputs (as already mentioned above) and the 404

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 16: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

16

locations of the respective synapses, which will require more precise connectomic studies of the cell 405

types and microcircuits in question. A more realistic model incorporating these aspects could allow 406

additional insights into the mechanisms underlying the observed spatio-temporal dendritic integration 407

rules. 408

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 17: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

17

METHODS 409

Animals and tissue preparation 410

Mice used in this study were purchased from Jackson Laboratory and housed under a standard 12 hour 411

day/night cycle. For all experiments, mice aged 5-8 weeks of either sex were used. We used the 412

transgenic mouse line B6;129P2-Pvalbtm1(cre)Arbr/J (“PV”, JAX 008069, The Jackson Laboratory, Bar Habor, 413

ME; (Hippenmeyer et al., 2005)) cross-bred with the red florescence Cre-dependent reporter line 414

Gt(ROSA)26Sortm9(CAG-tdTomato)Hze (“Ai9tdTomato“, JAX 007905) for all recordings of tOff mini, sOff and F-415

miniOff cells (n=25 animals). For alpha RGC recordings, we also used the wild-type line (C57Bl/6J, JAX 416

000664, n=3 animals), as alpha RGCs can be easily targeted due to their large soma size. All animal 417

procedures followed the laws governing animal experimentation issued by the German federal 418

government and were approved by the institutional animal welfare committee of the University of 419

Tübingen. 420

Mice were dark adapted ≥ 2 hours before tissue preparation, then anaesthetized with isoflurane 421

(Baxter, Hechingen Germany) and killed with cervical dislocation. The eyes were quickly enucleated in 422

carboxygenated (95% O2, 5% CO2) artificial cerebral spinal fluid (ACSF) solution containing (in mM): 125 423

NaCl, 2.5 KCl, 2 CaCl2, 1 MgCl2, 1.25 NaH2PO4, 26 NaHCO3, 20 glucose, and 0.5 L-glutamine (pH 7.4). 424

After removing cornea, sclera and vitreous body, the retina was flattened on an Anodisc (0.2 µm pore 425

size, GE Healthcare, Pittsburgh, PA) with the ganglion cell side facing up and then transferred to the 426

recording chamber of the microscope, where it was continuously perfused with carboxygenated ACSF 427

(at 35°C and 4 ml/min). All experimental procedures were carried out under very dim red light. 428

Loading of individual cells with calcium indicator 429

To visualize blood vessels and avoiding them when filling individual RGCs, 5 µl of a 50 mM sulforho-430

damine-101 (SR101, Invitrogen /Thermo Fisher Scientific, Dreieich, Germany) stock solution was added 431

per litre ACSF solution. Sharp electrodes for single-cell injection were pulled on a P-1000 micropipette 432

puller (Sutter Instruments, Novato, CA) with resistances ranging between 70 and 130 MΩ. Oregon 433

Green BAPTA-1 (OGB-1, hexapotassium salt; Life Technologies, Darmstadt, Germany; 15 mM in water) 434

was loaded into individual RGCs using the single-pulse function (500 ms, -10 nA) of a MultiClamp 900A 435

amplifier (Axon Instruments/Molecular Devices, Wokingham, UK). To allow the cells to completely fill 436

and recover, we started recordings 1 h post-injection. 437

Two-photon imaging and light stimulation 438

A MOM-type two-photon microscope (designed by W. Denk, MPI, Martinsried; purchased from Sutter 439

Instruments/Science Products) as described previously (Euler et al., 2009) was used for this study. 440

Briefly, the system was equipped with a mode-locked Ti:Sapphire laser (MaiTai-HP DeepSee, Newport 441

Spectra-Physics, Darmstadt, Germany), green and red fluorescence detection channels for OGB-1 (HQ 442

510/84, AHF, Tübingen, Germany) and SR101/tdTomato (HQ 630/60, AHF), and a water immersion 443

objective (W Plan-Apochromat 20x/1,0 DIC M27, Zeiss, Oberkochen, Germany). For all scans, we tuned 444

the laser to 927 nm, and used a custom-made software (ScanM, by M. Müller, MPI, Martinsried, and 445

T.E.) running under IGOR Pro 6.3 for Windows (Wavemetrics, Portland, OR). Time-elapsed dendritic 446

signals were recorded with 64×16 pixel image sequences (31.25 Hz). High-resolution morphology 447

stacks were acquired using 512×512 pixel image stacks with 0.8 or 1.0 µm z steps. 448

Light stimuli were projected through the objective lens (Euler et al., 2009). We used two alternative 449

digital light processing (DLP) projectors: a K11 (Acer, Ahrensburg, Germany) or a LightCrafter E4500 450

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 18: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

18

MKII (Texas Instruments, Dallas, TX; modified by EKB Technologies Ltd., Israel). Both were equipped 451

with light-emitting diodes (LEDs) – “green” (575 nm) and UV (390 nm) that match the spectral 452

sensitivities of mouse M- and S-opsins (for details, see (Baden et al., 2016; Franke et al., 2019)). Both 453

LEDs were intensity-calibrated to range from 0.1 x103 (“black“ background) to 20.0 x103 (“white“ full 454

field) photoisomerisations P*/s/cone. The light stimulus was centred before every experiment, 455

ensuring that its centre corresponded to the centre of the microscope’s scan field. For all experiments, 456

the tissue was kept at a constant mean stimulator intensity level for ≥ 15 s after the laser scanning 457

started and before light stimuli were presented. 458

Light stimuli were generated and presented using the Python-based software package QDSpy (see 459

Table 4). Three types of light stimuli were used: 460

(1) binary dense noise (20x15 matrix of 30 μm/pixel; each pixel displayed an independent, balanced 461

random sequence at 5 Hz for 5 minutes) for spatio-temporal receptive field (RF) mapping; 462

(2) full-field (800x600 µm) “chirp”, consisting of a bright step and two sinusoidal intensity 463

modulations, one with increasing frequency (0.5-8 Hz) and one with increasing contrast. 464

(3) local “chirp”; like (2) but with a diameter of 100 µm. 465

Reconstruction of cell morphologies 466

Directly after the recording, the complete dendritic morphology of the RGC was captured by acquiring 467

a high-resolution stack. In case the cell was not bright enough to see all branches in detail, a second 468

dye-injection was performed. Using semi-automatic neurite tracing (Longair et al., 2011) , we obtained 469

cell skeletons of the recorded RGCs. If necessary, we de-warped image stack and traced cell, as 470

described earlier (Baden et al., 2016). All further analysis, such as the extraction of morphological 471

parameters (see below), was done using custom Python scripts. 472

Relating recording positions to cell morphology 473

As the full dendritic morphology could not be imaged during calcium recordings, recorded dendrites 474

(i.e. ROIs, see below) were not necessarily well-aligned with the cell morphology reconstructed later. 475

Based on the relative position of each recording field, a region 9 times larger than this recording field 476

was cropped from the reconstructed morphology and z-projected. Next, the recording field was 477

automatically aligned to this cropped region using match_templates from scikit-image (Table 4). 478

The centre coordinates of all ROIs in that recording field were then calibrated to the closest dendritic 479

branch based on their Euclidean distance. In rare cases, when automatic matching failed, the matching 480

was done manually. 481

Morphological parameters and hierarchical clustering 482

To morphologically cluster the RGCs as described by Bae et al. (2018), we had to determine the relative 483

position of the two ChAT bands, the dendritic plexi of the starburst amacrine cells (Vaney, 1984). For 484

this, the blood vessel plexi in GCL and INL served as landmarks. With their positions defined as 0 and 485

1, the relative IPL depth of the On and Off ChAT bands is 0.48 and 0.77, respectively, as shown earlier 486

(Baden et al., 2016; Franke et al., 2017). The following parameters were extracted for each cell: 487

To determine the marginal-central arbour density difference, we defined the central IPL as the 488

portion between the ChAT bands and the remainder (On ChAT band to GCL, Off ChAT band to INL) 489

as marginal IPL (cf. Fig. 2b, inset). The marginal-central arbour density difference was calculated 490

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 19: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

19

using the sum of the dendritic length in central IPL minus the sum of the dendritic length located in 491

marginal IPL. 492

Dendritic arbour area was calculated as the area of the tightest convex hull containing the z-493

projected dendritic arbour. 494

Asymmetry of the dendritic arbour was calculated as the distance between the centre of mass of 495

dendritic density and the soma position. 496

Soma size was defined as soma area. For this, the image frame in which the soma appeared the 497

largest was used. 498

Dendritic distance between ROIs was defined as the shortest distance along the dendrite between 499

two ROIs. 500

Angular distance between ROIs was defined as the positive angle between two ROIs and the nearest 501

branching point (cf. Fig. 4b). 502

n Arbour density difference (a.u.)

Area (103 µm2)

Asymmetry (a.u.)

Soma size (µm2)

tOff alpha 17 -2.41 ± 0.44 53.2 ± 2.3 44.9 ± 6.3 322.5 ± 8.3

tOff mini 5 -5.10 ± 0.57 29.7 ± 1.9 18.9 ± 4.1 151.5 ± 14.8

sOff 4 8.58 ± 1.19 29.0 ± 1.5 14.2 ± 3.7 204.0 ± 34.1

F-miniOff 5 6.77 ± 0.62 11.5 ± 0.6 74.8 ± 6.2 102.7 ± 2.2

Table 1 | Morphological parameters describing the dendritic arbours of the clustered RGCs. For parameter 503 definitions, see Methods. 504

Hierarchical clustering was performed with 1D k-means clustering with k=2 for all splits, using KMeans 505

from the Python package scikit-learn (Table 4). First, cells were split into two clusters based on arbour 506

density difference (cf. Fig. 2b). Next, the group with lower arbour density difference was separated by 507

soma size, while the group with the higher arbour density difference was further split based on their 508

asymmetry index. Here, we refrained from further splitting, because the cells in each group displayed 509

highly consistent light responses. Thus, these four groups were used for further analysis. 510

Data analysis 511

All data were analysed using custom scripts: For data preprocessing, we used IGOR Pro; further analysis 512

and modelling was done using Python and R. Upon publication, all data, scripts and models will become 513

available (see links in Table 4). 514

Regions of interest (ROIs) – We used dense noise recordings to extract ROIs. First, for each recorded 515

field, the standard deviation (s.d.) of the fluorescence intensity for each pixel over time was calculated, 516

generating an s.d. image of the time-lapsed image stack. Pixels brighter than the mean of the s.d. image 517

plus 1 s.d. were considered dendritic pixels. Then, in each recorded field, the time traces of the 100 518

most responsive dendritic pixels (=100 brightest dendritic pixels in the s.d. image) were extracted and 519

cross-correlated. The mean of the resulting cross-correlation coefficients (𝜌) served as correlation 520

threshold (𝜌𝑇ℎ𝑟𝑒𝑠ℎ𝑜𝑙𝑑) for each field. Next, we grouped neighbouring pixels (within a distance of 3 µm) 521

with 𝜌 > 𝜌𝑇ℎ𝑟𝑒𝑠ℎ𝑜𝑙𝑑 into one ROI. Finally, each ROI’s Ca2+ trace was extracted using the image analysis 522

toolbox SARFIA for IGOR Pro (Dorostkar et al., 2010). A time marker embedded in the recorded data 523

served to align the traces relative to the visual stimulus with 2 ms precision. All stimulus-aligned traces 524

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 20: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

20

together with the relative ROI positions on the recorded cell’s dendritic arbour were exported for 525

further analysis. 526

Dendritic receptive fields – Dendritic RFs were estimated using Automatic Smoothness Determination 527

(ASD, (Sahani and Linden, 2003)), a linear-Gaussian encoding model within the empirical Bayes 528

framework. The relationship between stimulus and response was modelled as a linear function plus 529

Gaussian noise: 530

yi = 𝑘𝑇xi + 𝜀i, 𝜀i~ 𝑁(0, δ2) (1) 531

where x is the binary dense noise stimulus (20x15 matrix of 30 µm/pixel), y is the gradient of the Ca2+ 532

response, 𝑘 is the spatio-temporal RF (STRF) with a time lag ranging from -1,000 ms to 0 ms, and 𝜀 is 533

independent and identically distributed (i.i.d.) Gaussian noise with zero mean and δ2 variance. 534

The STRF was then calculated in two steps (Sahani and Linden, 2003): First, the ASD prior covariance 535

( 𝐶𝑖𝑗 = exp(-ρ-Δ𝑖𝑗/2δ ) , where Δ𝑖𝑗 is the squared distance between any two filter coefficients), 536

controlled by the spatial and temporal smoothness (δ) and scale (ρ), was optimized using evidence 537

optimization. Then, the STRF was estimated by maximum a posteriori linear regression between 538

response and stimulus using the optimized prior. The spatial RF maps shown represent the spatial 539

component of the singular value decomposition of the STRF. 540

To quantify the quality of spatial RFs, contours were drawn on up-sampled and normalized RF maps 541

with different thresholds (0.60, 0.65 and 0.70). The quality was then determined from the number of 542

contours, their sizes and their degree of irregularity. The irregularity index was defined as 543

𝐼𝑖 = 1 −𝐴𝑐𝑜𝑛𝑡𝑜𝑢𝑟

𝐴𝑐𝑜𝑛𝑣𝑒𝑥 ℎ𝑢𝑙𝑙 (2) 544

with 𝐴𝑐𝑜𝑛𝑡𝑜𝑢𝑟 corresponding the area of the RF contour, and 𝐴𝑐𝑜𝑛𝑣𝑒𝑥 ℎ𝑢𝑙𝑙 the 2D morphology’s convex 545

hull. Only data with a good RF (a single contour with 𝐼𝑖 < 0.1 , 𝐴𝑐𝑜𝑛𝑡𝑜𝑢𝑟 > 1.8 ∙ 1,000 µm2, at a 546

contour threshold of 0.60; see Suppl. Fig. S1) were used for further analyses. 547

Offset between ROI centre and its RF centre – RF offset was calculated as the linear distance between 548

ROI centre and the geometrical centre of its RF contour. 549

Dendritic RF overlap index – An RF overlap index (𝑂𝑖) was calculated as follows: 550

𝑂𝑖 =Ao

Amin[A1,A2] (3) 551

where 𝐴1 and 𝐴2 are the RF areas of the ROI pair, Ao is the overlap area between 𝐴1 and 𝐴2, and 552

𝐴𝑚𝑖𝑛[𝐴1, 𝐴2] corresponds to the smaller area (𝐴1 or 𝐴2). 553

Full-field chirp and local chirp – Ca2+ traces for full-field and local chirp stimuli were linearly up-sampled 554

(interpolated) to 500 Hz, baseline-subtracted (using the mean of 2,500 samples before light stimulus 555

onset) and normalized by the s.d. of this baseline. To estimate the signal-to-noise ratio, we calculated 556

the response quality index (𝑄𝑖) for both full-field and local chirps as described before (Franke et al., 557

2017): 558

𝑄𝑖 =Var[⟨𝐶⟩r]t

Var[⟨𝐶⟩t]r (4) 559

where 𝐶 is the T-by-R response matrix (time samples by stimulus repetitions) and ⟨⟩x and Var[]x 560

denote the mean and variance across the indicated dimension, respectively. If all trials were identical, 561

such that the mean response is a perfect representative of the response, 𝑄𝑖 = 1. If all trials were 562

completely random with fixed variance, such that the mean response is not informative about the 563

individual trials, 𝑄𝑖 ∝ 1/𝑅. 564

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 21: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

21

Signal correlation – To quantify temporal signal correlation, we cross-correlated the mean Ca2+ 565

responses. We noticed that some ROIs with good spatial RFs (see above) displayed a low signal-to-566

noise chirp responses. Hence, we repeated the analysis for 𝑄𝑖 > 0.4 or 0.5 with comparable result 567

(Supplemental Fig. S6). 568

Further temporal analysis – Using the responses to the step part of the chirp stimulus, we calculated a 569

transience index (𝑇𝑖, for local chirp) and a polarity index (𝑃𝑂𝑖, for both local and full-field chirps). Here, 570

only ROIs with 𝑄𝑖 > 0.4 were used for the analysis. Before the computation of these indices, the mean 571

traces were binomially smoothed (with 3,000 repetitions). Then, 2 s.d. of the baseline (2,500 samples 572

of the smoothed trace before light stimulus onset) were used to determine the time of response onset 573

(𝑇𝑅_𝑜𝑛𝑠𝑒𝑡) and offset (𝑇𝑅_𝑜𝑓𝑓𝑠𝑒𝑡). 574

𝑇𝑖 was calculated as 575

𝑇𝑖 = 1 −𝑇𝑅_𝑜𝑓𝑓𝑠𝑒𝑡−𝑇𝑅_𝑜𝑛𝑠𝑒𝑡

𝑇𝑠𝑡𝑖𝑚𝑢𝑙𝑢𝑠 (5) 576

where 𝑇𝑠𝑡𝑖𝑚𝑢𝑙𝑢𝑠 is the stimulus duration. 577

For 𝑃𝑂𝑖, data points before and after the response (see above) were set to zero, before calculating: 578

POI =∑ 𝑟(𝑡+𝑡𝑠𝑡𝑖𝑚 𝑜𝑛)− ∑ 𝑟𝑏

𝑡=0 (𝑡+𝑡𝑠𝑡𝑖𝑚 𝑜𝑓𝑓)𝑏𝑡=0

∑ 𝑟(𝑡+𝑡𝑠𝑡𝑖𝑚 𝑜𝑛)+∑ 𝑟𝑏𝑡=0 (𝑡+𝑡𝑠𝑡𝑖𝑚 𝑜𝑓𝑓)𝑏

𝑡=0 (6) 579

Where 𝑏 = 3 s, 𝑡𝑠𝑡𝑖𝑚 𝑜𝑛 and 𝑡𝑠𝑡𝑖𝑚 𝑜𝑓𝑓 are the time points of light stimulus onset and offset, 𝑟(𝑡) is the 580

mean response at time 𝑡. For ROIs responding only to the light-onset, 𝑃𝑂𝑖 = 1, whereas for ROIs only 581

responding during the light-offset, 𝑃𝑂𝑖 = −1. 582

Statistical analysis 583

We used Generalized Additive Models (GAMs) to analyse the relationships of RF size vs. dendritic 584

distance; RF offset vs. dendritic distance; RF overlap vs. dendritic distance and dendritic angle; 585

temporal correlation vs. dendritic distance and dendritic angle (for details, see Supplemental 586

Information). GAMs extend the generalized linear model by allowing the linear predictors to depend 587

on arbitrary smooth functions of the underlying variables (Wood, 2006): 588

𝑔(𝜇) = 𝛽0 + 𝑓1(𝑥1) + ⋯ + 𝑓𝑛(𝑥𝑛) (7) 589

Here, 𝑥𝑖 are the predictor variables, 𝑔 is a link function, and the 𝑓𝑖 are smooth functions of the 590

predictor variables. These smooth functions can also depend on more than one predictor variable. 591

To implement GAMs and perform statistical testing, we employed the mgcv package for R (Table 4). 592

Here, for smooth terms we used penalized regression splines. We modelled the dependence of our 593

variable of interest as a single smooth term per cell type for univariate dependencies and a tensor 594

product smooth for bivariate dependencies. The dimension of the basis was set high enough such that 595

the estimated degrees of freedom stayed sufficiently below the possible maximum. Smoothing 596

parameters were selected via the default methods of the package. All models also included a random 597

effect term for cell identity. Typically, we used models from the Gaussian family; for the dependence 598

of RF overlap on dendritic distance and dendritic angle, we instead used a scaled t-distribution, as this 599

improved BIC (Bayesian information criterion) and diagnostic plots. 600

Statistical significance for differences in the obtained smooths between cell types were performed 601

using plot_diff or plot_diff2 of the itsadug package for R (Table 4). 95% confidence intervals 602

were calculated using the simultaneous confidence intervals (CI) option, excluding the random effect 603

of cell identity. 604

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 22: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

22

Biophysical model 605

To explore the mechanisms underlying dendritic integration in different RGC types, we built a multi-606

compartmental 1D model. To get precise measurements of dendrite thickness and segment length for 607

tOff alpha and tOff mini cells, we extracted these information from published morphologies of “4ow” 608

and “4i” RGCs (cf. Fig. 6b), respectively, reconstructed from EM data (http://museum.eyewire.org). 609

Then we mapped the medium values of these parameters to the respective branch order of the model 610

(cf. Fig. 6c). The model was implemented in the NEURON simulation environment (Hines and Carnevale, 611

1997). Here, each dendritic portion (between two branch points) was represented as a “section” in the 612

simulator, which was further divided into multiple “segments” (compartments) with a maximal length 613

of 7 µm. The 1D model can be characterized by the cable equation, 614

𝑑

4𝑎

∂2𝑉

∂𝑥2 = 𝐶𝑚𝜕𝑉

𝜕𝑡+ 𝐼𝑖𝑜𝑛 − 𝐼𝑠𝑡𝑖𝑚 (8) 615

where 𝑉 is the voltage across the cell membrane, 𝑥 is the distance along the cable, 𝑑 is the dendritic 616

diameter, 𝑎 is the intracellular resistivity, and 𝐶𝑚 is the specific membrane capacitance. 𝐼𝑖𝑜𝑛 617

represents the sum of four voltage-gated cation currents (sodium, 𝐼𝑁𝑎; calcium, 𝐼𝐶𝑎; delayed rectifier 618

potassium, 𝐼𝐾; A-type potassium, 𝐼𝐾,𝐴), one calcium-activated potassium current (𝐼𝐾,𝐶𝑎), and one leak 619

current (𝐼𝐿𝑒𝑎𝑘). The current dynamics are described following Fohlmeister and Miller (1997) as: 620

𝐼𝑖𝑜𝑛 = 𝐼𝑁𝑎 + 𝐼𝐶𝑎 + 𝐼𝐾 + 𝐼𝐾,𝐴 + 𝐼𝐾,𝐶𝑎 + 𝐼𝐿𝑒𝑎𝑘 (9) 621

= 𝑁𝑎𝑚3ℎ(𝑉 − 𝑉𝑁𝑎) + 𝐶𝑎𝑐3(𝑉 − 𝑉𝐶𝑎) 622

+(𝐾𝑛4 + 𝐾,𝐴𝑎3ℎ𝐴 + 𝐾,𝐶𝑎)(𝑉 − 𝑉𝐾) 623

+𝐿𝑒𝑎𝑘(𝑉 − 𝑉𝑙𝑒𝑎𝑘) 624

The intracellular stimulation current (𝐼𝑠𝑡𝑖𝑚 in Eq. 8) was the product of a 5,000 ms x 200 µm 1D 625

Gaussian noise stimulus and a BC’s spatial RF with a Gaussian shape (with the width set by σ=6 and the 626

centre depends on the current injection location). The stimulation current was injected either at the 627

proximal (at 25 µm from soma) or distal dendrite (at 85% of the total dendrite length from soma), then 628

filtered by a soft rectification function. The magnitude of 𝐼𝑠𝑡𝑖𝑚 was scaled between 0 and 15 nA with a 629

mean of 6.17 nA and an s.d. of 1.88 nA to ensure that the input stimulus would elicit spikes in the soma 630

for all parameter combinations. 631

Model parameters (Table 2) and channel conductances (Table 3) were taken from Fohlmeister and 632

Miller (1997). To identify parameters that explain the experimental data, we grid-searched 633

combinations of ion channel densities by multiplying the reference parameters with different scaling 634

factors (for 𝐾 and 𝐶𝑎, the scaling factors ranged from 0 to 2, and were incremented by 0.25 each 635

step; for 𝑁𝑎, we used [0.1, 1, 2], for 𝐾,𝐶𝑎, [10, 100, 1000, 5000, 9000]). 636

Parameters Values

Temperature T = 32oC

Intracellular axial resistivity 𝑅𝑎 = 110 Ω∙cm

Specific membrane resistance 𝑅𝑚= 15000 Ω∙cm2

Specific membrane Capacitance 𝐶𝑚= 1 μF/cm2

Potassium reversal potential 𝑉𝐾 = -75 mV

Sodium reversal potential 𝑉𝑁𝑎= 35 mV

Table 2 | Model parameters. 637

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 23: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

23

Channel type Conductance in [S/cm2]

Soma Dendrites

𝑁𝑎 0.08 0.025

𝐶𝑎 0.0015 0.002

𝐾 0.018 0.012

𝐾,𝐴 0.054 0.036

𝐾,𝐶𝑎 0.000065 0.000001

Table 3 | Reference distribution of ion channels in cell compartments. 638

Voltage changes in the dendrites were read out at the centre of each dendritic “section” and used to 639

estimate the local dendritic RF with a maximum likelihood method, then smoothed by a Savitzky-Golay 640

filter (window length of 31; 3rd degree polynomial). The peak amplitudes of dendritic RFs were 641

measured and normalized to the RF at the position closest to the current input. 642

Part Description (link) Company /Author Item number (RRID)

ScanM 2P imaging software running under IGOR Pro

Written by M. Müller (MPI Neuro-biology, Martinsried), and T.E.

IGOR Pro (v6)

https://www.wavemetrics.com Wavemetrics, Lake Oswego, OR IGOR Pro v6 (SCR_000325)

Python (v3.6.7)

http://www.python.org/

R (v2.3)

The R project http://www.r-project.org/

QDSpy (v0.77)

Visual stimulation software https://github.com/eulerlab/QDSpy

Written by T.E, supported by Tom Boissonnet (EMBL, Monterotondo)

- (SCR_016985)

Scikit-learn (v0.20.0)

Software package for Python (Pedregosa et al., 2011)

Scikit-Image (v0.14.2)

Software package for Python (van der Walt et al., 2014)

itsadug (v2.3)

Software package for R (van Rij et al., 2017)

mgcv (v1.8-24)

Software package for R (Wood, 2006)

NEURON

(v7.7.0)

Simulation environment for modelling individual neurons and networks of neurons

(Carnevale and and Hines, 2006)

Custom scripts and data

http://retinal-functomics.net/

Table 4 | Software used and repositories for custom scripts and data. 643

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 24: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

24

ACKNOWLEDGEMENTS 644

We thank Huayu Ding for helping us with single-cell microinjections, Luke Rogerson for help with 645

statistics und discussion, Zhijian Zhao and Gordon Eske for excellent technical support, and Timm 646

Schubert for discussion. 647

This research was supported by Deutsche Forschungsgemeinschaft (DFG, EXC307 to TE and EXC 2064, 648

project number 390727645 to PB; BE5601/4-1, BE5601/6-1 to PB; EU 42/10-1 to TE); NINDS of the 649

National Institutes of Health (U01NS090562 to TE); BMBF (01GQ1601 and 01IS18052C to PB; 650

01GQ1002 to KF); BWSF (AZ 1.16101.09 to TB); MPG (M.FE.A.KYBE0004 to KF). 651

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 25: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

25

REFERENCES 652

Ames, A.I., and Nesbett, F.B. (1981). In Vitro Retina as an Experimental Model of the Central Nervous 653 System. Journal of Neurochemisry 37, 867-877. 654

Baden, T., Berens, P., Franke, K., Roman Roson, M., Bethge, M., and Euler, T. (2016). The functional 655 diversity of retinal ganglion cells in the mouse. Nature 529, 345-350. 656

Baden, T., Schubert, T., Berens, P., and Euler, T. (2018). The Functional Organization of Vertebrate 657 Retinal Circuits for Vision. Oxford Research Encyclopedia of Neuroscience, 658 10.1093/acrefore/9780190264086.013.68. 659

Bae, J.A., Mu, S., Kim, J.S., Turner, N.L., Tartavull, I., Kemnitz, N., Jordan, C.S., Norton, A.D., 660 Silversmith, W.M., Prentki, R., et al. (2018). Digital Museum of Retinal Ganglion Cells with 661 Dense Anatomy and Physiology. Cell 173, 1293-1306 e1219. 662

Barlow, H.B., and Levick, W.R. (1965). The mechanism of directionally selective units in rabbit's 663 retina. J Physiol 178, 477-504. 664

Behabadi, B.F., and Mel, B.W. (2014). Mechanisms underlying subunit independence in pyramidal 665 neuron dendrites. Proc Natl Acad Sci U S A 111, 498-503. 666

Bernander, O., Douglas, R.J., Martin, K.A., and Koch, C. (1991). Synaptic background activity 667 influences spatiotemporal integration in single pyramidal cells. Proc Natl Acad Sci U S A 88, 668 11569-11573. 669

Branco, T., and Hausser, M. (2010). The single dendritic branch as a fundamental functional unit in 670 the nervous system. Curr Opin Neurobiol 20, 494-502. 671

Carnevale, N.T., and and Hines, M.L. (2006). The NEURON Book (Cambridge, UK: Cambridge 672 University Press). 673

Chapot, C.A., Behrens, C., Rogerson, L.E., Baden, T., Pop, S., Berens, P., Euler, T., and Schubert, T. 674 (2017). Local Signals in Mouse Horizontal Cell Dendrites. Curr Biol 27, 3603-3615 e3605. 675

Deny, S., Ferrari, U., Mace, E., Yger, P., Caplette, R., Picaud, S., Tkacik, G., and Marre, O. (2017). 676 Multiplexed computations in retinal ganglion cells of a single type. Nat Commun 8, 1964. 677

Diamond, J.S. (2017). Inhibitory Interneurons in the Retina: Types, Circuitry, and Function. Annu Rev 678 Vis Sci 3, 1-24. 679

Dorostkar, M.M., Dreosti, E., Odermatt, B., and Lagnado, L. (2010). Computational processing of 680 optical measurements of neuronal and synaptic activity in networks. J Neurosci Meth 188, 681 141-150. 682

Euler, T., Detwiler, P.B., and Denk, W. (2002). Directionally selective calcium signals in dendrites of 683 starburst amacrine cells. Nature 418, 845-852. 684

Euler, T., Hausselt, S.E., Margolis, D.J., Breuninger, T., Castell, X., Detwiler, P.B., and Denk, W. (2009). 685 Eyecup scope--optical recordings of light stimulus-evoked fluorescence signals in the retina. 686 Pflugers Arch 457, 1393-1414. 687

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 26: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

26

Ferrante, M., Migliore, M., and Ascoli, G.A. (2013). Functional impact of dendritic branch-point 688 morphology. J Neurosci 33, 2156-2165. 689

Field, G.D., Gauthier, J.L., Sher, A., Greschner, M., Machado, T.A., Jepson, L.H., Shlens, J., Gunning, 690 D.E., Mathieson, K., Dabrowski, W., et al. (2010). Functional connectivity in the retina at the 691 resolution of photoreceptors. Nature 467, 673-U654. 692

Fohlmeister, J.F., and Miller, R.F. (1997). Impulse encoding mechanisms of ganglion cells in the tiger 693 salamander retina. J Neurophysiol 78, 1935-1947. 694

Franke, K., Berens, P., Schubert, T., Bethge, M., Euler, T., and Baden, T. (2017). Inhibition decorrelates 695 visual feature representations in the inner retina. Nature 542, 439-444. 696

Franke, K., Chagas, A.M., Zhao, Z., Zimmermann, M.J.Y., Qiu, Y., Szatko, K., Baden, T., and Euler, T. 697 (2019). Preprint: An arbitrary-spectrum spatial visual stimulator for vision research. bioRxiv 698 10.1101/649566. 699

Goldstein, S.S., and Rall, W. (1974). Changes of action potential shape and velocity for changing core 700 conductor geometry. Biophys J 14, 731-757. 701

Grimes, W.N., Zhang, J., Graydon, C.W., Kachar, B., and Diamond, J.S. (2010). Retinal parallel 702 processors: more than 100 independent microcircuits operate within a single interneuron. 703 Neuron 65, 873-885. 704

Guo, T., Tsai, D., Bai, S., Morley, J.W., Suaning, G.J., Lovell, N.H., and Dokos, S. (2014). Understanding 705 the retina: a review of computational models of the retina from the single cell to the network 706 level. Crit Rev Biomed Eng 42, 419-436. 707

Hausselt, S.E., Euler, T., Detwiler, P.B., and Denk, W. (2007). A dendrite-autonomous mechanism for 708 direction selectivity in retinal starburst amacrine cells. PLoS Biol 5, e185. 709

Hausser, M., Spruston, N., and Stuart, G.J. (2000). Diversity and dynamics of dendritic signaling. 710 Science 290, 739-744. 711

Helmstaedter, M., Briggman, K.L., Turaga, S.C., Jain, V., Seung, H.S., and Denk, W. (2013). 712 Connectomic reconstruction of the inner plexiform layer in the mouse retina. Nature 500, 713 168-174. 714

Hines, M.L., and Carnevale, N.T. (1997). The NEURON simulation environment. Neural Comput 9, 715 1179-1209. 716

Hippenmeyer, S., Vrieseling, E., Sigrist, M., Portmann, T., Laengle, C., Ladle, D.R., and Arber, S. (2005). 717 A developmental switch in the response of DRG neurons to ETS transcription factor signaling. 718 PLoS Biol 3, e159. 719

Koch, C., Poggio, T., and Torre, V. (1982). Retinal ganglion cells: a functional interpretation of 720 dendritic morphology. Philos Trans R Soc Lond B Biol Sci 298, 227-263. 721

Koren, D., Grove, J.C.R., and Wei, W. (2017). Cross-compartmental Modulation of Dendritic Signals 722 for Retinal Direction Selectivity. Neuron 95, 914-927 e914. 723

Krieger, B., Qiao, M., Rousso, D.L., Sanes, J.R., and Meister, M. (2017). Four alpha ganglion cell types 724 in mouse retina: Function, structure, and molecular signatures. PLoS One 12, e0180091. 725

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 27: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

27

Lai, H.C., and Jan, L.Y. (2006). The distribution and targeting of neuronal voltage-gated ion channels. 726 Nat Rev Neurosci 7, 548-562. 727

Lee, S., Kim, K., and Zhou, Z.J. (2010). Role of ACh-GABA Cotransmission in Detecting Image Motion 728 and Motion Direction. Neuron 68, 1159-1172. 729

Longair, M.H., Baker, D.A., and Armstrong, J.D. (2011). Simple Neurite Tracer: open source software 730 for reconstruction, visualization and analysis of neuronal processes. Bioinformatics 27, 2453-731 2454. 732

Magee, J.C. (2000). Dendritic integration of excitatory synaptic input. Nat Rev Neurosci 1, 181-190. 733

Maturana, M.I., Kameneva, T., Burkitt, A.N., Meffin, H., and Grayden, D.B. (2014). The effect of 734 morphology upon electrophysiological responses of retinal ganglion cells: simulation results. 735 J Comput Neurosci 36, 157-175. 736

Oesch, N., Euler, T., and Taylor, W.R. (2005). Direction-selective dendritic action potentials in rabbit 737 retina. Neuron 47, 739-750. 738

Pang, J.J., Gao, F., and Wu, S.M. (2003). Light-evoked excitatory and inhibitory synaptic inputs to ON 739 and OFF alpha ganglion cells in the mouse retina. J Neurosci 23, 6063-6073. 740

Pedregosa, F., Varoquaux, G., Gramfort, A., Michel, V., Thirion, B., Grisel, O., Blondel, M., 741 Prettenhofer, P., Weiss, R., Dubourg, V., et al. (2011). Scikit-learn: Machine Learning in 742 Python. J Mach Learn Res 12, 2825-2830. 743

Poleg-Polsky, A., Ding, H., and Diamond, J.S. (2018). Functional Compartmentalization within 744 Starburst Amacrine Cell Dendrites in the Retina. Cell Rep 22, 2898-2908. 745

Polsky, A., Mel, B.W., and Schiller, J. (2004). Computational subunits in thin dendrites of pyramidal 746 cells. Nat Neurosci 7, 621-627. 747

Rall, W., and Rinzel, J. (1973). Branch Input Resistance and Steady Attenuation for Input to One 748 Branch of a Dendritic Neuron Model. Biophys J 13, 648-688. 749

Rheaume, B.A., Jereen, A., Bolisetty, M., Sajid, M.S., Yang, Y., Renna, K., Sun, L., Robson, P., and 750 Trakhtenberg, E.F. (2018). Single cell transcriptome profiling of retinal ganglion cells identifies 751 cellular subtypes. Nat Commun 9, 2759. 752

Roska, B., Molnar, A., and Werblin, F.S. (2006). Parallel processing in retinal ganglion cells: How 753 integration of space-time patterns of excitation and inhibition form the spiking output. J 754 Neurophysiol 95, 3810-3822. 755

Roska, B., and Werblin, F. (2001). Vertical interactions across ten parallel, stacked representations in 756 the mammalian retina. Nature 410, 583-587. 757

Rousso, D.L., Qiao, M., Kagan, R.D., Yamagata, M., Palmiter, R.D., and Sanes, J.R. (2016). Two Pairs of 758 ON and OFF Retinal Ganglion Cells Are Defined by Intersectional Patterns of Transcription 759 Factor Expression. Cell Rep 15, 1930-1944. 760

Sahani, M., and Linden, J.F. (2003). Evidence Optimization Techniques for Estimating Stimulus-761 Response Functions. Advances in neural information processing systems, 317-324. 762

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 28: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

28

Sanes, J.R., and Masland, R.H. (2015). The types of retinal ganglion cells: current status and 763 implications for neuronal classification. Annu Rev Neurosci 38, 221-246. 764

Schachter, M.J., Oesch, N., Smith, R.G., and Taylor, W.R. (2010). Dendritic spikes amplify the synaptic 765 signal to enhance detection of motion in a simulation of the direction-selective ganglion cell. 766 PLoS Comput Biol 6(8), e1000899. 767

Schmidt-Hieber, C., Toleikyte, G., Aitchison, L., Roth, A., Clark, B.A., Branco, T., and Hausser, M. 768 (2017). Active dendritic integration as a mechanism for robust and precise grid cell firing. Nat 769 Neurosci 20, 1114-1121. 770

Schwartz, G.W., Okawa, H., Dunn, F.A., Morgan, J.L., Kerschensteiner, D., Wong, R.O., and Rieke, F. 771 (2012). The spatial structure of a nonlinear receptive field. Nat Neurosci 15, 1572-1580. 772

Siegert, S., Cabuy, E., Scherf, B.G., Kohler, H., Panda, S., Le, Y.Z., Fehling, H.J., Gaidatzis, D., Stadler, 773 M.B., and Roska, B. (2012). Transcriptional code and disease map for adult retinal cell types. 774 Nat Neurosci 15, 487-495, S481-482. 775

Sivyer, B., and Williams, S.R. (2013). Direction selectivity is computed by active dendritic integration 776 in retinal ganglion cells. Nat Neurosci 16, 1848-1856. 777

Stuart, G.J., and Spruston, N. (2015). Dendritic integration: 60 years of progress. Nat Neurosci 18, 778 1713-1721. 779

Sümbül, U., Song, S., McCulloch, K., Becker, M., Lin, B., Sanes, J.R., Masland, R.H., and Seung, H.S. 780 (2014). A genetic and computational approach to structurally classify neuronal types. Nat 781 Commun 5, 3512. 782

Tran-Van-Minh, A., Abrahamsson, T., Cathala, L., and DiGregorio, D.A. (2016). Differential Dendritic 783 Integration of Synaptic Potentials and Calcium in Cerebellar Interneurons. Neuron 91, 837-784 850. 785

Tran-Van-Minh, A., Caze, R.D., Abrahamsson, T., Cathala, L., Gutkin, B.S., and DiGregorio, D.A. (2015). 786 Contribution of sublinear and supralinear dendritic integration to neuronal computations. 787 Front Cell Neurosci 9, 67. 788

Trenholm, S., Johnson, K., Li, X., Smith, R.G., and Awatramani, G.B. (2011). Parallel Mechanisms 789 Encode Direction in the Retina. Neuron 71, 683-694. 790

van der Walt, S., Schonberger, J.L., Nunez-Iglesias, J., Boulogne, F., Warner, J.D., Yager, N., Gouillart, 791 E., Yu, T., and Contributors, S.I. (2014). scikit-image: image processing in Python. Peerj 2. 792

Van Hook, M.J., Nawy, S., and Thoreson, W.B. (2019). Voltage- and calcium-gated ion channels of 793 neurons in the vertebrate retina. Progress in Retinal and Eye Research, 794 10.1016/j.preteyeres.2019.05.001. 795

van Rij, J., Wieling, M., Baayen, R., and van Rijn, H. (2017). itsadug: Interpreting Time Series and 796 Autocorrelated Data Using GAMMs. R package version 2.3. 797

van Rossum, M.C., O'Brien, B.J., and Smith, R.G. (2003). Effects of noise on the spike timing precision 798 of retinal ganglion cells. J Neurophysiol 89, 2406-2419. 799

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint

Page 29: Type-specific dendritic integration in mouse retinal ... · 84 individual cells (Fig. 1a). After recording dendritic activity, the cells were 3D-reconstructed (Fig. 1b), 85 allowing

29

Van Wyk, M., Wassle, H., and Taylor, W.R. (2009). Receptive field properties of ON- and OFF-ganglion 800 cells in the mouse retina. Vis Neurosci 26, 297-308. 801

Vaney, D.I. (1984). Coronate Amacrine Cells in the Rabbit Retina Have the Starburst Dendritic 802 Morphology. Proc R Soc Ser B-Bio 220, 501-508. 803

Velte, T.J., and Masland, R.H. (1999). Action potentials in the dendrites of retinal ganglion cells. J 804 Neurophysiol 81, 1412-1417. 805

Vetter, P., Roth, A., and Hausser, M. (2001). Propagation of action potentials in dendrites depends on 806 dendritic morphology. J Neurophysiol 85, 926-937. 807

Wässle, H., Peichl, L., and Boycott, B.B. (1981). Morphology and Topography of on-Alpha and Off-808 Alpha Cells in the Cat Retina. Proc R Soc Ser B-Bio 212, 157-175. 809

Williams, S.R. (2004). Spatial compartmentalization and functional impact of conductance in 810 pyramidal neurons. Nat Neurosci 7, 961-967. 811

Wood, S.N. (2006). Generalized additive models: an introduction with R (Boca Raton, Florida, U. S. A.: 812 Chapman & Hall). 813

Yu, W.Q., El-Danaf, R.N., Okawa, H., Pacholec, J.M., Matti, U., Schwarz, K., Odermatt, B., Dunn, F.A., 814 Lagnado, L., Schmitz, F., et al. (2018). Synaptic Convergence Patterns onto Retinal Ganglion 815 Cells Are Preserved despite Topographic Variation in Pre- and Postsynaptic Territories. Cell 816 Rep 25, 2017-2026 e2013. 817

.CC-BY-NC-ND 4.0 International licenseavailable under awas not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made

The copyright holder for this preprint (whichthis version posted September 3, 2019. ; https://doi.org/10.1101/753335doi: bioRxiv preprint